Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2012 Oct 7.
Published in final edited form as: J Mol Biol. 2011 Mar 1;412(5):754–771. doi: 10.1016/j.jmb.2011.01.018

Mechanism of Bacterial Transcription Initiation: RNA Polymerase - Promoter Binding, Isomerization to Initiation-Competent Open Complexes, and Initiation of RNA Synthesis

Ruth M Saecker 1, M Thomas Record Jr 1,2, Pieter L deHaseth 3,4,*
PMCID: PMC3440003  NIHMSID: NIHMS276927  PMID: 21371479

Abstract

Initiation of RNA synthesis from DNA templates by RNA polymerase (RNAP) is a multi-step process, in which initial recognition of promoter DNA by RNAP triggers a series of conformational changes in both RNAP and promoter DNA. The bacterial RNAP functions as a molecular isomerization machine, using binding free energy to remodel the initial recognition complex, placing downstream duplex DNA in the active site cleft and then separating the nontemplate and template strands in the region surrounding the start site of RNA synthesis. In this initial unstable “open” complex the template strand appears correctly positioned in the active site. Subsequently, the nontemplate strand is repositioned and a clamp is assembled on duplex DNA downstream of the open region to form the highly stable open complex, RPo. The transcription initiation factor, σ70, plays critical roles in promoter recognition and RPo formation as well as in early steps of RNA synthesis.

Keywords: RNA polymerase, Promoter, Sigma factor, Transcription initiation, Intermediate complexes

Introduction

We begin this perspective with a brief overview of transcription initiation by bacterial RNA polymerase (RNAP), summarizing the players and the major steps in the process. Excellent review articles provide a more detailed coverage of many aspects of transcription initiation.19 Here we focus on current advances in understanding the process of isomerization of the initial closed complex to form the stable open complex RPo and the many crucial roles of the specificity subunit σ70 in all steps of initiation.

Initiation of RNA Synthesis in Bacteria

The essential players

The bacterial RNAP “core enzyme” (E) consists of five subunits, ββ′α2ω (see Fig. 1). The core enzyme is capable of nonspecific DNA binding and initiation of RNA synthesis from DNA ends or nicks, but requires a sigma factor to initiate specific transcription from promoter DNA. Sigma assembles with core to form the “holoenzyme” (or Eσ).18,19 Sigma factors recognize specific promoter DNA sequences, interact with transcription activators, participate in promoter DNA opening, and influence the early phases of transcription (e.g., Gruber and Gross;5 the latter two roles of sigma are further discussed in this review). The vast majority of studies of bacterial initiation have been carried out using Escherichia coli as model system. A model of the structure of the open complex formed by E. coli70 RNAP (shown in Fig. 1) highlights: (i) the positioning of σ70 on the core enzyme (Fig. 1a); (ii) the deep, wide cleft formed by β and β′ that binds the transcription bubble (Fig. 1b); and (iii) the flexible domains of β and β′ at the downstream end of the cleft proposed to assemble on the downstream duplex DNA to stabilize the open complex(es) (Fig. 1b).

Fig. 1.

Fig. 1

Model of the E. coli RNAP (σ70 α2ββ′ω) open complex RPo based on Protein Data Bank IDs 3IYD10 and 3LU0.11 (a) View of RPo illustrating the interactions between promoter DNA [nontemplate strand (NT), black; template (T), dark green] and σ2, σ3, and σ4 (wheat). Linker σ3.2 is buried in the RNA exit channel. The N-terminal domains of α (bright green, yellow) form a hinge at the bottom of the cleft. σNCD is a folded nonconserved domain connecting σ1.2 and σ2. ω is shown in light gray. Missing from the figure are σ1.1 and the flexibly tethered αCTD (not resolved in any holoenzyme structure to date). (b) View down into the active-site channel highlighting mobile regions on the periphery of the cleft and in the cleft. At the upstream entrance to the cleft, β′ clamp helices (black) tightly interact with σ1.2 and σ2. The open transcription bubble (–11 to +3 in this model) binds in the cleft, with the template strand start site (+1) next to the active site Mg2+ (red sphere) at the bottom. βSI1 (magenta) and β′SI3 (blue) are species-specific sequence insertions (SIs) present in E. coli. The remaining colored regions are highly conserved in bacteria.12 Along with βSI and β′SI3, β′ jaw (yellow) and β′ clamp (red) appear positioned to clamp on the downstream duplex DNA after the bubble has opened.1315 Flexible elements in the cleft that likely bind and stabilize the DNA single strands in RPo include the bridge helix (visible under the double-stranded–single-stranded boundary of the downstream DNA; pink), rudder (green), fork loop 2 (teal), and switch 2 (light blue). Other mobile elements shown are the β′ upclamp (hot pink; see Supplementary Fig. 1), which is proposed to interact with upstream DNA in forming I1 (the first kinetically-significant intermediate at the λPR promoter),16 and the trigger loop (orange), which is known to be critically involved in the RNA synthesis steps.17

All bacteria have a primary sigma factor that suffices for growth under nutrient-rich conditions. In E. coli, the primary sigma factor is σ70 (also called σD), reflecting its molecular mass of approximately 70,000 Da. In many other bacteria, the analogous primary sigma factor is designated σA. Most bacteria also have a complement of “alternative” or “minor” sigma factors (six in E. coli). Holoenzymes containing minor sigma factors recognize promoters of genes that can mitigate the effects of various adverse conditions.5,6 Most bacterial sigma factors exhibit significant homology to E. coli σ706 and, as such, belong the σ70 class. Others belong to the σ54 class due to their similarity to E. coli σ54 (also called σN, which is responsible for the expression of genes involved in nitrogen utilization), which has little sequence similarity with σ70.3 The evolution of these two distinct lineages of sigma factors is not understood.

The structure of σ70 is shown in Fig. 1a. The four regions of sequence conservation common to the σ70 class sigma factors20 and the architecture of promoter DNA sequences that they recognize are shown in Fig. 2. (Regions of σ70 are designated in this review as subscripts; i.e., σ2 refers to region 2 of σ70.) In addition to the –10 and –35 hexameric recognition sequences22 (Fig. 2), σ70 factors also recognize a TG sequence upstream of –10 (together called the extended –10)2326 and guanines in the discriminator region (see the text below) at –6 and –5.27,28 The spacer length (i.e., the number of base pairs separating the – 10 and – 35 elements, optimally 17 bp) and the number of base pairs separating the –10 element from the transcription start site (optimally 7 bp)2931 both modulate the interactions of Eσ70 with the promoter. Some promoters also include ~20 bp of A/T-rich sequence upstream of the –35 element, referred to as an “UP element” (see Fig. 2). The UP element is recognized by the flexibly tethered α-subunit C-terminal domain (αCTD).21 The αCTD also can bind nonspecifically to upstream DNA,3235 making contacts up to ~–90.

Fig. 2.

Fig. 2

Promoter recognition by amino acids of the α subunit and σ70. Orange and blue arrows indicate recognition of promoter regions as double-stranded DNA elements by the α and σ70 subunits, respectively. The two red arrows delineate a region of the nontemplate strand DNA recognized by σ70 subsequent to strand separation. In the linear representations (not drawn to scale) of both σ70 and α, the N-termini are on the right. Only the sequence of the nontemplate strand is shown (5′ end on the left). A typical E. coli promoter does not have all elements shown and exhibits deviations from the consensus sequences shown here for the –10, –35, and UP21 elements, as well as the consensus spacer length (17 bp).

Steps of transcription initiation

Specific binding of Eσ70 RNAP to promoter DNA, forming an initial closed complex RPc, triggers a series of conformational changes in both biomolecules. This series of events, often collectively called “isomerization,” opens ~13 bp from the –10 element to just beyond the transcription start site, creating the initiation “bubble” and an unstable open complex.36 In this step or in subsequent steps of forming the final stable open complex (RPo), the +1 template (T) strand base is placed in the active site of the polymerase, and the nontemplate (NT) strand is placed in its binding track. RPo is stabilized by the assembly and DNA binding of a downstream jaw/clamp,1315,36 which presumably is important for processive transcription (see Fig. 1b and the text below).

During isomerization, contacts between σ2 and the duplex form of the –10 region in the closed complex are replaced by interactions between conserved aromatic residues in σ2 and NT strand bases from –11 to –7 during or after DNA opening (see the text below). Work to date indicates that the –10 element (with the exception of –12, which remains base-paired) is primarily recognized as single-stranded DNA.37 Open bases at positions –6 and –5 on the NT strand (the discriminator region; see Fig. 2) interact with σ1.2,27,28 as judged by cross-linking experiments.38 Base identity at these positions has very large effects on the rate of dissociation of the open complex at the ribosomal rrnB P1 promoter, but only small effects on the binding and isomerization steps that determine the association kinetics.28 Thus, bases on the NT strand from –11 to –5 appear to be largely recognized in the single-stranded state28 after the opening of the initiation bubble. Importantly, no such recognition occurs on the T strand. The difference in interactions with the T strand versus the NT strand has consequences for the later steps of NTP addition and promoter escape. σ3 (also called σ2.5) interacts with the extended –10 TG sequences;23 it appears to also play a role in the steps after promoter binding (Fig. 2). Although the extended –10 element remains duplex throughout initiation, changes in this sequence primarily affect the rate of isomerization and not closed complex formation.39,40

Templated RNA synthesis (transcription) involves covalent bond formation between the 3′ OH end of the nascent RNA and the α phosphate of the incoming NTP (nucleoside triphosphate). Phosphodiester bond synthesis results in the extension of the chain by one residue and in the release of pyrophosphate. At most Eσ70 promoters, transcription is initiated with ATP or GTP, but promoters at which initiation occurs with CTP or UTP have also been characterized. The initial transcribing complex may go down either productive or abortive paths. As a short RNA chain is synthesized, contacts between the RNAP and the promoter DNA upstream of –11 remain intact, while the promoter DNA is progressively “scrunched”41,42 and the transcription bubble is expanded43 as additional DNA is pulled in and copied into RNA. This process builds up stress and sets up a competition between extending the RNA chain and increasing the size of the DNA bubble, or releasing both the small product and the stress in the scrunched DNA to revert to RPo.41,42 When the nascent RNA reaches a critical length of about 11 nucleotides, the stress is instead relieved by disruption of the contacts between the RNAP and the promoter DNA.

The number and the length of abortive products produced prior to productive initiation are a function of promoter sequence and conditions,44,45 but the precise “rules” governing this behavior remain unknown. For the single subunit phage T7 RNAP, single-molecule and fluorescence studies demonstrate that the probability of transition from an initiating complex to an elongation complex strongly depends on RNA length.46 Abortive initiation was once thought to be an in vitro artifact or an inconsequential aspect of promoter escape. However, abortive RNAs have now been detected in vivo.47 This discovery raises the intriguing possibility that the small products (e.g., 2–4 mers) may rebind the open complex and thus prime initiation in vivo and alter gene expression in a concentration-dependent fashion.47

σ70 is not required for elongation and is typically released from the transcription complex when the RNA reaches a length of 12–15 nt.48 Release of σ70 is likely triggered by events set in motion when the nascent RNA–DNA hybrid reaches 8–9 bp. Further extension requires displacing the sigma linker connecting σ2 and σ4 that lies in the RNA exit channel.49,50 While the competition with the growing RNA chain is thought to release σ3.2 and σ4, it is unclear how the remaining interaction between σ2 and β′ clamp helices51 is disrupted. Indeed, when σ70 is retained, σ2 induces promoter-proximal pausing at promoters with a –10-like sequence in the NT strand downstream of the start site.5255 Under some conditions, σ70 release is delayed beyond the transition from initiation to productive elongation. The events governing σ70 release versus retention remain to be defined. They are likely regulated and thus motivate ongoing investigations.

X-ray and electron microscopy structures of Eσ RNAP: Implications for transcription initiation

X-ray structural data for core,56A,50,57 and several nucleic acid–thermophilic RNAP complexes17,49 have had considerable impact and influence on the understanding of bacterial transcription initiation. In common with other nucleic acid polymerases (but on a larger scale), the active site lies at the bottom of a deep cleft (~70 Å deep and >100 Å long; see Fig. 1b). In addition, the EσA structures detail the extensive interface formed between the highly conserved regions of sigma and core, as first deduced by biochemical and genetic studies.58

The highly conserved multisubunit RNAP architecture12 itself appears to play a key role in discriminating promoter DNA from nonpromoter DNA during initiation. First, the arrangement of sigma on core and, in particular, the positioning of the promoter-recognition regions of sigma relative to the active-site cleft create a series of obstacles for the DNA to overcome to form an open complex. Interactions of σ2 and σ4 with the –10 and –35 elements (Fig. 2) define a promoter DNA trajectory in the initial “short footprint” closed complex RPc (see the text below), which is at 90° with respect to the cleft. Consequently, a sharp bend at –1159 and/or DNA opening outside the cleft50,60,61 must be introduced for DNA to enter the cleft.

Does promoter DNA containing the start site of transcription (+1) enter the cleft as separated single strands or as a double helix? The ongoing debate regarding this question is driven, in part, by the narrow width (>25 Å) of the cleft seen in the EσA crystal structures. This observation motivates the hypothesis that the cleft “screens” the state of the DNA by only allowing single-stranded, but not duplex, DNA entry.50,60 However, this narrow width appears to be a snapshot of just one conformational state. In general, structures of the bacterial RNAP and eukaryotic RNAP II in various states of ligation and/or crystal forms exhibit a range of cleft widths, with distances (between the β′ clamp and the β pincer lobes) varying from <25 Å (e.g., “open”) to >15 Å (“closed”) (Mukhopadhyay et al.62 and references therein). Recent single-molecule fluorescence resonance energy transfer (FRET) experiments confirm that clamp opening and closing occur, demonstrating that the “hinges” in β and β′ at the base of the cleft are flexible in solution (A. Chakraborty and R. Ebright, personal communication).

Comparison of RNAP structures indicates that cleft width is controlled, in part, by the conformation of “switch 2” at the base of β′ (see Fig. 1b) (Mukhopadhyay et al.62 and references therein). Recent work has demonstrated that the bacterial transcription initiation inhibitors myxopyronin (myxo), corallopyronin, ripostatin, and lipiarmycin (lpm) target switch 2.6265 Although RNAP–myxo crystal structures indicate that myxo binding stabilizes a “partly closed” conformation of the clamp,62,63 footprinting data on myxo–RNAP–promoter DNA complexes suggest that myxo inhibits melting of the start site region (–2 to +2) but does not prevent the entry of duplex DNA into the cleft.63 Like myxo ternary complexes, lpm–RNAP–promoter DNA complexes are protected downstream to at least +15 from DNase I cleavage.65 However, unlike myxo–RNAP– promoter DNA complexes, no permanganate-reactive thymines are detected in the presence of lpm.65 These data, along with results summarized below, argue that cleft width does not preclude duplex DNA entry during formation of RPo.

An additional block to forming the open complex is created by the acidic N-terminal domain of σ70. The single-stranded nucleic acid mimic σ1.1 binds in the cleft, blocking access to the active site.66,67 For proper orientation of the start site base with respect to the active site Mg2+, σ1.1 must be repositioned and the T strand must descend ~70 Å from its location in the closed complex (see the text below). After NTP binding and short RNA synthesis, the transition from initiation to elongation (promoter escape) requires displacing the flexible linker (σ3.2) connecting σ2 and σ4 from the RNA exit channel,50,57 and breaking the contacts of σ2 and σ4 with the –10 and –35 elements of promoter DNA, respectively.

Although high-resolution structures are not available for E. coli RNAP, a recent 20-Å electron microscopy structure of a ternary complex (E. coli holoenzyme-CRP-DNA)10 and a complete model of the E. coli core enzyme11 reveal the locations of three large sequence insertions (SIs) in the E. coli β and β′ subunits that are absent in the thermophilic RNAP.12 Two of these insertions, βSI1 and β′SI3, lie at the downstream end of the cleft (see Fig. 1b). β′SI3 occupies a particularly prominent position: it forms a tethered independent domain with the highly conserved β′ “jaw.” The β′ jaw/SI3 domain is highly mobile68 and likely provides an additional steric “block” to prevent nonpromoter duplex DNA from accessing the active site.16

Steps in RPc-to-RPo Isomerization

Mechanistic studies

How is the start site DNA opened, placed in the active site, and stabilized? During RPo formation, how and when are the obstacles that prevent nonpromoter DNA from accessing the cleft and being opened overcome? For several decades, kinetic mechanistic and footprinting studies have been employed to determine the sequence of conformational changes and the nature of intermediate complexes on the pathway from the initial promoter-recognition complex RPc to RPo. At the lac UV5 and λPR promoters, at least two steps are required to convert the initial closed complex into the final stable open complex RPo.6971 However, the “isomerization” intermediates that separate the closed complex and RPo are relatively unstable and short lived (<1 ms to 1 s; see Fig. 3). To date, they have resisted characterization by crystallography, cross-linking, FRET, and single-molecule approaches. Their size currently precludes NMR characterization.

Fig. 3.

Fig. 3

Summary of the proposed isomerization steps that form the initiating complex (RPinit) after recruitment of RNAP to form an initial complex at the promoter (RPc). Formation of the closed complex RPc triggers a series of subsequent large-scale conformational changes. The RNAP molecular machine places start-site duplex DNA in the active-site cleft in I1, opens it to form I2, and stabilizes the open form by assembling a clamp in I3 and RPo (model based on Gries et al.,36 Kontur et al.,13,15 Davis et al.,16 and Craig et al.72). Once promoter DNA is open, NTPs can bind, and transcription initiates. I2 and I3 are open complexes; current studies are addressing whether they can bind NTPs and initiate transcription.

While methods for characterizing transient isomerization intermediates were being developed, attention was focused on initial promoter recognition (forming RPc) and its regulation by promoter sequence and by activator or repressor proteins. Information about RPc and other potential intermediates has been obtained by equilibrium footprinting either at low temperatures (0–15 °C) (e.g., Kovacic,73 Cowing et al.,74 and Schickor et al.75) or with variant RNAPs40 unable to effect promoter DNA melting. These complexes, all closed, exhibit different hydroxyl radical (·OH) or DNase footprint boundaries at different temperatures or promoters (see the text below) and have been given different names.59,69,70,72,7678 Because of the challenges in characterizing kinetically significant but unstable intermediates in real time, the mechanism of forming RPo has often been condensed into two steps: R +P→RPc→RPo. This mechanism collapses all of isomerization, including DNA opening and placement of the start site base of the T strand in the active site, into a single step.

In the association direction of the three-step mechanism for the lac UV5 and λPR promoters,70,71 the first kinetically-significant intermediate (designated I1 at the λPR promoter) is found to be closed16. I1 is more “advanced” than RPc, protecting DNA to +20. While RPc likely forms first, it never accumulates at the λPR promoter. I1 isomerizes in the rate-determining step to a second intermediate (designated I2 at the λPR promoter and found to be open36), which rapidly converts to RPo. In the dissociation direction, the reverse direction of this same step (I2→I1) is rate determining. Thus, the I1→I2 and I2→ I1 steps are the bottleneck steps in each direction. In the forward direction, use of high concentrations of RNAP creates a “burst” in the population of the closed complex immediately preceding this rate-limiting isomerization step.7779 To create bursts of intermediates formed after I1, RPo is destabilized by using salts and solutes that do not destabilize these intermediates. To trap the elusive second intermediate, temperature downshifts were attempted.70,72,80,81 However, because I2 is an open complex, this approach also destabilized it rather than leading to its accumulation.14,36

DNA footprinting studies of intermediate complexes

Methods for investigating late intermediates of isomerization have been developed recently.14,36 These methods, combined with rapid-quench (<2 ms) mixing, allow one to perform “real-time” kinetic and chemical footprinting experiments on the timescale of the formation and disappearance of transient intermediates.36,77,78 To date, all structural information about complexes known to be on-pathway intermediates in RPo formation has come from chemical and enzymatic DNA footprinting methods.

Recent advances in describing the large-scale conformational changes that occur after recruitment of RNAP to the promoter and initial specific binding are summarized in Fig. 3. In addition to opening of the promoter DNA, strong evidence is obtained for DNA wrapping and for coupled folding and domain repositioning of RNAP. Because all of these conformational changes are driven by binding free energy, the motions in the RNAP machinery are linked to the DNA sequence and structure in the interfaces that form, as well as to solution conditions. Below we detail our current understanding of the progression of conformational changes, the structures and stabilities of the intermediates, and the controversies and questions that remain.

RPc formation: Initial recognition of duplex promoter DNA sequences

Initial specific interactions of RNAP with promoter DNA form a closed complex in which the promoter DNA is fully duplex. Based on the low-temperature footprinting data (see the text above) and the structures for holoenzyme50,57 and for a complex of RNAP with a “fork junction” promoter fragment,49 a model of such an initial closed complex has been proposed. In this model, duplex promoter DNA interacts with σ2 and σ4, and a continuous DNA duplex extends downstream of the –10 element, projecting away from the active-site cleft and therefore is cleavable by DNase I or ·OH.60,82

Real-time ·OH footprinting after mixing high concentrations of RNAP and T7A1 promoter DNA shows progression in the protection of the downstream boundary after mixing.77,78 In these studies, a series of closed intermediates, which initially establish protection between –80 and –55 and then progressively extend it downstream, has been proposed. These snapshots suggest that interactions involving αCTDs and the T7A1 UP element and/or other far-upstream DNA contacts with RNAP are established first (see also Borukhov and Nudler8). The footprint then extends downstream as contacts between σ2, σ3, and σ4 and the –10 element, and between the spacer and the –35 element of the promoter DNA, respectively, are established to form a RPc complex with a downstream boundary of –5.77,78

More advanced closed complexes, including the upstream-wrapped closed intermediate I1 with downstream duplex in the active-site cleft

Downstream boundary and its implications for closed complexes

More “advanced” closed complexes with downstream boundaries extending to +15–20 have also been characterized.16,72,74,80,8386 Conversion from an “RPc”-like complex into one that protects one to two turns of the DNA helix downstream of the start site (+1) has been observed by increasing the temperature to 16–20 °C74,80,83 or, in the case of the rrnB P1 promoter, by increasing the temperature to 37 °C in the absence of NTPs.84,86 In all cases, chemical probes (dimethyl sulfate and permanganate) do not detect open/unstacked bases in these “advanced” complexes, and the periodic patterns of protection observed from ~–55 to –12 transition to full protection of both strands of the helix between –11 and +15.

At the λPR promoter, both the transient, kinetically significant intermediate closed complex (I1) that accumulates early in the time course of open complex formation16 and the low-temperature (0 °C) closed complex,87 also known to be I1 by extrapolation of thermodynamic data from 7 °C and higher temperatures, have an extended downstream footprint (to +20–25). Based on the holoenzyme structures, we proposed that protection of both strands from –11 to +15 results from a sharp ~90° bend at the upstream end of the –10 element that inserts downstream duplex DNA into the active-site cleft.59 Additional protection to +20 and +25 likely arises from mobile elements at the downstream end of the cleft that block access to the DNA backbone (e.g., β′ SI3; see Fig. 1b). Conversion of RPc to a more I1-like complex appears to be driven by increasing temperature, by favoring the bend at –11/–12 and/or the interactions that stabilize the bend. At λPR, I1 is the most advanced closed complex because it opens in the next kinetic step.

In closed complexes with downstream boundaries between –5 and +15,33,84,86 downstream DNA is presumably only partially inserted into the cleft. Is this because of the difficulty in bending the DNA or the difficulty in inserting the duplex in the cleft? Possible examples of both scenarios are available. Davis et al. found that the I1 intermediate formed by RNAP at an upstream-truncated (UT-47) λPR promoter exhibits a downstream boundary of +2(NT)/ +7(T).16 Since the interactions of the –10 element are presumably identical in UT-47 and full-length λPR, the difference in the insertion of the downstream duplex in the cleft may therefore indicate an obstacle in the cleft that is removed as a result of interactions with far upstream DNA (above –47).

Upstream boundary and its implications for closed complexes

Interactions with DNA upstream of the –35 hexamer are established in the steps of promoter recognition.16,77,78 They influence the stability of the intermediate I1 and the rate of converting it to the next intermediate I2. The αCTDs (see the text above) mediate upstream interactions by binding DNA either specifically21,88 or nonspecifically3235 and by interacting with activator proteins.1,89 Intriguingly, all proteins (including the αCTDs, based on DNase I enhancements) that bind upstream of the –35 hexamer and modulate transcription bend DNA [e.g., CRP (cyclic AMP receptor protein), IHF (integration host factor), and FIS (factor for inversion stimulation)]. The function of these DNA bends may be simply to provide better interactions with the activator and/or the αCTDs. However, many transcription factors bind to sites upstream of –90, presumably out of “reach” of the flexibly tethered αCTDs. In addition, the phasing of upstream binding sites for these factors (e.g., see Dethiollaz et al.90 and Giladi et al.91) affects their action. Shifting the UP element in the upstream direction relative to the –35 hexamer of a given promoter abolishes UP element activation of transcription, regardless of phasing; lengthening the αCTD–αN-terminal domain linker does not restore full-length transcripts to their nondisplaced UP element levels.92 Moving the UP element upstream prevents the formation of a complementary interface between σ4 and the adjacent (proximal) αCTD.93,94 Likewise, shifting sites for transcription factors that naturally abut the –35 hexamer upstream destroys the interface that they form with σ4.95,96 In the wild-type context, formation of these complementary interfaces is a key event in transcription initiation.

The above data have engendered several hypotheses about the role of upstream interactions. One hypothesis is that protein–protein interfaces communicate “allosterically” with the active-site channel to affect steps in initiation. Alternatively, or in addition to possible allosteric effects, we have proposed that the network of interactions between σ4 and the αCTDs (and transcription factors, when present) and DNA bends the DNA from ~–30 to –55 and thereby sets the trajectory of far upstream DNA in the early steps of RPo formation16 (see Supplementary Fig. 1). The importance of the upstream DNA trajectory is based on several observations. First, the presence of DNA upstream of the –35 element at the lac UV5 and λPR promoters profoundly accelerates the bottleneck isomerization step (see Fig. 3), now established as the DNA opening step at λPR.36 The isomerization rate constant k2 (conversion of I1 into I2) for full-length λPR is ~50-fold larger than that for truncated λPR, with the upstream DNA deleted beyond –47 (UT-47), an effect as large as or larger than that exerted by activator binding. Surprisingly, deletion of this upstream DNA has little effect to no effect on the stability of I1. 33,34 Second, deletion of upstream DNA leads to a “less advanced” closed complex that only protects downstream DNA to +2 (T)/+7 (NT)16 relative to ~+20 observed for the full-length λPR promoter.

How might upstream interactions facilitate the loading of downstream DNA in the cleft? In the final stable open complex at the λPR promoter, the upstream boundary defined by DNase I or ·OH cleavage ends at ~–65. However, ·OH footprints of I1 on full-length λPR reveal modest protection of the DNA backbone on both strands to at least –85.16 DNase I hypersensitive sites in I116,72 indicate that a bend occurs just upstream of the –35 element. Mapping the I1 protection pattern and the inferred bend onto available X-ray structures indicates most simply that RNAP wraps DNA around the “back” of the β′ subunit (Supplementary Fig. 1). Alternatively, the upstream interactions responsible for the large effect on the kinetics of the DNA opening step discussed above and for the far upstream ·OH footprint could involve the mobile αCTDs (see the text below).

In the upstream wrapping model,16 the bend induced by σ4 and the two αCTDs places far upstream DNA near the downstream end of the cleft (see Supplementary Fig. 1). Because the pattern of protection from –65 to –85 is not periodic (as typically observed for the αCTDs97), we proposed that this region of DNA is directed into a surface groove formed by β′ and the N-terminal domain of the associated α subunit. If so positioned, the upstream DNA lies near a conserved mobile element in β′, termed the upstream clamp. The upstream clamp is directly connected to other dynamic elements at the downstream end of the cleft: conserved jaw, trigger loop, and β′SI3. Based on our kinetic and footprinting data,16,33 we hypothesize that interactions between upstream DNA and upstream clamp restrain the movements of the jaw domain, trigger loop, and β′SI3. Without this constraint, it appears that these elements sterically interfere with the loading of the downstream DNA in the cleft (see Fig. 1b; Supplementary Fig. 1). Predictions of this model are currently being tested.

Alternative hypotheses for the role of upstream DNA invoke a direct role for the αCTDs in mediating the acceleration of the DNA melting step beyond bending upstream DNA. In the absence of the αCTDs, the presence of upstream DNA only increases k2 by ~2.5-fold.34 Cross-linking of RPo indicates that the αCTDs can occupy multiple sites on the upstream DNA.35 Discerning whether the role of upstream DNA is to simply provide additional nonspecific αCTD binding sites or whether the αCTDs and σ4 together set a trajectory required for wrapping interactions between upstream DNA and other elements of RNAP in I1 or other early complexes awaits further experiments.

I1→I2: DNA opening is the bottleneck step in RPo formation at the λPR promoter

How is DNA opened by RNAP? Two conflicting hypotheses describing this critical step exist in the field. One hypothesis, based on structural data, posits that opening is nucleated by DNA breathing above the active-site cleft, after which the T strand enters the cleft and diffuses to the active site.50,60 Evidence in support of this proposal has been obtained from molecular dynamics simulations on modeled structures formed by the bacterial RNAP61 and from a comparison of the time evolution of downstream ·OH and MnO4- footprints in association experiments at the T7A1 promoter.77 A second hypothesis proposes that the DNA duplex is first loaded in the cleft, where it is then actively opened by elements on RNAP.15,36 This proposal is supported by extensive kinetic and footprinting (equilibrium and real time) experiments on the λPR promoter (cf. Fig. 3), equilibrium footprinting experiments at other promoters (see the text above), and equilibrium footprinting experiments performed in the presence of antibiotics that block DNA melting (discussed above).

Closed promoter DNA–RNAP complexes that protect the DNA backbone to at least +15 demonstrate that duplex DNA can bind in the active-site cleft. However, in most cases, evidence that complexes populated at equilibrium were on-pathway kinetic intermediates was not obtained. However, extensive kinetic studies (filter binding) of RPo formation at λPR, combined with real-time footprinting experiments, provide strong evidence that duplex DNA (–11 to +20) occupies the active-site cleft in the final closed on-pathway intermediate I1. Once bound in the cleft, the next step (I1→I2) opens DNA (–11 to +2), as detailed below.

At λPR, DNA in I1 is not MnO4- reactive and is continuously protected from ·OH and DNase I on both strands from –11 to positions +20–25.16,72 Thymines in the subsequent kinetically significant intermediate I2 are MnO4- reactive at all positions reactive in RPo.36 In addition, real-time footprinting experiments reveal that the extended downstream footprint (protected from ·OH cleavage to ~+20) of I1 develops in 100 ms. In contrast, the MnO4- reactivities of thymines detected in RPo develop much more slowly (tens of seconds; Heitkamp, Drennan, et al., in preparation). Therefore, we conclude that duplex DNA binds in the cleft in I1, and that the entire bubble opens concertedly in the cleft at λPR in one kinetic step.15,16,36,59

In addition to evidence cited directly above, the following data also indicate that DNA opens in the cleft. The rate constant for I1→ I2 is strongly temperature dependent.59 The corresponding activation energy (34 kcal) is consistent with the cooperative opening of at least 6–7 bp in the I1–I2 transition state. While salt and other solutes exhibit large effects on DNA opening in solution, they only exert small effects on the DNA opening (k2) and closing (k–2) steps for the λPR promoter.1315,36,59 Most simply, these data indicate that DNA opening occurs in the sequestered environment of the cleft and not outside it. Alternatively, compensating for the stabilizing and destabilizing effects of these salts and other solutes may accompany opening.

I2→RPo: Evidence for the assembly and DNA binding of a downstream clamp/jaw to stabilize the open complex

Evidence to date indicates that the final steps of isomerization involve the interconversion of multiple different open complexes, including on-pathway intermediates I2 and I3, as well as the final open complex RPo. This striking discovery has significant implications for the regulation of transcription initiation.14,36 Major changes in DNA and RNAP in the conversion of I2 into RPo were revealed for the first time by burst footprinting of the dissociation intermediate I2 and by analysis of dissociation data as a function of both stabilizing and destabilizing solutes and salts.1315,36 These include downstream folding and assembly of >100 residues of mobile elements of RNAP (Figs. 1 and 3; Supplementary Fig. 1) to form a clamp/jaw on downstream DNA, as well as establishment of in-cleft interactions. Evidence for the latter includes a 2-fold increase in the MnO4- reactivity of thymine bases in the downstream region of the NT strand (–4, –3, and +2) in the conversion of I2 into RPo. Thymine bases in the upstream half of the NT strand remain protected from MnO4- oxidation by being bound to RNAP (σ2) in both I2 and RPo. The start site thymine (+1, T strand) is equally MnO4- reactive in I2 and RPo, suggesting that it is correctly positioned in I2.

Downstream interactions in the open complexes

Numerous lines of evidence demonstrate that RNAP undergoes a large-scale conformational change in the steps following DNA melting. The extreme effects of solutes, temperature, and salts on the steps converting I2 into RPo suggest that the late steps of RPo formation create a new protein–DNA interface in a process that involves coupled folding.98 Quantitative analysis of the effects of multiple solute probes on the dissociation rate constant Kd indicates that 75–100 residues fold in the conversion of I2 to RPo; the effects of salt are consistent with a burial of 10 or more DNA phosphates.15 The conformational changes that occur in these steps appear to be comparable in scale to those that occur in the conversion of the initiation complex to the elongation complex of the T7 phage RNAP (~300 amino acids refold99).

To interpret these results, we proposed1315 that several large (50–70 residue) mobile regions of the β′ and β subunits fold and assemble into a jaw/ clamp superstructure that binds to duplex DNA (see Fig. 1b; Supplementary Fig. 1) after DNA opening. Given the tight binding interactions established upstream in the early steps, assembly of the downstream clamp is likely delayed to allow unimpeded rotation of the downstream DNA on its helical axis by 1.3 turns (470°).1315 Regions in β′ include: (i) the jaw; (ii) a highly positively charged helix hairpin helix; (iii) β′SI3; and (iv) a C-terminal region adjacent to (i). Individual deletions of (i)–(iii) all destabilize RPo;13,84,100 deletions/mutations in (iv) have not been studied. In addition, βSI1 also likely forms part of the clamp.14 In the recent model of RPo10 (as well as in the complete model of an E. coli transcription elongation complex11), all of the regions above are positioned near the downstream DNA, but do not necessarily directly interact with it. This may be a consequence of the low resolution of DNA in electron microscopy images and of basing the model on the transcription elongation complex, which only protects downstream DNA to ~+10–15101 (~1 turn of DNA shorter than RPo). Nonetheless, these models clearly show that β′SI3, the jaw, and βSI1 are positioned to clamp downstream DNA from +10 to +20.10,11

In both the RPo and the transcription elongation complex models, the trigger loop, which connects to β′SI3 through flexible linkers, is unfolded. Intriguingly, using the folded form of the trigger loop17 creates steric clash, leading to the proposal that the jaw/β′SI3 domain likely toggles between two positions as the trigger loop folds and unfolds with each cycle of NTP addition.10,11 In addition, the RPo model shows density for βSI1 near +15–20. The volume of this density increases relative to the partially disordered state in free E. coli RNAP,68 supporting our hypothesis that βSI1 folds on binding downstream DNA.13

In-cleft interactions in the open complexes

Are these large conformational changes in the RNAP downstream machinery during the conversion of I2 to RPo connected/correlated with smaller-scale but very significant conformational changes in the active site and surrounding regions of the cleft? Differences in the MnO4- reactivities of bases on the downstream portion of the NT strand in I2 and RPo suggest that rearrangements in the NT strand are coupled to the formation of the downstream DNA clamp in the conversion of I2 to RPo.1315,59,102 Details of the communication between the cleft and the clamp/jaw remain to be established. This communication, if established in the isomerization steps, would likely persist in regulating the subsequent steps of the transcription cycle. Does the sequence/length of the discriminator region (–6 to +1) affect the “repositioning” of the NT strand and, thus, the stability of RPo at different promoters? Are I2 and I3 functional in transcription? Do the different open complexes (one unstable and one stable) play distinct functional roles of open complex formation (e.g., Which open complex is the target of the stress sensor protein DksA?85)? How do the contacts established in the late steps of RPo formation affect promoter escape? Addressing these questions will likely bring new challenges and surprises, and will advance our understanding of the regulation of these late steps as a function of promoter sequence and solution conditions.

The role of σ70 in promoter interactions, open complex formation, and early RNA synthesis

Promoter recognition

Structures of RNAP holoenzyme from Thermus aquaticus and Thermus thermophilus reveal that σ70 consists of several independently folded domains (σ1.2, σ2, σ4, and likely the N-terminal ~60 residues of σ1.1 as well103) connected by flexible linkers (σ3.2 and the highly negatively charged C-terminal residues of σ1.1). Recent evidence reveals that the structure of the free sigma factors is compact, and that σ1.1 and σ4 interact.103 This interaction may lead to the observed autoinhibition of promoter DNA binding by free primary sigma factors.104 Autoinhibition has not been observed for interaction with the double-stranded promoter DNA of free sigma factors lacking region 1.1 by deletion105 or naturally,106,107 or for free σ70 interacting with the NT strand of promoter DNA.108

In the holoenzyme, the interactions between sigma and the core cover an extended surface area of both proteins58 (see Fig. 1a): various regions of σ70, including the linkers, interact with core RNAP, thus affording the bound σ70 considerable structural rigidity compared to the free σ70. For example, σ2.3/2.4 (–10 recognition) and σ4.2 (–35 recognition) are now at a fixed orientation with respect to each other, imposing a rather strict limitation on the length of the spacer DNA separating the –10 and –35 elements (17±1 bp). Indeed, there are conditions where only σ2 of sigma is bound to RNAP in a paused transcription complex where the RNA is 16–17 nt—long enough to have displaced σ4 from the β flap. At this point, –10-like and –35-like elements that occur just downstream and upstream of the start site of transcription, respectively, can be jointly contacted by the now flexibly tethered σ2 and σ4 even if they are separated by only 1 bp.109

Redundancy in σ70 promoter elements

The –10 and –35 elements together constitute the classical prokaryotic promoter. However, in addition to the –10 and –35 elements mentioned above, there are several other regions of promoter DNA contacted by σ70 or the α subunit (see the text above). This raises the question of whether other pairs of promoter elements can also constitute an active promoter. Promoter DNA melting initiates within the –10 element, rendering this the most important and the least dispensable of the promoter elements. Can other regions substitute for the –35 element? This has indeed been found to be the case (reviewed by Hook-Barnard and Hinton7). Notable among these are the extended –10 (TG+–10) and UP+–10, which has so far been only characterized as an artificial construct. Under in vitro conditions, promoter DNA strand separation has been observed with DNA containing just the –10 element.110 Even the combination of –35+TG has been found to be active, if provided with an A+T-rich region that has the all important –11A and –7T in the NT strand.110

Stringent promoter requirements for holoenzymes containing “minor” sigma factors

In addition to the primary (“housekeeping”) sigma factor, most bacteria have one or more minor sigma factors that can impart to RNAP the ability to transcribe genes whose products allow cells to deal with various types of stress. An example is the heat shock sigma factor σH (usually called σ32),111,112 which helps to mitigate the cytoplasmic consequences of transient exposure to higher-than-optimal temperatures. Promoter recognition occurs through a –35 element similar to that of σ70 promoters and an extended –10 element of which the sequence deviates considerably from that recognized by Eσ70.113 Importantly, while Eσ70 is relaxed in its ability to recognize promoter DNA sequences, Eσ32 is found to be stringent in requiring promoters with consensus or near-consensus sequences. The reason for the difference is the DNA melting region (2.3) of σ70: a broader promoter-recognition spectrum for Eσ32 is generated by replacing just two residues of region 2.3 with homologous aromatic σ70 residues.114 These experiments indicate not only that σ32 is intrinsically melting deficient but also that this deficiency can be overcome by the use of consensus promoters. Indeed, similar behavior has also been observed for σ70 mutants rendered melting deficient by substitutions in region 2.3: if provided with a very good promoter, RNAP containing a defective σ70 could still form an open complex.79 Observations similar to those described above for σ32 have also been made for RNAP containing σ28, another minor E. coli sigma factor.114

Nucleation of DNA melting: Role of conserved residues in σ70

The –11A element in the NT strand of E. coli σ70 promoters plays an important role in the formation of an open complex, as detailed in a number of studies.76,115119 It is the most conserved base pair in the –10 element.120 Substitution of the –11A element by 2-AP117 or loss of the base at this position121,122 has much larger negative effects than at other positions. At the upstream end of the –10 region, the A-T base pair at –12 remains double-stranded in RPo. A clear indication that strand separation is initiated at the –11 position is derived from studies demonstrating a correlation between the reduced stability of base pairing at –11 bp and the ability of promoter DNA to be melted by RNAP.116

Residues Y421, Y425, F427, T429, Y430, Y425, W433, and W434 (E. coli σ70 numbering; see Fig. 4) are nearly invariant among 53 sigma factors analyzed124 and are found within a short distance of each other and of the –11A base49 (at the upstream single-stranded–double-stranded DNA boundary over σ2 in Fig. 1a). T429, Y430, and W433 are near the double-stranded–single-stranded junction of the model DNA cocrystallized with the RNAP, consistent with their involvement in the initiation of DNA melting. Evidence for the vital roles of Y430 and W433 includes deleterious effects of substitutions on open complex formation125127 and their high extent of conservation. Compelling evidence has been obtained for an interaction of Y430 with –11A.79 However, it is likely that Y430 (and W433) additionally also recognizes other bases, and that other amino acid residues recognize –11A. Support for such a network of interactions, with multiple roles for the participating groups, is derived from two sets of observations. First, the effects of various substitutions for Y430 and W433 on the ability of RNAP to form stable promoter complexes are evident even in the absence of the base at –11 of the NT strand.122 Second, a variant RNAP containing a multiply substituted σ70 (alanine substitutions for F427, Y430, W433, and W434) has retained the capability for sequence recognition at –11.40

Fig. 4.

Fig. 4

Structure of region 2.3 of σ70.123 In the N→C direction are helix 13 (lower helix), a loop, and helix 14 (upper helix). The side chains of K414, K418, Y425, T429, Y430, W433, W434, and Q437 (Lys, green; Tyr, red; Thr, blue; Trp, purple; Gln, pink) stick out towards the viewer from approximately the same face of the protein, where they can interact with promoter DNA. Y430 has been shown to stack with –11A of the –10 region.79 Y421 sticks out in another direction but may be able to interact with DNA. The structures of the T. aquaticus and T. thermophilus σ2.3 are very similar.50,57

It is envisioned that interactions with basic amino acid residues of σ2 (including K414 and K418) anchor the promoter DNA to the surface of the RNAP in the closed complex.128 Promoter DNA melting is likely initiated by the rotation (or “flipping”) of –11A out of the DNA helix76,129 so that it now can stack onto Y430. Aromatic amino acid residues T429122 and W433125,128 of σ2.3 are likely closely involved in the actual process of flipping –11A out of the DNA helix.79 From the flipped –11A, DNA strand opening would proceed in downstream direction to +2.

Melting 12–14 bp of duplex DNA at 25–37 °C in the absence of RNAP has a large enthalpic cost (~70–84 kcal), thought to arise primarily from base unstacking and not from breaking of hydrogen bonds.130 However, the activation enthalpy for opening the transcription bubble at λPR is approximately half as large (~34 kcal59). Preservation of intrastrand stacking and favorable interactions between bases on the NT strand and aromatic residues (see the text above) likely reduce the enthalpic cost of opening the initiation bubble. Evidence for this hypothesis comes from the lack of permanganate reactivity of thymines at –7 and –10, indicating that these bases, once opened, either remain stacked with their neighbors or interact with residues in σ2. In addition, NT strand bases (–4, –3, and +2) may be partially stacked in I2, since they are only half as MnO4- reactive as in RPo.36 Thus, the model presented above is perhaps best described as “bind–bend/flip–melt,” followed by clamping. In this model, RNAP is an active participant in achieving DNA strand separation: both RNAP-induced DNA bending and the side chains of amino acids T429, Y430, and W43379,122,125,126,128 facilitate the DNA strand separation reaction. In addition, various elements in the cleft, such as the “fork loop 2” of β and multiple “tracks” of positively charged residues, appear positioned to capture and stabilize the open state via interactions with the DNA phosphate backbone.15

Sigma release versus retention

Since the discovery of sigma factor over 40 years ago,18,19 it has been thought that an obligate late step in transcription initiation was the release of sigma factor from RNAP. However, two studies131,132 clearly demonstrated the retention of σ70 in transcription complexes beyond the early phases of transcription. Compelling evidence for the presence of σ70 in elongation complexes was obtained from both FRET experiments and analysis of the proteins in the complexes. Global analysis techniques133 provided support for σ70 retention in vivo. While these studies did not demonstrate a function for the retained sigma factor, other work demonstrated that the retained sigma factor was instrumental in generating a promoter-proximal pause of transcription during the synthesis of bacteriophage (i.e., bacterial virus) λ mRNAs. Such a pause is vital for endowing the transcription complexes with the ability to read through termination signals.134 Subsequently, similar sigma-dependent pausing was demonstrated for the transcription of various bacterial genes as well5355,135,136 (see also a recent review by Artsimovitch135).

The role of the retained σ70 is to recognize –10-like sequences on the NT strand of the transcribed DNA.5355,134 The interaction of σ70 with such regions was found to be similar to its interaction with bases of the –10 element on the NT DNA. The interaction may be further strengthened by contacts to G-C base pairs52 positioned similarly to the G-C base pairs at –5 and –6, which are contacted by σ1.2 in RNAP–promoter complexes.27,28 Interestingly, the –10-like element may not be absolutely necessary, although it greatly stabilizes the interaction of sigma with the transcription complex.54 Recently described was an atypical example of σ70-dependent pausing where the –10 element was lacking but pausing was shown to be dependent on the TG sequence of the extended –10 element of the actual promoter and contacts to a C-rich region at +2–6 of the NT strand.136 The σ70–NT DNA contacts serve to lock the elongating RNAP in position, thus impeding further movement of the elongation complex. The duration of the sigma-facilitated pause is reduced by GreA and GreB proteins in vitro, and evidence is consistent with this also being the case in vivo.5355,137 This behavior is indicative of backtracking of transcription complexes during the pause.

It remains to be established whether sigma retention is characteristic of most, or all, σ70 promoters. It may be that retained σ70 is not detected unless the NT strand has the proper –10-like sequence for σ70-dependent pausing. Then the interaction of σ70 with the transcription complex would be stabilized,53,54 further delaying sigma release. Alternatively, delayed release of sigma may be a promoter-specific event for which the signals have not yet been discerned. In support of the former, experimental evidence has been obtained for a stochastic release of sigma factor from the elongating complex, which would manifest itself as a certain half-life (estimated to be on the order of 5 s) for σ70 release from RNAP during transcription elongation.138 A decreased stability of the RNAP– sigma complex is consistent with the suggestion that sigma's attachment to core during elongation differs from that in free holoenzyme.53 Indeed, it is likely that the nascent RNA, by the time it has reached 10 nt in length, will have pried loose the contacts between σ3.2 and the β′ subunit in the RNA channel of the core enzyme. At about 16 nucleotides, the contacts between σ4 and the core enzyme β flap would be disrupted. Just the interaction of σ2 and the clamp helices would then tie the sigma to the transcription complex. Interestingly, recent work indicates that release of σ2 from the β′ clamp helices is required to load the elongation factors NusG139 and rfaH,140 which thus may play a role in the release of sigma from the elongation complex. Regardless of the nature of the contacts, sigma apparently is held in a position that allows it to scan the sequence of the NT strand.

Conclusion

Determination of high-resolution structures of free and promoter-bound holoenzymes, together with advances in our understanding of how salts and solutes interact with biopolymer surfaces and perturb biopolymer processes, has led to rapid progress in our understanding of the events of RNAP recruitment and promoter recognition to form the initial closed complex RPc, and the massive conformational changes in RNAP and promoter DNA that occur to convert it to the most stable open complex RPo. Challenges for the future include developing a molecular understanding of how the start site region is opened and how the T strand is placed in the active site; how conformational changes in the cleft involving σ1.1 and the down stream NT strand in the conversion of the initial open complex (I2) to RPo are sensed by the assembling downstream clamp/jaw apparatus; how upstream DNA trajectory and interactions with the αCTDs and the upstream clamp allow the entry of downstream duplex DNA into the cleft; and how all these steps of isomerization are regulated by DNA sequence, factors, ligands, and environmental variables in the response of the cell to changing growth conditions or stress.

Supplementary Material

supp_figure_1

Acknowledgements

We thank the reviewers for helpful comments, R. H. Ebright for conversations and communication of unpublished results, and our current and former colleagues for many fruitful collaborations. We thank Caroline Davis, Wayne Kontur, Theodore Gries, and Lisa Schroeder for their significant contributions to the mechanistic work reviewed here. We gratefully acknowledge Craig Bingman and Irina Artsimovitch for discussion of and assistance with Figs. 1 and 3 and Supplementary Fig. 1, and Olga Kourennaia for discussion of and assistance with Fig. 4. R.M.S. and M.T.R. acknowledge research support from National Institutes of Health grant GM23467.

Abbreviations used

RNAP

RNA polymerase

NT

nontemplate strand

T

template strand

αCTD

α-subunit C-terminal domain

FRET

fluorescence resonance energy transfer

myxo

myxopyronin

lpm

lipiarmycin

SI

sequence insertion

Footnotes

Supplementary data associated with this article can be found, in the online version, at doi:10.1016/j. jmb.2011.01.018

References

  • 1.Browning DF, Busby SJ. The regulation of bacterial transcription initiation. Nat. Rev. Microbiol. 2004;2:57–65. doi: 10.1038/nrmicro787. [DOI] [PubMed] [Google Scholar]
  • 2.Haugen SP, Ross W, Gourse RL. Advances in bacterial promoter recognition and its control by factors that do not bind DNA. Nat. Rev. Microbiol. 2008;6:507–519. doi: 10.1038/nrmicro1912. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Wigneshweraraj S, Bose D, Burrows PC, Joly N, Schumacher J, Rappas M, et al. Modus operandi of the bacterial RNA polymerase containing the σ54 promoter-specificity factor. Mol. Microbiol. 2008;68:538–546. doi: 10.1111/j.1365-2958.2008.06181.x. [DOI] [PubMed] [Google Scholar]
  • 4.deHaseth PL, Zupancic ML, Record MT., Jr. RNA polymerase–promoter interactions: the comings and goings of RNA polymerase. J. Bacteriol. 1998;180:3019–3025. doi: 10.1128/jb.180.12.3019-3025.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Gruber TM, Gross CA. Multiple sigma subunits and the partitioning of bacterial transcription space. Annu. Rev. Microbiol. 2003;57:441–466. doi: 10.1146/annurev.micro.57.030502.090913. [DOI] [PubMed] [Google Scholar]
  • 6.Paget MS, Helmann JD. The σ70 family of sigma factors. Genome Biol. 2003;4:203. doi: 10.1186/gb-2003-4-1-203. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Hook-Barnard IG, Hinton DM. Transcription initiation by mix and match elements: flexibility for polymerase binding to bacterial promoters. Gene Regul. Syst. Biol. 2007;1:275–293. [PMC free article] [PubMed] [Google Scholar]
  • 8.Borukhov S, Nudler E. RNA polymerase: the vehicle of transcription. Trends Microbiol. 2008;16:126–134. doi: 10.1016/j.tim.2007.12.006. [DOI] [PubMed] [Google Scholar]
  • 9.Ghosh T, Bose D, Zhang X. Mechanisms for activating bacterial RNA polymerase. FEMS Microbiol. Rev. 2010;34:611–627. doi: 10.1111/j.1574-6976.2010.00239.x. [DOI] [PubMed] [Google Scholar]
  • 10.Hudson BP, Quispe J, Lara-Gonzalez S, Kim Y, Berman HM, Arnold E, et al. Three-dimensional EM structure of an intact activator-dependent transcription initiation complex. Proc. Natl Acad. Sci. USA. 2009;106:19830–19835. doi: 10.1073/pnas.0908782106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Opalka N, Brown J, Lane WJ, Twist KA, Landick R, Asturias FJ, Darst SA. Complete structural model of Escherichia coli RNA polymerase from a hybrid approach. PLoS Biol. 2010;8 doi: 10.1371/journal.pbio.1000483. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Lane WJ, Darst SA. Molecular evolution of multisubunit RNA polymerases: structural analysis. J. Mol. Biol. 2010;395:686–704. doi: 10.1016/j.jmb.2009.10.063. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Kontur WS, Saecker RM, Davis CA, Capp MW, Record MT., Jr. Solute probes of conformational changes in open complex (RPo) formation by Escherichia coli RNA polymerase at the λPR promoter: evidence for unmasking of the active site in the isomerization step and for large-scale coupled folding in the subsequent conversion to RPo. Biochemistry. 2006;45:2161–2177. doi: 10.1021/bi051835v. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.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]
  • 15.Kontur WS, Capp MW, Gries TJ, Saecker RM, Record MT., Jr Probing DNA binding, DNA opening and assembly of downstream clamp/jaw in Escherichia coli RNA polymerase–λPR promoter complexes using salt and the physiological anion glutamate. Biochemistry. 2010;49:4361–4373. doi: 10.1021/bi100092a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.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]
  • 17.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]
  • 18.Burgess RR, Travers AA, Dunn JJ, Bautz EK. Factor stimulating transcription by RNA polymerase. Nature. 1969;221:43–46. doi: 10.1038/221043a0. [DOI] [PubMed] [Google Scholar]
  • 19.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]
  • 20.Lonetto M, Gribskov M, Gross CA. The σ70 family: sequence conservation and evolutionary relationships. J. Bacteriol. 1992;174:3843–3849. doi: 10.1128/jb.174.12.3843-3849.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Gourse RL, Ross W, Gaal T. UPs and downs in bacterial transcription initiation: the role of the α 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]
  • 22.Siebenlist U, Simpson RB, Gilbert W. E. coli RNA polymerase interacts homologously with two different promoters. Cell. 1980;20:269–281. doi: 10.1016/0092-8674(80)90613-3. [DOI] [PubMed] [Google Scholar]
  • 23.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]
  • 24.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]
  • 25.Sanderson A, Mitchell JE, Minchin SD, Busby SJ. Substitutions in the Escherichia coli RNA polymerase σ70 factor that affect recognition of extended -10 elements at promoters. FEBS Lett. 2003;544:199–205. doi: 10.1016/s0014-5793(03)00500-3. [DOI] [PubMed] [Google Scholar]
  • 26.Keilty S, Rosenberg M. Constitutive function of a positively regulated promoter reveals new sequences essential for activity. J. Biol. Chem. 1987;262:6389–6395. [PubMed] [Google Scholar]
  • 27.Feklistov A, Barinova N, Sevostyanova A, Heyduk E, Bass I, Vvedenskaya I, et al. A basal promoter element recognized by free RNA polymer-ase sigma 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]
  • 28.Haugen SP, Berkmen MB, Ross W, Gaal T, Ward C, Gourse RL. rRNA promoter regulation by nonoptimal binding of sigma 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]
  • 29.Jeong W, Kang C. Start site selection at lacUV5 promoter affected by the sequence context around the initiation sites. Nucleic Acids Res. 1994;22:4667–4672. doi: 10.1093/nar/22.22.4667. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Liu J, Turnbough CL., Jr. Effects of transcriptional start site sequence and position on nucleotide-sensitive selection of alternative start sites at the pyrC promoter in Escherichia coli. J. Bacteriol. 1994;176:2938–2945. doi: 10.1128/jb.176.10.2938-2945.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Walker KA, Osuna R. Factors affecting start site selection at the Escherichia coli fis promoter. J. Bacteriol. 2002;184:4783–4791. doi: 10.1128/JB.184.17.4783-4791.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Tang Y, Murakami K, Ishihama A, deHaseth PL. Upstream interactions at the λPRM promoter are sequence nonspecific and activate the promoter to a lesser extent than an introduced UP element of an rRNA promoter. J. Bacteriol. 1996;178:6945–6951. doi: 10.1128/jb.178.23.6945-6951.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.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]
  • 34.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]
  • 35.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]
  • 36.Gries TJ, Kontur WS, Capp MW, Saecker RM, Record MT., Jr. One-step DNA melting in the RNA polymerase cleft opens the initiation bubble to form an unstable open complex. Proc. Natl Acad. Sci. USA. 2010;107:10418–10423. doi: 10.1073/pnas.1000967107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.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]
  • 38.Haugen SP, Ross W, Manrique M, Gourse RL. Fine structure of the promoter–sigma 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]
  • 39.Burr T, Mitchell J, Kolb A, Minchin S, Busby S. DNA sequence elements located immediately upstream of the -10 hexamer in Escherichia coli promoters: a systematic study. Nucleic Acids Res. 2000;28:1864–1870. doi: 10.1093/nar/28.9.1864. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Cook VM, deHaseth PL. Strand opening-deficient Escherichia 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]
  • 41.Kapanidis AN, Margeat E, Ho SO, Kortkhonjia E, Weiss S, Ebright RH. 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]
  • 42.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]
  • 43.Tchernaenko V, Halvorson HR, Kashlev M, Lutter LC. DNA bubble formation in transcription initiation. Biochemistry. 2008;47:1871–1884. doi: 10.1021/bi701289g. [DOI] [PubMed] [Google Scholar]
  • 44.Deuschle U, Kammerer W, Gentz R, Bujard H. Promoters of Escherichia coli: a hierarchy of in vivo strength indicates alternate structures. EMBO J. 1986;5:2987–2994. doi: 10.1002/j.1460-2075.1986.tb04596.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Hsu LM, Cobb IM, Ozmore JR, Khoo M, Nahm G, Xia L, et al. Initial transcribed sequence mutations specifically affect promoter escape properties. Biochemistry. 2006;45:8841–8854. doi: 10.1021/bi060247u. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Tang GQ, Roy R, Bandwar RP, Ha T, Patel SS. Real-time observation of the transition from transcription initiation to elongation of the RNA polymerase. Proc. Natl Acad. Sci. USA. 2009;106:22175–22180. doi: 10.1073/pnas.0906979106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Goldman SR, Ebright RH, Nickels BE. Direct detection of abortive RNA transcripts in vivo. Science. 2009;324:927–928. doi: 10.1126/science.1169237. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.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]
  • 49.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]
  • 50.Vassylyev DG, Sekine S, Laptenko O, Lee J, Vassylyeva MN, Borukhov S, Yokoyama S. 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]
  • 51.Young BA, Anthony LC, Gruber TM, Arthur TM, Heyduk E, Lu CZ, 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]
  • 52.Ring BZ, Roberts JW. Function of a nontranscribed DNA strand site in transcription elongation. Cell. 1994;78:317–324. doi: 10.1016/0092-8674(94)90300-x. [DOI] [PubMed] [Google Scholar]
  • 53.Brodolin K, Zenkin N, Mustaev A, Mamaeva D, Heumann H. The σ70 subunit of RNA polymerase induces lacUV5 promoter-proximal pausing of transcription. Nat. Struct. Mol. Biol. 2004;11:551–557. doi: 10.1038/nsmb768. [DOI] [PubMed] [Google Scholar]
  • 54.Nickels BE, Mukhopadhyay J, Garrity SJ, Ebright RH, Hochschild A. The σ70 subunit of RNA polymerase mediates a promoter-proximal pause at the lac promoter. Nat. Struct. Mol. Biol. 2004;11:544–550. doi: 10.1038/nsmb757. [DOI] [PubMed] [Google Scholar]
  • 55.Hatoum A, Roberts J. Prevalence of RNA polymerase stalling at Escherichia coli promoters after open complex formation. Mol. Microbiol. 2008;68:17–28. doi: 10.1111/j.1365-2958.2008.06138.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Zhang G, Campbell EA, Minakhin L, Richter C, Severinov K, Darst SA. 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]
  • 57.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]
  • 58.Sharp MM, Chan CL, Lu CZ, Marr MT, Nechaev S, Merritt EW, et al. The interface of sigma 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]
  • 59.Saecker RM, Tsodikov OV, McQuade KL, Schlax PE, Jr, Capp MW, Record MT., Jr Kinetic studies and structural models of the association of E. coli σ70 RNA polymerase with the lambdaP(R) 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]
  • 60.Murakami KS, Masuda S, Darst SA. Crystallographic analysis of Thermus aquaticus RNA polymerase holoenzyme and a holoenzyme/promoter DNA complex. Methods Enzymol. 2003;370:42–53. doi: 10.1016/S0076-6879(03)70004-4. [DOI] [PubMed] [Google Scholar]
  • 61.Chen J, Darst SA, Thirumalai D. Promoter melting triggered by bacterial RNA polymerase occurs in three steps. Proc. Natl Acad. Sci. USA. 2010;107:12523–12528. doi: 10.1073/pnas.1003533107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.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]
  • 63.Belogurov GA, Vassylyeva MN, Sevostyanova A, Appleman JR, Xiang AX, Lira R, et al. Transcription inactivation through local refolding of the RNA polymerase structure. Nature. 2009;457:332–335. doi: 10.1038/nature07510. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Ho MX, Hudson BP, Das K, Arnold E, Ebright RH. Structures of RNA polymerase–antibiotic complexes. Curr. Opin. Struct. Biol. 2009;19:715–723. doi: 10.1016/j.sbi.2009.10.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Tupin A, Gualtieri M, Leonetti JP, Brodolin K. The transcription inhibitor lipiarmycin blocks DNA fitting into the RNA polymerase catalytic site. EMBO J. 2010;29:2527–2537. doi: 10.1038/emboj.2010.135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Nagai H, Shimamoto N. Regions of the Escherichia coli primary sigma factor σ70 that are involved in interaction with RNA polymerase core enzyme. Genes Cells. 1997;2:725–734. doi: 10.1046/j.1365-2443.1997.1600357.x. [DOI] [PubMed] [Google Scholar]
  • 67.Mekler V, Kortkhonjia E, Mukhopadhyay J, Knight J, Revyakin A, Kapanidis AN, 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]
  • 68.Darst SA, Opalka N, Chacon P, Polyakov A, Richter C, Zhang G, Wriggers W. Conformational flexibility of bacterial RNA polymerase. Proc. Natl Acad. Sci. USA. 2002;99:4296–4301. doi: 10.1073/pnas.052054099. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Roe JH, Burgess RR, Record MT., Jr. Kinetics and mechanism of the interaction of Escherichia coli RNA polymerase with the λPR promoter. J. Mol. Biol. 1984;176:495–522. doi: 10.1016/0022-2836(84)90174-8. [DOI] [PubMed] [Google Scholar]
  • 70.Buc H, McClure WR. Kinetics of open complex formation between Escherichia coli RNA polymerase and the lac UV5 promoter. Evidence for a sequential mechanism involving three steps. Biochemistry. 1985;24:2712–2723. doi: 10.1021/bi00332a018. [DOI] [PubMed] [Google Scholar]
  • 71.Roe JH, Burgess RR, Record MT., Jr. Temperature dependence of the rate constants of the Escherichia coli RNA polymerase–λPR promoter interaction. Assignment of the kinetic steps corresponding to protein conformational change and DNA opening. J. Mol. Biol. 1985;184:441–453. doi: 10.1016/0022-2836(85)90293-1. [DOI] [PubMed] [Google Scholar]
  • 72.Craig ML, Tsodikov OV, McQuade KL, Schlax PE, Jr, Capp MW, Saecker RM, Record MT., Jr DNA footprints of the two kinetically significant intermediates in formation of an RNA polymerase–promoter open complex: evidence that interactions with start site and downstream DNA induce sequential conformational changes in polymerase and DNA. J. Mol. Biol. 1998;283:741–756. doi: 10.1006/jmbi.1998.2129. [DOI] [PubMed] [Google Scholar]
  • 73.Kovacic RT. The 0 °C closed complexes between Escherichia coli RNA polymerase and two promoters, T7-A3 and lacUV5. J. Biol. Chem. 1987;262:13654–13661. [PubMed] [Google Scholar]
  • 74.Cowing DW, Mecsas J, Record MT, Jr., Gross CA. Intermediates in the formation of the open complex by RNA polymerase holoenzyme containing the sigma factor σ32 at the groE promoter. J. Mol. Biol. 1989;210:521–530. doi: 10.1016/0022-2836(89)90128-9. [DOI] [PubMed] [Google Scholar]
  • 75.Schickor P, Metzger W, Werel W, Lederer H, Heumann H. Topography of intermediates in transcription initiation of E. coli. EMBO J. 1990;9:2215–2220. doi: 10.1002/j.1460-2075.1990.tb07391.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.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]
  • 77.Rogozina A, Zaychikov E, Buckle M, Heumann H, Sclavi B. DNA melting by RNA polymerase at the T7A1 promoter precedes the rate-limiting step at 37 °C and results in the accumulation of an off-pathway intermediate. Nucleic Acids Res. 2009;37:5390–5404. doi: 10.1093/nar/gkp560. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Sclavi B, Zaychikov E, Rogozina A, Walther F, Buckle M, Heumann H. 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]
  • 79.Schroeder LA, Gries TJ, Saecker RM, Record MT, Jr, Harris ME, deHaseth PL. Evidence for a tyrosine–adenine stacking interaction and for a short-lived open intermediate subsequent to initial binding of Escherichia coli RNA polymerase to promoter DNA. J. Mol. Biol. 2009;385:339–349. doi: 10.1016/j.jmb.2008.10.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Spassky A, Kirkegaard K, Buc H. Changes in the DNA structure of the lac UV5 promoter during formation of an open complex with Escherichia coli RNA polymerase. Biochemistry. 1985;24:2723–2731. doi: 10.1021/bi00332a019. [DOI] [PubMed] [Google Scholar]
  • 81.Tsodikov OV, Craig ML, Saecker RM, Record MT., Jr. Quantitative analysis of multiple-hit footprinting studies to characterize DNA conformational changes in protein–DNA complexes: application to DNA opening by Eσ70 RNA polymerase. J. Mol. Biol. 1998;283:757–769. doi: 10.1006/jmbi.1998.2130. [DOI] [PubMed] [Google Scholar]
  • 82.Murakami KS, Masuda S, Darst SA. Crystallographic analysis of Thermus aquaticus RNA polymerase holoenzymeand a holoenzyme/promoter DNA complex. Methods Enzymol. 2003;370:42–53. doi: 10.1016/S0076-6879(03)70004-4. [DOI] [PubMed] [Google Scholar]
  • 83.Mecsas J, Cowing DW, Gross CA. Development of RNA polymerase–promoter contacts during open complex formation. J. Mol. Biol. 1991;220:585–597. doi: 10.1016/0022-2836(91)90102-c. [DOI] [PubMed] [Google Scholar]
  • 84.Bartlett MS, Gaal T, Ross W, Gourse RL. RNA polymerase mutants that destabilize RNA polymerase–promoter complexes alter NTP-sensing by rrn P1 promoters. J. Mol. Biol. 1998;279:331–345. doi: 10.1006/jmbi.1998.1779. [DOI] [PubMed] [Google Scholar]
  • 85.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]
  • 86.Rutherford ST, Villers CL, Lee JH, Ross W, Gourse RL. Allosteric control of Escherichia coli rRNA promoter complexes by DksA. Genes Dev. 2009;23:236–248. doi: 10.1101/gad.1745409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Craig ML, Suh WC, Record MT., Jr. HO and DNase I probing of Eσ70 RNA polymerase–λPR promoter open complexes: Mg2+ binding and its structural consequences at the transcription start site. Biochemistry. 1995;34:15624–15632. doi: 10.1021/bi00048a004. [DOI] [PubMed] [Google Scholar]
  • 88.Ross W, Gosink KK, Salomon J, Igarashi K, Zou C, Ishihama A, et al. A third recognition element in bacterial promoters: DNA binding by the α subunit of RNA polymerase. Science. 1993;262:1407–1413. doi: 10.1126/science.8248780. [DOI] [PubMed] [Google Scholar]
  • 89.Busby S, Ebright RH. Transcription activation by catabolite activator protein (CAP). J. Mol. Biol. 1999;293:199–213. doi: 10.1006/jmbi.1999.3161. [DOI] [PubMed] [Google Scholar]
  • 90.Dethiollaz S, Eichenberger P, Geiselmann J. Influence of DNA geometry on transcriptional activation in Escherichia coli. EMBO J. 1996;15:5449–5458. [PMC free article] [PubMed] [Google Scholar]
  • 91.Giladi H, Igarashi K, Ishihama A, Oppenheim AB. Stimulation of the phage lambda pL promoter by integration host factor requires the carboxy terminus of the α-subunit of RNA polymerase. J. Mol. Biol. 1992;227:985–990. doi: 10.1016/0022-2836(92)90514-k. [DOI] [PubMed] [Google Scholar]
  • 92.Meng W, Belyaeva T, Savery NJ, Busby SJ, Ross WE, Gaal T, et al. UP element-dependent transcription at the Escherichia coli rrnB P1 promoter: positional requirements and role of the RNA polymerase α subunit linker. Nucleic Acids Res. 2001;29:4166–4178. doi: 10.1093/nar/29.20.4166. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.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]
  • 94.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 sigma region 4. Genes Dev. 2003;17:1293–1307. doi: 10.1101/gad.1079403. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.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]
  • 96.Lonetto MA, 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]
  • 97.Kolb A, Igarashi K, Ishihama A, Lavigne M, Buckle M, Buc H. E. coli RNA polymerase, deleted in the C-terminal part of its α-subunit, interacts differently with the cAMP–CRP complex at the lacP1 and at the galP1 promoter. Nucleic Acids Res. 1993;21:319–326. doi: 10.1093/nar/21.2.319. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Spolar RS, Record MT., Jr. Coupling of local folding to site-specific binding of proteins to DNA. Science. 1994;263:777–784. doi: 10.1126/science.8303294. [DOI] [PubMed] [Google Scholar]
  • 99.Yin YW, Steitz TA. Structural basis for the transition from initiation to elongation transcription in T7 RNA polymerase. Science. 2002;298:1387–1395. doi: 10.1126/science.1077464. [DOI] [PubMed] [Google Scholar]
  • 100.Ederth J, Artsimovitch I, Isaksson LA, Landick R. The downstream DNA jaw of bacterial RNA polymerase facilitates both transcriptional initiation and pausing. J. Biol. Chem. 2002;277:37456–37463. doi: 10.1074/jbc.M207038200. [DOI] [PubMed] [Google Scholar]
  • 101.Zaychikov E, Denissova L, Heumann H. Translocation of the Escherichia coli transcription complex observed in the registers 11 to 20: “jumping” of RNA polymerase and asymmetric expansion and contraction of the “transcription bubble”. Proc. Natl Acad. Sci. USA. 1995;92:1739–1743. doi: 10.1073/pnas.92.5.1739. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Brodolin K, Zenkin N, Severinov K. Remodeling of the σ70 subunit non-template DNA strand contacts during the final step of transcription initiation. J. Mol. Biol. 2005;350:930–937. doi: 10.1016/j.jmb.2005.05.048. [DOI] [PubMed] [Google Scholar]
  • 103.Schwartz EC, Shekhtman A, Dutta K, Pratt MR, Cowburn D, Darst S, Muir TW. A full-length group 1 bacterial sigma factor adopts a compact structure incompatible with DNA binding. Chem. Biol. 2008;15:1091–1103. doi: 10.1016/j.chembiol.2008.09.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Camarero JA, Shekhtman A, Campbell EA, Chlenov M, Gruber TM, Bryant DA, et al. Autoregulation of a bacterial sigma factor explored by using segmental isotopic labeling and NMR. Proc. Natl Acad. Sci. USA. 2002;99:8536–8541. doi: 10.1073/pnas.132033899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Dombroski AJ, Walter WA, Record MT, Jr, Siegele DA, Gross CA. Polypeptides containing highly conserved regions of transcription initiation factor σ70 exhibit specificity of binding to promoter DNA. Cell. 1992;70:501–512. doi: 10.1016/0092-8674(92)90174-b. [DOI] [PubMed] [Google Scholar]
  • 106.Chen YF, Helmann JD. The Bacillus subtilis flagellar regulatory protein σD: overproduction, domain analysis and DNA-binding properties. J. Mol. Biol. 1995;249:743–753. doi: 10.1006/jmbi.1995.0333. [DOI] [PubMed] [Google Scholar]
  • 107.Sevim E, Gaballa A, Belduz AO, Helmann JD. DNA-binding properties of the Bacillus subtilis and Aeribacillus pallidus AC6 σD proteins. J. Bacteriol. 2011;193:575–579. doi: 10.1128/JB.01193-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Mekler V, Pavlova O, Severinov K. The interaction of Escherichia coli RNA polymerase σ70 subunit with promoter elements in the context of free σ70, RNA polymerase holoenzyme, and the β′–σ70 complex. J. Biol. Chem. 2011;286:270–279. doi: 10.1074/jbc.M110.174102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Devi PG, Campbell EA, Darst SA, Nickels BE. Utilization of variably spaced promoter-like elements by the bacterial RNA polymerase holoenzyme during early elongation. Mol. Microbiol. 2010;75:607–622. doi: 10.1111/j.1365-2958.2009.07021.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Niedziela-Majka A, Heyduk T. Escherichia coli RNA polymerase contacts outside the -10 promoter element are not essential for promoter melting. J. Biol. Chem. 2005;280:38219–38227. doi: 10.1074/jbc.M507984200. [DOI] [PubMed] [Google Scholar]
  • 111.Grossman AD, Erickson JW, Gross CA. The htpR gene product of E. coli is a sigma factor for heat-shock promoters. Cell. 1984;38:383–390. doi: 10.1016/0092-8674(84)90493-8. [DOI] [PubMed] [Google Scholar]
  • 112.Grossman AD, Straus DB, Walter WA, Gross CA. σ32 synthesis can regulate the synthesis of heat shock proteins in Escherichia coli. Genes Dev. 1987;1:179–184. doi: 10.1101/gad.1.2.179. [DOI] [PubMed] [Google Scholar]
  • 113.Koo BM, Rhodius VA, Campbell EA, Gross CA. Dissection of recognition determinants of Escherichia coli σ32 suggests a composite -10 region with an ‘extended -10’ motif and a core -10 element. Mol. Microbiol. 2009;72:815–829. doi: 10.1111/j.1365-2958.2009.06690.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Koo BM, Rhodius VA, Nonaka G, deHaseth PL, Gross CA. Reduced capacity of alternative sigmas to melt promoters ensures stringent promoter recognition. Genes Dev. 2009;23:2426–2436. doi: 10.1101/gad.1843709. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Guo Y, Gralla JD. Promoter opening via a DNA fork junction binding activity. Proc. Natl Acad. Sci. USA. 1998;95:11655–11660. doi: 10.1073/pnas.95.20.11655. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.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]
  • 117.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]
  • 118.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]
  • 119.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]
  • 120.Shultzaberger RK, Chen Z, Lewis KA, Schneider TD. Anatomy of Escherichia coli σ70 promoters. Nucleic Acids Res. 2007;35:771–788. doi: 10.1093/nar/gkl956. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Li XY, McClure WR. Stimulation of open complex formation by nicks and apurinic sites suggests a role for nucleation of DNA melting in Escherichia coli promoter function. J. Biol. Chem. 1998;273:23558–23566. doi: 10.1074/jbc.273.36.23558. [DOI] [PubMed] [Google Scholar]
  • 122.Schroeder LA, Choi AJ, deHaseth PL. The -11A of promoter DNA and two conserved amino acids in the melting region of σ70 both directly affect the rate limiting step in formation of the stable RNA polymerase–promoter complex, but they do not necessarily interact. Nucleic Acids Res. 2007;35:4141–4153. doi: 10.1093/nar/gkm431. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Malhotra A, Severinova E, Darst SA. Crystal structure of a σ70 subunit fragment from E. coli RNA polymerase. Cell. 1996;87:127–136. doi: 10.1016/s0092-8674(00)81329-x. [DOI] [PubMed] [Google Scholar]
  • 124.Campbell EA, Muzzin O, Chlenov M, Sun JL, Olson CA, Weinman O, et al. Structure of the bacterial RNA polymerase promoter specificity sigma subunit. Mol. Cell. 2002;9:527–539. doi: 10.1016/s1097-2765(02)00470-7. [DOI] [PubMed] [Google Scholar]
  • 125.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]
  • 126.Panaghie G, Aiyar SE, Bobb KL, Hayward RS, de Haseth PL. Aromatic amino acids in region 2.3 of Escherichia coli σ70 participate collectively in the formation of an RNA polymerase–promoter open complex. J. Mol. Biol. 2000;299:1217–1230. doi: 10.1006/jmbi.2000.3808. [DOI] [PubMed] [Google Scholar]
  • 127.Juang YL, Helmann JD. A promoter melting region in the primary sigma factor of Bacillus subtilis. Identification of functionally important aromatic amino acids. J. Mol. Biol. 1994;235:1470–1488. doi: 10.1006/jmbi.1994.1102. [DOI] [PubMed] [Google Scholar]
  • 128.Tomsic M, Tsujikawa L, Panaghie G, Wang Y, Azok J, deHaseth PL. 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]
  • 129.Helmann JD, Chamberlin MJ. Structure and function of bacterial sigma factors. Annu. Rev. Biochem. 1988;57:839–872. doi: 10.1146/annurev.bi.57.070188.004203. [DOI] [PubMed] [Google Scholar]
  • 130.Holbrook JA, Capp MW, Saecker RM, Record MT., Jr. Enthalpy and heat capacity changes for formation of an oligomeric DNA duplex: interpretation in terms of coupled processes of formation and association of single-stranded helices. Biochemistry. 1999;38:8409–8422. doi: 10.1021/bi990043w. [DOI] [PubMed] [Google Scholar]
  • 131.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]
  • 132.Mukhopadhyay J, Kapanidis AN, Mekler V, Kortkhonjia E, Ebright YW, Ebright RH. 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]
  • 133.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]
  • 134.Ring BZ, Yarnell WS, Roberts JW. Function of E. coli RNA polymerase sigma factor σ70 in promoter-proximal pausing. Cell. 1996;86:485–493. doi: 10.1016/s0092-8674(00)80121-x. [DOI] [PubMed] [Google Scholar]
  • 135.Artsimovitch I. Post-initiation control by the initiation factor sigma. Mol. Microbiol. 2008;68:1–3. doi: 10.1111/j.1365-2958.2008.06136.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Stepanova E, Wang M, Severinov K, Borukhov S. Early transcriptional arrest at Escherichia coli rplN and ompX promoters. J. Biol. Chem. 2009;284:35702–35713. doi: 10.1074/jbc.M109.053983. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Marr MT, Roberts JW. Function of transcription cleavage factors GreA and GreB at a regulatory pause site. Mol. Cell. 2000;6:1275–1285. doi: 10.1016/s1097-2765(00)00126-x. [DOI] [PubMed] [Google Scholar]
  • 138.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]
  • 139.Mooney RA, Schweimer K, Rosch P, Gottesman M, Landick R. Two structurally independent domains of E. coli NusG create regulatory plasticity via distinct interactions with RNA polymerase and regulators. J. Mol. Biol. 2009;391:341–358. doi: 10.1016/j.jmb.2009.05.078. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Belogurov GA, Mooney RA, Svetlov V, Landick R, Artsimovitch I. Functional specialization of transcription elongation factors. EMBO J. 2009;28:112–122. doi: 10.1038/emboj.2008.268. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

supp_figure_1

RESOURCES