Elsevier

Methods

Volume 47, Issue 1, January 2009, Pages 13-24
Methods

Review Article
Analysis of RNA polymerase-promoter complex formation

https://doi.org/10.1016/j.ymeth.2008.10.018Get rights and content

Abstract

Bacterial promoter identification and characterization is not as straightforward as one might presume. Promoters vary widely in their similarity to the consensus recognition element sequences, in their activities, and in their utilization of transcription factors, and multiple approaches often must be used to provide a framework for understanding promoter regulation. Characterization of RNA polymerase-promoter complex formation in the absence of additional regulatory factors (basal promoter function) can provide a basis for understanding the steps in transcription initiation that are ultimately targeted by nutritional or environmental factors. Promoters can be localized using genetic approaches in vivo, but the detailed properties of the RNAP-promoter complex are studied most productively in vitro. We first describe approaches for identification of bacterial promoters and transcription start sites in vivo, including promoter-reporter fusions and primer-extension. We then describe a number of methods for characterization of RNAP-promoter complexes in vitro, including in vitro transcription, gel mobility shift assays, footprinting, and filter binding. Utilization of these methods can result in determination of not only basal promoter strength but also the rates of transcription initiation complex formation and decay.

Introduction

Bacterial transcription is carried out by a multisubunit RNA polymerase holoenzyme, Eσ, comprised of the core enzyme E (subunits α2, β, β′ and ω), and one of several σ specificity subunits (a major “housekeeping” sigma factor, such as Escherichia coli σ70, or one of a variable number of alternative sigma factors present in different bacterial species) [1], [2], [3], [4]. While the response of RNAP to specialized DNA-binding transcription factors plays a critical role in determining the cell’s transcription program under different growth conditions, the promoter sequences with which RNAP interacts are variable and are a major determinant of the wide range of strength and regulation of gene expression. RNAP-promoter complex formation and transcription initiation in the absence of trans-acting factors are referred to as basal promoter function. In this chapter we focus on methods for observing and studying basal promoter function as a key component of the overall mechanism of transcription regulation. The mechanism of action of regulatory factors (activators or repressors) must ultimately be understood in terms of how they alter the interactions of RNAP with a particular promoter.

An RNAP-promoter complex capable of transcription initiation (an open complex; RPO) is formed through interactions that involve several regions of RNAP and that span 70–80 bp of promoter DNA (∼−60 to +20 with respect to the transcription start site, +1). Each RNAP subunit, with the exception of ω, participates in these interactions, although a majority of the sequence-specific interactions occur with the σ subunit. In an E. coli70-promoter complex, sequence-specific interactions with σ70 occur at the −10 element (with σ70 regions 2.3–2.4), the extended −10 element (σ70 region 3.0), the −35 element (σ70 region 4.2), and the discriminator element immediately downstream of the −10 hexamer (σ70 region 1.2) (Fig. 1; [1], [5]). In addition, the C-terminal domain of α subunit can interact sequence-specifically with the UP element, located upstream of the −35 hexamer [6]. Sequence non-specific DNA contacts are mediated by several regions of the large β and β′ subunits in the active site channel and downstream of the start site (as observed in elongation complexes; [7], [8]) and in the spacer region between the −10 and −35 hexamers [9].

Promoters vary widely in overall strength (∼4 orders of magnitude) and in the extent of similarity to the consensus sequences for the recognition elements [1], [10], [11]. The −10 element is the most highly conserved of the promoter elements, but similarities to consensus usually are found in one or more of the other recognition elements as well [6], [12], [13]. The contribution of each recognition element to promoter function depends upon the context in which it is found.

In this review we discuss several methods for investigating RNAP-promoter complex formation, including preliminary identification of the promoter itself using in vivo methods, followed by the characterization of properties of complex formation with RNAP in vitro. Formation of the RNAP-promoter complex can be strongly affected by several parameters in vitro, and effects of these conditions vary with different promoter sequences. These parameters will be discussed as well.

Section snippets

Promoter identification and activity determination in vivo

While bioinformatic analyses can predict some promoters correctly, definitive identification of promoters from sequence information alone remains difficult [14], [15]. In addition, while in vitro methods that examine RNA polymerase-promoter complex formation can be very informative, they can identify interactions that do not correspond to promoters that function in vivo (e.g., “tight-binding sites”, or end-binding sites; [16], [17]). Thus, correct identification of a promoter is established

RNAP-promoter complex formation in vitro

The interaction of purified RNA polymerase with a promoter to form a complex in vitro can be detected and studied by several methods, including in vitro transcription, gel mobility shift (EMSA), footprinting, or filter binding assays. The process of RNAP-promoter complex formation is affected greatly by several parameters, including salt concentration, temperature, template topology, and RNA polymerase concentration. The optimum conditions for RNAP complex formation with a particular promoter

Acknowledgments

We thank Pete Chandrangsu and Justin Lemke for Fig. 2, Fig. 3, and members of our laboratory for helpful discussions. This work was supported by National Institutes of Health Grant R37 GM37048 to R.L.G.

References (86)

  • S.P. Haugen et al.

    Cell

    (2006)
  • A.M. Huerta et al.

    J. Mol. Biol.

    (2003)
  • N.B. Reppas et al.

    Mol. Cell

    (2006)
  • D.A. Schneider et al.

    Methods Enzymol.

    (2003)
  • A.L. Lloyd et al.

    J. Microbiol. Methods

    (2005)
  • J.T. Newlands et al.

    J. Mol. Biol.

    (1991)
  • L. Rao et al.

    J. Mol. Biol.

    (1994)
  • R.W. Simons et al.

    Gene

    (1987)
  • J. Lodge et al.

    FEMS Microbiol. Lett.

    (1992)
  • R.L. Gourse et al.

    Cell

    (1986)
  • V. Bergendahl et al.

    Protein Expression and Purification

    (2003)
  • T.T. Su et al.

    J. Biol. Chem.

    (1994)
  • J. Davison

    Gene

    (1984)
  • J.-H. Roe et al.

    J. Mol. Biol.

    (1985)
  • R.M. Saecker et al.

    J. Mol. Biol.

    (2002)
  • M.M. Barker et al.

    J. Mol. Biol.

    (2001)
  • S. Borukhov et al.

    J. Biol. Chem.

    (1993)
  • H.D. Murray et al.

    Mol. Cell

    (2003)
  • P.H. von Hippel

    Cell

    (2006)
  • D.C. Straney et al.

    Cell

    (1985)
  • P. Hershberger et al.

    J. Mol. Biol.

    (1991)
  • X. Zhang et al.

    J. Mol. Biol.

    (1996)
  • D.C. Straney et al.

    J. Mol. Biol.

    (1987)
  • E. Severinova et al.

    J. Mol. Biol.

    (1998)
  • V. Cook et al.

    J. Biol. Chem.

    (2007)
  • M.S. Bartlett et al.

    J. Mol. Biol.

    (1998)
  • W.J. Dixon et al.

    Methods in Enzymology

    (1991)
  • S. Sasse-Dwight et al.

    Methods in Enzymology

    (1991)
  • U. Siebenlist et al.

    Cell

    (1980)
  • A. Maxam et al.

    Methods Enzymol.

    (1980)
  • X.-Y. Li et al.

    J. Biol. Chem.

    (1998)
  • M.M. Barker et al.

    J. Mol. Biol.

    (2001)
  • M.T. Record et al.
  • P.L. deHaseth et al.

    J. Bacteriol.

    (1998)
  • J.D. Helmann et al.

    Biochemistry

    (1999)
  • T.M. Gruber et al.

    Annu. Rev. Microbiol.

    (2003)
  • S.P. Haugen et al.

    Nat. Rev. Microbiol.

    (2008)
  • R.L. Gourse et al.

    Mol. Microbiol.

    (2000)
  • N. Korsheva et al.

    Curr. Opin. Microbiol.

    (2001)
  • D.G. Vassylyev et al.

    Nature

    (2007)
  • K.S. Murakami et al.

    Science

    (2002)
  • S. Lisser et al.

    Nucleic Acids Res.

    (1993)
  • R.K. Shultzaberger et al.

    Nucleic Acids Res.

    (2007)
  • Cited by (55)

    • Structural Basis for Virulence Activation of Francisella tularensis

      2021, Molecular Cell
      Citation Excerpt :

      Strikingly, examination of the DNA sequences of promoters known to be regulated by PigR, such as the iglA, pdpA, and FTL_0026 promoters, revealed that all contain AT-rich regions flanking their PRE sites (Figure 4B). Studies in model bacteria have revealed the importance of AT-rich regions located upstream of the −35 element (Newlands et al., 1991; Ross et al., 1993; Estrem et al., 1998; Ross and Gourse, 2009), termed Upstream (UP) elements, as contact sites for the C-terminal domains of the α subunits (αCTDs) of RNAP (Newlands et al., 1991; Ross et al., 1993; Estrem et al., 1998; Ross and Gourse, 2009). Consistent with the assignment of these densities in our structure as FtαCTDs, lower contoured density maps revealed a connection between the densities bound to the AT-rich DNA sites and the FtRNAP αNTDs, and the EcαCTD structure could be readily docked into these densities (Figure 4C).

    • Stepwise Promoter Melting by Bacterial RNA Polymerase

      2020, Molecular Cell
      Citation Excerpt :

      The upstream protection boundary (−54) remained unchanged, indicating that an intermediate complex is formed at rpsT P2 either by reducing the temperature or by TraR. The pattern of unstacked thymines detected by KMnO4 footprinting (Ross and Gourse, 2009) in these complexes suggested that the transcription bubble was partially melted. At both 23°C and 37°C with RNAP alone, non-template strand (nt-strand) Ts at −10, −8, and −4 were KMnO4 reactive with RNAP alone (Figure 1C, lanes 3 and 11; Ts at +3, +4, and +5 in the 37°C complex were also reactive; lane 11, indicating “scrunching” of a minor fraction of the complexes at this temperature; Winkelman et al., 2016).

    View all citing articles on Scopus
    View full text