Transcription is the principal control point for bacterial gene expression, and it enables a global cellular response to an intracellular or environmental trigger. Transcriptional regulation is orchestrated by transcription factors, which activate or repress transcription of target genes by modulating the activity of RNA polymerase. Dissecting the nature and precise choreography of these interactions is essential for developing a molecular understanding of transcriptional regulation. While the contribution of X-ray crystallography has been invaluable, the ‘resolution revolution’ of cryo-electron microscopy has transformed our structural investigations, enabling large, dynamic and often transient transcription complexes to be resolved that in many cases had resisted crystallisation. In this review, we highlight the impact cryo-electron microscopy has had in gaining a deeper understanding of transcriptional regulation in bacteria. We also provide readers working within the field with an overview of the recent innovations available for cryo-electron microscopy sample preparation and image reconstruction of transcription complexes.

The expression of a gene to assemble a protein (the Central Dogma) is a fundamental process of life. Transcription is the first step in gene expression, which is coordinated by a complex of proteins that cooperate at the promoter region to transcribe the DNA gene sequence into mRNA. The lead actor of transcription, RNA Polymerase (RNAP), comprises a core subunit of proteins labelled α2ββ′ω (Figure 1A) [1]. Upon interaction with a sigma (σ)-factor, RNAP forms the active holoenzyme [2,3]. Sigma factors are large, multi-domain proteins that bind various sites across the core RNAP and promoter DNA [4], and are responsible for guiding RNAP to the transcription start sites by locating the −35 element (consensus sequence TTGACA) and the −10 element (consensus sequence TATAAT) within the promoter [5] (Figure 1A). These two DNA elements constitute major features of a bacterial promoter and serve as notable controllers of transcriptional activity. Bacterial promoters are also decorated with other regulatory components, such as −10 extension (EXT) [6], discriminator (DISC) [7,8], and the upstream element (UP element) [9]. Once RNAP holoenzyme binds DNA, a series of conformational changes serve to manipulate the DNA and unwind a 13 base pair region (promoter melting) to initiate transcription.

Schematic model of bacterial transcription initiation and regulation by transcription factors.

Figure 1.
Schematic model of bacterial transcription initiation and regulation by transcription factors.

(A) In bacteria, transcription is initiated by the formation of the RNA polymerase (RNAP) holoenzyme, which comprises the RNAP core and a sigma (σ) factor (e.g. σ70). Upon interaction with promoter DNA, the RNAP holoenzyme forms a closed transcription complex. The subunits of the RNAP core, σ factor and sequence-specific interactions with the promoter DNA are illustrated. RNAP core: α N-terminal/C-terminal domains (αNTD/CTD) — blue; β and β′ — grey; ω — black. σ factor — orange, linker regions (black lines). Promoter DNA (grey): UP element — cyan; −35 element — yellow; spacer — white; extended −10 (EXT) — violet; −10 element — dark red; discriminator (DISC) — peach; transcription start site (+1, arrow). Sequence-specific interactions are highlighted with dash lines. (B) Transcription factors, known as activators or repressors can activate or repress transcription initiation, respectively. The general mechanism of each is shown here. Transcriptional activators (left panel, green) bind to a site upstream of the promoter (−35 and −10 element) and transcription start site (+1, arrow) where they can recruit the RNAP holoenzyme by interacting with the C-terminal of the α-subunit (αCTD) of RNAP. This process can be enhanced by small molecules or effectors (purple hexagon) to increase the rate of transcription. In contrast, transcriptional repressors (right panel, red) bind to a site that overlaps the core −35 and −10 elements of the promoter to directly block the binding of RNAP to the promoter, switching gene transcription off (shown by red cross). In the presence of an effector (blue hexagon) the DNA affinity is reduced, and the repressor dissociates from the promoter. This allows RNAP to be recruited, switching gene transcription on.

Figure 1.
Schematic model of bacterial transcription initiation and regulation by transcription factors.

(A) In bacteria, transcription is initiated by the formation of the RNA polymerase (RNAP) holoenzyme, which comprises the RNAP core and a sigma (σ) factor (e.g. σ70). Upon interaction with promoter DNA, the RNAP holoenzyme forms a closed transcription complex. The subunits of the RNAP core, σ factor and sequence-specific interactions with the promoter DNA are illustrated. RNAP core: α N-terminal/C-terminal domains (αNTD/CTD) — blue; β and β′ — grey; ω — black. σ factor — orange, linker regions (black lines). Promoter DNA (grey): UP element — cyan; −35 element — yellow; spacer — white; extended −10 (EXT) — violet; −10 element — dark red; discriminator (DISC) — peach; transcription start site (+1, arrow). Sequence-specific interactions are highlighted with dash lines. (B) Transcription factors, known as activators or repressors can activate or repress transcription initiation, respectively. The general mechanism of each is shown here. Transcriptional activators (left panel, green) bind to a site upstream of the promoter (−35 and −10 element) and transcription start site (+1, arrow) where they can recruit the RNAP holoenzyme by interacting with the C-terminal of the α-subunit (αCTD) of RNAP. This process can be enhanced by small molecules or effectors (purple hexagon) to increase the rate of transcription. In contrast, transcriptional repressors (right panel, red) bind to a site that overlaps the core −35 and −10 elements of the promoter to directly block the binding of RNAP to the promoter, switching gene transcription off (shown by red cross). In the presence of an effector (blue hexagon) the DNA affinity is reduced, and the repressor dissociates from the promoter. This allows RNAP to be recruited, switching gene transcription on.

Close modal

In bacteria, the regulation of gene expression occurs primarily at transcription initiation, allowing bacteria to maintain homeostasis and adapt to changing environmental conditions, such as nutrient availability, by changing which genes are expressed, and which are not [excellent reviews on gene regulation focused on initiation can be found here [10,11]]. Transcriptional regulation in bacteria is predominantly modulated by transcription factors, an important class of trans-acting factor, that bind DNA and function as either activators or repressors of gene expression (Figure 1B). Whereas transcriptional activators generally bind upstream of the RNAP-binding site to co-opt RNAP and enhance activity, transcriptional repressors bind to the operator region of target genes to directly obstruct the binding and activity of the RNAP [10,12].

Transcription factors largely share a common domain architecture, comprising an N-terminal DNA-binding domain and a C-terminal effector-binding domain, typically connected by a flexible linker (Figure 1B). The DNA-binding domain recognises a specific DNA sequence and most often contains the highly conserved, helix-turn-helix motif, while the effector-binding domain functions as a signal sensor [13]. In general, the effector is a small molecule or pathway metabolite that allosterically binds the protein to trigger a conformational change that alters the affinity of the DNA-binding domain to the target sequence [12,13]. This collective function enables transcription factors to act as molecular switches, enabling bacteria to rapidly respond to sudden environmental challenges [14].

In addition to transcription factors that directly modulate RNAP activity, σ-factors also serve as an important class of trans-acting factor. By acting in complex with RNAP, σ-factors facilitate broader changes, such as the expression of genes required for bacterial cell viability [10,12] (reviewed here [10,15,16]). Overall, RNAP serves as the core element of gene regulation, combining information from an assortment of sensory systems to appropriately modulate gene expression — the net outcome is to determine which genes are transcribed, and to what extent, under any specific growth condition [17].

Over the past few decades, structural biology has been instrumental for defining these regulatory mechanisms and the biological function of RNAP during transcription [18]. Yet, despite a focused effort to delineate these processes, principally through X-ray crystallography, technical limitations have restricted our molecular understanding of transcriptional regulation. Notably, these include the intrinsic flexibility, conformational/compositional heterogeneity, and transient nature of the transcription complexes [19,20]. Crystal growth is usually easiest when the macromolecules are stably folded, can be concentrated, are homogenous (i.e. purified away from all other contaminants and in a stable oligomeric state), and have limited flexibility. Fuelled by technological breakthroughs in data collection and imaging processing (reviewed in [21]), cryo-electron microscopy (cryo-EM) now offers a powerful alternative to overcome these challenges. This capacity has granted researchers the power to ‘see’ transient or heterogeneous complexes that are unamenable to crystallisation and determine their structure at or near atomic-level resolution.

Herein, we review the contribution of cryo-EM to further our understanding of transcriptional regulation in bacteria, with a focus on studies that have provided key mechanistic insights into transcription initiation. We also highlight state-of-the-art sample preparation and 3D reconstruction strategies for structure determination with a particular focus on ‘tricks’ for protein–nucleic acid complexes.

Visualising transcription complexes at the atomic level is essential for unravelling their mechanism of function. Over the past few years, cryo-EM has been indispensable for resolving large and heterogeneous complexes, where previous crystallographic studies have come up short, providing over 65 structures to date (summarised in Table 1). Here, we highlight pivotal transcription complexes active during transcription initiation, which contain transcriptional activators and repressors that until the advent of cryo-EM weren't fully understood.

Table 1
Summary of bacterial transcription complex structures resolved by cryo-EM
Transcription complexFamilyOrganismEMDB IDPDB IDReference
EcmrR-RPo2 MerR Escherichia coli EMD-22234 6XL5 [22
EcmrR-RPo2 (EcmrR-spacer DNA complex) EMD-22235 6XL6 
EcmrR-RPint3 with 3 nt RNA transcript EMD-22236 6XL9 
EcmrR-RPint3 with 3 nt RNA transcript (EcmrR-spacer DNA complex) EMD-22237 6XLA 
EcmrR-RPint3 with 4 nt RNA transcript EMD-22245 6XLJ 
EcmrR-RPint3 with 4 nt RNA transcript (EcmrR-spacer DNA complex) EMD-22246 6XLK 
EcmrR-RPo2 (clearer σ70 density) EMD-23291 
BmrR-RNA polymerase complex MerR Bacillus subtilis EMD-30390 7CKQ [23
CueR-RNA polymerase complex (without RNA transcript) MerR Escherichia coli EMD-22184 6XH7
6XH8 
[24
CueR-RNA polymerase complex (with RNA transcript) EMD-22185 
CueR-RNA polymerase complex (clearer σ70 density) EMD-22289 
CueR- RNA polymerase complex MerR Escherichia coli EMD-30268 6LDI [25
CueR- RNA polymerase complex (with fully duplex DNA) EMD-0874 7C17 
NanR-dimer1/DNA complex GntR Escherichia coli EMD-21652 6WFQ [26
NanR-dimer3/DNA complex EMD-21661 6WG7 
BusR-tetramer1/pAB DNA complex GntR Streptococcus agalactiae EMD-13119 7OZ3 [27
BusR-tetramer1/pAB1 DNA complex EMD-12051 7B5Y 
TraR-Eσ70 (state I) LuxR Escherichia coli EMD-0348 6N57 [28
TraR-Eσ70 (state II) EMD-0349 6N58 
TraR-Eσ70 (state III) EMD-20231 N/A 
MmfR-dimer2/DNA complex TetR Streptomyces coelicolor EMD-20781 N/A [29
Class-II CAP-TAC1 without RNA transcript (state I) CAP Escherichia coli EMD-20287 6PB5 [30
Class-II CAP-TAC1 without RNA transcript (state II) EMD-20288 6PB6 
Class-II CAP-TAC1 with RNA transcript (state II) EMD-20286 6PB4 
Class-I CAP-TAC1 CAP Escherichia coli EMD-7059 6B6F [31
Class-I CAP-TAC1 (focused map on αCTD-CAP region) EMD-7060 
Crl-EσS-RNA polymerase complex Crl Escherichia coli EMD-200090 6OMF [32
Spx-RNA polymerase complex Spx Bacillus subtilis EMD-31485 7F75 [33
WhiB7-RPo2 WhiB Mycobacterium tuberculosis EMD-22886 7KIF [34
WhiB7-RPc4 EMD-22887 7KIM 
Rgg2-short hydrophobic peptide complex Rgg Streptococcus thermophilus EMD-22341 7JI0 [35
GreB-RNA polymerase elongation complex (pre-RNA cleavage) Gre Escherichia coli EMD-4892 6RIN [36
GreB-RNA polymerase elongation complex (post-RNA cleavage) EMD-4885 6RI7 
GreB-RNA polymerase reactivated complex (before RNA extension) EMD-4882 6RH3 
CarD-RPo2 CarD Mycobacterium tuberculosis EMD-9037 6EDT [37
CarD-RNA polymerase intermediate (with 8-nt RNA transcript) EMD-9039 6EE8 
CarD-RPo2 (with corallopyronin A) EMD-9041 6EEC 
CarD-RNA polymerase holoenzyme (with corallopyronin A) EMD-9047 6M7J 
CarD-RPo2 (with Sorangicin A) CarD Mycobacterium tuberculosis EMD-21407 6VVY [38
CarD-S456LRPo2 (with Sorangicin A) EMD-21408 6VW0 
CarD-RPo2 (with Sorangicin A) (with 8-nt RNA transcript) EMD-21406 6VVX 
CarD-S456LRPo2 (with Sorangicin A) (with 8-nt RNA transcript) EMD-21409 6VVZ 
SspA-σ70-RPo2 GST Escherichia coli EMD-30307 7C97 [39
DksA-RPo2 (State I) with guanosine tetraphosphate (ppGpp) DksA Escherichia coli EMD-21881 7KHI [40
DksA-RPo2 (State II) with guanosine tetraphosphate (ppGpp) EMD-21883 7KHE 
NusG-opsEC NusG Escherichia coli EMD-7351 6C6U [41
RfaH-NusG-N-Term-opsEC RfaH EMD-7350 6C6T 
RfaH-full-length-opsEC RfaH EMD-7349 6C6S 
RNAP-HelD HelD Bacillus subtillus EMD-21921 6WVK [42
Msm HelD–RNAP complex State I HelD Mycobacterium smegmatis EMD-10996 6YXU [43
Msm HelD–RNAP complex State II EMD-11004 6YYS 
Msm HelD–RNAP complex State III EMD- 11026 6Z11 
Spt4/5-RNAP complex (with antibodies) Spt4/5 Pyrococcus furiosus EMD- 1840 N/A [44
Mfd-dependent transcription termination complex MfD Thermus thermophilus EMD- 30117 6M6A [45
Mfd-dependent transcription termination complex with ATPγS MfD EMD- 30118 6M6B 
Mfd-bound RNA polymerase elongation complex — L1 state (with ATP) MfD Escherichia coli EMD-21996 6X26 [46
Mfd-bound RNA polymerase elongation complex — L2 state (with ADP) EMD-22006 6X2F 
Mfd-bound RNA polymerase elongation complex — I state EMD-22012 6X2N 
Mfd-bound RNA polymerase elongation complex — II state EMD-22039 6X43 
Mfd-bound RNA polymerase elongation complex — III state EMD-22043 6X4W 
Mfd-bound RNA polymerase elongation complex — IV state EMD-22044 6X4Y 
Mfd-bound RNA polymerase elongation complex — V state EMD-22045 6X50 
σ70-RPo2 σ-factor Klebsiella pneumoniae EMD-0001 6GH5 [47
σ70-RNA polymerase (intermediate partially loaded) complex EMD-0002 6GH6 
σ70-RNA polymerase (initially transcribing) complex EMD-4397 6GFW 
Transcription complexFamilyOrganismEMDB IDPDB IDReference
EcmrR-RPo2 MerR Escherichia coli EMD-22234 6XL5 [22
EcmrR-RPo2 (EcmrR-spacer DNA complex) EMD-22235 6XL6 
EcmrR-RPint3 with 3 nt RNA transcript EMD-22236 6XL9 
EcmrR-RPint3 with 3 nt RNA transcript (EcmrR-spacer DNA complex) EMD-22237 6XLA 
EcmrR-RPint3 with 4 nt RNA transcript EMD-22245 6XLJ 
EcmrR-RPint3 with 4 nt RNA transcript (EcmrR-spacer DNA complex) EMD-22246 6XLK 
EcmrR-RPo2 (clearer σ70 density) EMD-23291 
BmrR-RNA polymerase complex MerR Bacillus subtilis EMD-30390 7CKQ [23
CueR-RNA polymerase complex (without RNA transcript) MerR Escherichia coli EMD-22184 6XH7
6XH8 
[24
CueR-RNA polymerase complex (with RNA transcript) EMD-22185 
CueR-RNA polymerase complex (clearer σ70 density) EMD-22289 
CueR- RNA polymerase complex MerR Escherichia coli EMD-30268 6LDI [25
CueR- RNA polymerase complex (with fully duplex DNA) EMD-0874 7C17 
NanR-dimer1/DNA complex GntR Escherichia coli EMD-21652 6WFQ [26
NanR-dimer3/DNA complex EMD-21661 6WG7 
BusR-tetramer1/pAB DNA complex GntR Streptococcus agalactiae EMD-13119 7OZ3 [27
BusR-tetramer1/pAB1 DNA complex EMD-12051 7B5Y 
TraR-Eσ70 (state I) LuxR Escherichia coli EMD-0348 6N57 [28
TraR-Eσ70 (state II) EMD-0349 6N58 
TraR-Eσ70 (state III) EMD-20231 N/A 
MmfR-dimer2/DNA complex TetR Streptomyces coelicolor EMD-20781 N/A [29
Class-II CAP-TAC1 without RNA transcript (state I) CAP Escherichia coli EMD-20287 6PB5 [30
Class-II CAP-TAC1 without RNA transcript (state II) EMD-20288 6PB6 
Class-II CAP-TAC1 with RNA transcript (state II) EMD-20286 6PB4 
Class-I CAP-TAC1 CAP Escherichia coli EMD-7059 6B6F [31
Class-I CAP-TAC1 (focused map on αCTD-CAP region) EMD-7060 
Crl-EσS-RNA polymerase complex Crl Escherichia coli EMD-200090 6OMF [32
Spx-RNA polymerase complex Spx Bacillus subtilis EMD-31485 7F75 [33
WhiB7-RPo2 WhiB Mycobacterium tuberculosis EMD-22886 7KIF [34
WhiB7-RPc4 EMD-22887 7KIM 
Rgg2-short hydrophobic peptide complex Rgg Streptococcus thermophilus EMD-22341 7JI0 [35
GreB-RNA polymerase elongation complex (pre-RNA cleavage) Gre Escherichia coli EMD-4892 6RIN [36
GreB-RNA polymerase elongation complex (post-RNA cleavage) EMD-4885 6RI7 
GreB-RNA polymerase reactivated complex (before RNA extension) EMD-4882 6RH3 
CarD-RPo2 CarD Mycobacterium tuberculosis EMD-9037 6EDT [37
CarD-RNA polymerase intermediate (with 8-nt RNA transcript) EMD-9039 6EE8 
CarD-RPo2 (with corallopyronin A) EMD-9041 6EEC 
CarD-RNA polymerase holoenzyme (with corallopyronin A) EMD-9047 6M7J 
CarD-RPo2 (with Sorangicin A) CarD Mycobacterium tuberculosis EMD-21407 6VVY [38
CarD-S456LRPo2 (with Sorangicin A) EMD-21408 6VW0 
CarD-RPo2 (with Sorangicin A) (with 8-nt RNA transcript) EMD-21406 6VVX 
CarD-S456LRPo2 (with Sorangicin A) (with 8-nt RNA transcript) EMD-21409 6VVZ 
SspA-σ70-RPo2 GST Escherichia coli EMD-30307 7C97 [39
DksA-RPo2 (State I) with guanosine tetraphosphate (ppGpp) DksA Escherichia coli EMD-21881 7KHI [40
DksA-RPo2 (State II) with guanosine tetraphosphate (ppGpp) EMD-21883 7KHE 
NusG-opsEC NusG Escherichia coli EMD-7351 6C6U [41
RfaH-NusG-N-Term-opsEC RfaH EMD-7350 6C6T 
RfaH-full-length-opsEC RfaH EMD-7349 6C6S 
RNAP-HelD HelD Bacillus subtillus EMD-21921 6WVK [42
Msm HelD–RNAP complex State I HelD Mycobacterium smegmatis EMD-10996 6YXU [43
Msm HelD–RNAP complex State II EMD-11004 6YYS 
Msm HelD–RNAP complex State III EMD- 11026 6Z11 
Spt4/5-RNAP complex (with antibodies) Spt4/5 Pyrococcus furiosus EMD- 1840 N/A [44
Mfd-dependent transcription termination complex MfD Thermus thermophilus EMD- 30117 6M6A [45
Mfd-dependent transcription termination complex with ATPγS MfD EMD- 30118 6M6B 
Mfd-bound RNA polymerase elongation complex — L1 state (with ATP) MfD Escherichia coli EMD-21996 6X26 [46
Mfd-bound RNA polymerase elongation complex — L2 state (with ADP) EMD-22006 6X2F 
Mfd-bound RNA polymerase elongation complex — I state EMD-22012 6X2N 
Mfd-bound RNA polymerase elongation complex — II state EMD-22039 6X43 
Mfd-bound RNA polymerase elongation complex — III state EMD-22043 6X4W 
Mfd-bound RNA polymerase elongation complex — IV state EMD-22044 6X4Y 
Mfd-bound RNA polymerase elongation complex — V state EMD-22045 6X50 
σ70-RPo2 σ-factor Klebsiella pneumoniae EMD-0001 6GH5 [47
σ70-RNA polymerase (intermediate partially loaded) complex EMD-0002 6GH6 
σ70-RNA polymerase (initially transcribing) complex EMD-4397 6GFW 

1CAP-TAC, cAMP receptor protein-dependent transcription activation complex.

2RPo, RNA polymerase-promoter open complex.

3RPint, RNA polymerase-promoter initial transcribing complex.

3RPc, RNA polymerase-promoter closed complex.

N/A, Not available.

Transcriptional activators

Activators serve to increase transcription by binding at, or upstream, of a promoter region, where they can positively interact with and recruit RNAP to initiate transcription of target genes (Figure 1B, left panel). This process can be achieved by the activator distorting promoter DNA to facilitate RNAP binding, or by directly tethering RNAP to the promoter region. To illustrate how each regulatory mechanism enhances RNAP binding, we outline two recent cryo-EM structures, respectively below.

The cryo-EM structure of MerR family regulator EcmrR provides our first example of DNA distortion (Figure 2A). Promoters that bind this family contain an additional 2 to 3 base pairs (or non-canonical space) between the −35 and −10 elements, which prevents optimal promoter recognition by RNAP and transcription initiation [48,49]. In contrast, a canonical promoter contains a 17 base pair spacer region between the −35 and −10 elements. MerR regulators bind and twist the non-optimal DNA spacer, such that the DNA promoter elements are readily recognisable by the RNAP holoenzyme (Figure 2A, inset). While previous crystal structures of MerR regulators in the absence of RNAP reported this DNA distortion [50,51], multiple cryo-EM structures of EcmrR in complex with RNAP provided a deeper molecular understanding of this promoter remodelling [22]. Similarly, cryo-EM was also used to dissect the DNA distortion mechanism from another MerR family regulator, BmrR [23].

Recent bacterial transcription complexes solved by cryo-EM.

Figure 2.
Recent bacterial transcription complexes solved by cryo-EM.

Examples (A,B) of transcriptional activator complexes that distort DNA. (A) Dimeric EcmrR in complex with promoter DNA (cartoon inset, PDB-6XL6), remodels (58° kink) the promoter DNA to create the optimal promoter architecture for E. coli holoenzyme [σ70 (grey surface) RNAP (α2ββ′ω surface colour shown in shaded box)] to form the EcmrR-RPo (PDB-6XL5) [22]. (B) Overview of the E. coli class-1 CAP-TAC (PDB-6B6H). The cyclic adenosine 3′,5′-monophosphate receptor (CAP) protein dimer (dark blue and dark red, cartoon inset), binds its cognate DNA and αCTD of RNAP (green) to introduce three DNA kinks (33°, 55°, 21°). This results in a full 92° turn, optimally orienting the promoter DNA for σ70-RNAP to bind [31]. Examples of transcription factors that stabilise aspects of the transcription complex are shown in C and D. (C) Crl (dark orange cartoon) binds residues of the β′clamp (beige cartoon) on E. coli RNAP and alternative σ factor σS (grey cartoon) through a distinct interface (shown by spheres) in the left inset. This tethering action creates a Crl-σS-RNAP complex that binds alternate promoter DNA to form the Crl-σS-RPo complex (PDB-6OMF) shown to the right of the inset [32]. (D) WhiB7 (cyan, cartoon) is a transcriptional activator in Mycobacterium tuberculosis (Mtb) that binds an AT-rich ‘hook’ sequence of DNA (shown by arrow) and σA (cartoon inset). By binding the active Mtb RNAP holoenzyme (surfaces coloured as E. coli RNAP), it creates the WhiB7-RPo (PDB-7KIF) [34]. Examples of steric occlusion in transcriptional repressors are shown in E and F. (E) Three E. coli NanR dimers binds three GGTATA repeats to form a NanR-dimer3/DNA complex. Their close proximity allows intramolecular protein–protein interactions to stabilise the multimeric assembly (PDB-6WG7). The 70.5 kDa cryo-EM structure of dimeric NanR in complex with cognate DNA (PBD-6WFQ) [26]. (F) Streptococcus agalactiae (Stag) BusR binds palindromic promoter DNA as a tetramer to repress transcription (PDB-7OZ3) [27]. The 5′ and 3′ DNA strands have been annotated throughout.

Figure 2.
Recent bacterial transcription complexes solved by cryo-EM.

Examples (A,B) of transcriptional activator complexes that distort DNA. (A) Dimeric EcmrR in complex with promoter DNA (cartoon inset, PDB-6XL6), remodels (58° kink) the promoter DNA to create the optimal promoter architecture for E. coli holoenzyme [σ70 (grey surface) RNAP (α2ββ′ω surface colour shown in shaded box)] to form the EcmrR-RPo (PDB-6XL5) [22]. (B) Overview of the E. coli class-1 CAP-TAC (PDB-6B6H). The cyclic adenosine 3′,5′-monophosphate receptor (CAP) protein dimer (dark blue and dark red, cartoon inset), binds its cognate DNA and αCTD of RNAP (green) to introduce three DNA kinks (33°, 55°, 21°). This results in a full 92° turn, optimally orienting the promoter DNA for σ70-RNAP to bind [31]. Examples of transcription factors that stabilise aspects of the transcription complex are shown in C and D. (C) Crl (dark orange cartoon) binds residues of the β′clamp (beige cartoon) on E. coli RNAP and alternative σ factor σS (grey cartoon) through a distinct interface (shown by spheres) in the left inset. This tethering action creates a Crl-σS-RNAP complex that binds alternate promoter DNA to form the Crl-σS-RPo complex (PDB-6OMF) shown to the right of the inset [32]. (D) WhiB7 (cyan, cartoon) is a transcriptional activator in Mycobacterium tuberculosis (Mtb) that binds an AT-rich ‘hook’ sequence of DNA (shown by arrow) and σA (cartoon inset). By binding the active Mtb RNAP holoenzyme (surfaces coloured as E. coli RNAP), it creates the WhiB7-RPo (PDB-7KIF) [34]. Examples of steric occlusion in transcriptional repressors are shown in E and F. (E) Three E. coli NanR dimers binds three GGTATA repeats to form a NanR-dimer3/DNA complex. Their close proximity allows intramolecular protein–protein interactions to stabilise the multimeric assembly (PDB-6WG7). The 70.5 kDa cryo-EM structure of dimeric NanR in complex with cognate DNA (PBD-6WFQ) [26]. (F) Streptococcus agalactiae (Stag) BusR binds palindromic promoter DNA as a tetramer to repress transcription (PDB-7OZ3) [27]. The 5′ and 3′ DNA strands have been annotated throughout.

Close modal

Our second example of DNA distortion is illustrated by the role of the global transcription factor, cyclic AMP (cAMP) receptor protein (CAP) to promote transcription. Two major classes of CAP exist, each of which can activate and initiate transcription by bending DNA to optimise RNAP recruitment [52]. These classes are differentiated by their promoter site and interaction mode with RNAP; class-I CAPs bind a −61 site and interact predominantly with αCTD subunit of RNAP (Figure 1A), while class-II CAPs bind at a −41 site and interact with multiple RNAP subunits. As the CAP–RNAP interactions are small, and the full CAP–RNAP–DNA complex is dynamic, any CAP-induced conformational changes in the presence of RNAP are difficult to capture by the freeze-frame feature of crystallography [52,53]. Recently, an intact transcription activation complex, containing a class-I CAP along with RNAP was resolved by cryo-EM (Figure 2B). This structure revealed extensive remodelling of the promoter DNA (∼90° kink) induced by CAP-binding (Figure 2B, inset) to wrap upstream DNA and co-opt RNAP via αCTD binding [31]. An analogous study reported the cryo-EM structure of the class-II CAP–RNAP complex [30]. Taken together, these structures shed light into how class-I and -II CAP activation complexes assemble to activate transcription.

Alternatively, an unconventional mode of activation involving protein tethering can stabilise and activate RNAP via a DNA-independent or -dependent process. DNA-independent modes involve activators that stabilise σ-factor and RNAP association solely through protein–protein interactions. This unusual mode of activation was first hypothesised using crystallography, but many interactions were absent due to crystal packing [54,55]. However, numerous cryo-EM structures of transcriptional activators, Crl [32], RbpA/CarD [37,38], Spx [33] and SspA [39], can now detail the precise interactions and conformational changes between the σ-factor that facilitate the formation of the RNAP holoenzyme. To highlight this protein–protein tethering mode, we present the structure of Crl bound to σS and a small domain of the β′ subunit to stabilise the holoenzyme (Figure 2C). Conversely, structural insight into DNA-dependent tethering was revealed through cryo-EM structures of WhiB7, which play a role in antibiotic resistance in mycobacteria. In addition to forming protein–protein contacts with the σ-factor, WhiB7 was observed to interact with promoter DNA via an AT-hook motif [34] (Figure 2D). This was unexpected as AT-hooks are rare in bacteria, yet common in eukaryotes [56]. Thus, these structures expand our understanding of how WhiB7 serves to regulate antibiotic resistance in mycobacteria, but also unearths a novel mode of transcriptional regulation in bacteria.

Transcriptional repressors

Repressors function to sterically occlude RNAP binding to DNA by occupying a site that overlaps the -35 and -10 promoter elements to prevent σ-factor recognition, switching gene transcription off [11] (Figure 1B, right panel). We illustrate this mode of regulation below.

Recently, it was shown that Escherichia coli NanR, which regulates bacterial sialic acid metabolism [57–60], cooperatively binds a three-repeat sequence overlapping the −10 element [26]. Through cryo-EM, three NanR dimers were observed to assemble in close proximity across the promoter, where intramolecular protein–protein interactions stabilise the repressor complex (Figure 2E). This multimeric assembly is unique among reported GntR-type regulators [26]. The lower-order NanR-dimer1/DNA complex (70.5 kDa) was also resolved within the study at near atomic resolution, demonstrating the power of cryo-EM in this so-called ‘resolution revolution’ (Figure 2E, inset). Similarly, cooperative binding was also observed for the TetR-type regulator, MmfR, where two dimers bound DNA at an obtuse angle of 140° in the cryo-EM structure [29].

BusR is a transcriptional repressor that binds the c-di-AMP molecule; a vital molecule in normal cellular growth conditions and a target for antibiotic development [61]. A recent cryo-EM structure of BusR revealed how it binds bipartite DNA motifs (Figure 2F) as a tetramer and proposes a new regulator family, as the protein architecture is unlike any other transcriptional regulator described [27].

Briefly, we would be remiss not to mention how cryo-EM has had a marked impact on functionally understanding the transcription factor TraR, which functions both as an activator and repressor. Using cryo-EM, Chen et al. [28] resolved a series of structures that, alongside in vitro experiments, helped elucidate the transition of active RNAP from RPc (closed complex) to RPo (open complex) in the presence of TraR. The deconvolution of heterogeneous intermediate conformations allowed researchers to propose a mechanism, and use biochemical techniques to validate it [62]. The understanding of this conformational landscape by cryo-EM is a pivotal discovery within the field, as it has structurally illuminated the complicated and multifaceted mechanism of transcription initiation.

Having highlighted how cryo-EM has transformed our understanding of transcriptional regulation in bacteria, we now outline contemporary cryo-EM strategies for determining the structure of these dynamic macromolecular assemblies, particularly protein–DNA complexes. This overview will include the recent and ongoing developments in sample preparation and structure determination (summarised in Figure 3). Further discussion on data acquisition, image processing, and refinement are beyond the scope of this review (but are reviewed in [63,64]).

Contemporary cryo-EM strategies for solving protein–DNA complexes.

Figure 3.
Contemporary cryo-EM strategies for solving protein–DNA complexes.

Recent innovations in sample preparation (left panel), 3D reconstruction (middle panel) and model building (right panel) are illustrated. During sample preparation (left panel) there are three facets that must be optimised: sample stability (e.g. optimising the length and number of binding sites within the DNA scaffold); sample homogeneity (using size exclusion chromatography); and preferential orientation (optimising spot-to-plunge time). To combat structural heterogeneity during 3D reconstruction (middle panel) and explore dynamics, multi-body refinement [65], masked refinement, 3D variability analysis [66] or cryoDRGN, using a neural network [67] can be employed. Density map is shown in grey, with hypothetical flexibility shown in orange. To annotate density maps during model building (right panel), Haruspex [68] or Emap2sec+ [69] can be utilised. As transcription complexes are dynamic assemblies, flexible fitting tools (cryo_fit [70], ISOLDE [71], iMODFIT [72] or Namdinator [73]) are required to fit PDB components (e.g. protein and DNA) into a cryo-EM density map (shown in grey) to generate a model (shown on the bottom right of panel). Cryo-EM data, maps and models are from ref. 63 (EMD-21652, PDB-6WFQ).

Figure 3.
Contemporary cryo-EM strategies for solving protein–DNA complexes.

Recent innovations in sample preparation (left panel), 3D reconstruction (middle panel) and model building (right panel) are illustrated. During sample preparation (left panel) there are three facets that must be optimised: sample stability (e.g. optimising the length and number of binding sites within the DNA scaffold); sample homogeneity (using size exclusion chromatography); and preferential orientation (optimising spot-to-plunge time). To combat structural heterogeneity during 3D reconstruction (middle panel) and explore dynamics, multi-body refinement [65], masked refinement, 3D variability analysis [66] or cryoDRGN, using a neural network [67] can be employed. Density map is shown in grey, with hypothetical flexibility shown in orange. To annotate density maps during model building (right panel), Haruspex [68] or Emap2sec+ [69] can be utilised. As transcription complexes are dynamic assemblies, flexible fitting tools (cryo_fit [70], ISOLDE [71], iMODFIT [72] or Namdinator [73]) are required to fit PDB components (e.g. protein and DNA) into a cryo-EM density map (shown in grey) to generate a model (shown on the bottom right of panel). Cryo-EM data, maps and models are from ref. 63 (EMD-21652, PDB-6WFQ).

Close modal

Sample preparation

Sample preparation for cryo-EM is paramount. In an ideal case, upon vitrification, the sample would be free of contamination, randomly oriented and evenly distributed in a monolayer of thin ice [74,75]. In reality, two main elements must be optimised to achieve this outcome — sample preparation and grid preparation. Sample preparation involves the purification of the protein or macromolecular complex in an intact and stable manner [74,76]. Given this process isolates the sample from its cellular environment, the buffering conditions must be optimised (e.g. salt, pH) to emulate native conditions. This can be achieved systematically and in high throughput using thermal stability assays [77], such as ProteoPlex [78], which is based on differential scanning fluorimetry. Common additives such as glycerol, which mimic a crowded environment for stability is largely avoided within the cryo-EM community, because it significantly decreases contrast during data collection [76,79,80]. However, recent evidence suggests ≤20% glycerol can improve the stability of large complexes that are prone to disassembly, without compromising data quality and therefore should not be fully discounted as an additive in cryo-EM [81]. Enhanced stability can also be afforded by the addition of small effector molecules (particularly for transcriptional activators), co-factors or inhibitors that may stabilise or biochemically arrest the macromolecule in a unique functional state [22,82].

Knowledge of the specific DNA sequence that your protein–DNA system binds is critical. This includes the kinetics of your system (KD, kon, koff) and stoichiometry of the protein–DNA interaction, which can vary by the length or number of binding sites within the DNA scaffold and by protein concentration. Thus, these elements must be carefully evaluated both biochemically and biophysically to inform the optimal DNA scaffold and concentrations you should use for grid preparation.

To perform their function, protein–DNA complexes undergo dynamic conformational rearrangements across one or more subunits [18,83]. While this conformational heterogeneity can typically be tackled during image processing through independent 3D classification and masked refinement (discussed below), sources of compositional heterogeneity must be mitigated. These sources include variations in the stoichiometry of the interaction partners, partially assembled complexes or the presence of assembly intermediates [75,76,80]. At the biochemical level, these phenomena can be addressed through sample preparation procedures, mostly commonly via size exclusion chromatography to remove aggregates or unbound components and isolate a target complex [22,23,26,29]. If complexes are inherently fragile, researchers have successfully employed a fractionation technique, named GraFix, to prepare homogeneous cryo-EM samples [39,84–86]. Here, using centrifugation, a density gradient (e.g. glycerol) is combined with weak chemical fixation (e.g. glutaraldehyde), which leads to the formation of monodisperse and chemically stabilised complexes [85,87]. The use of a weak fixation reagent is advantageous as it largely favours the formation of intramolecular crosslinks, which can prevent complex dissociation. To avoid reduced contrast, the density solution is removed by buffer exchange (Zeba) spin columns [85]. If the sample is scarce or unstable following removal of the density solution, agarose fixation offers an alternative strategy [88].

Following grid preparation [reviewed in [74,89]], which is largely dependent on user expertise and experience, sample stability, heterogeneity and particle distribution can be assessed by iterative negative stain experiments or under cryogenic temperatures [74,75,79]. When vitrified, protein–DNA complexes often preferentially adhere to the air–water interface or grid support [90]. This preferred particle orientation can lead to under sampling of some structural features, sample denaturation and anisotropic resolution in the density map [91,92]. Experimentally, this can be addressed by reducing the time interval between sample application to the grid and vitrification (spot-to-plunge time) [90,93,94], the use of support layer (e.g. graphene) to sequester the complex from the air–water interface [92,95,96] or an affinity support, such as streptavidin to immobilise biotinylated molecules [97], and by data collection at a fixed tilt-angle [98]. When acquiring data at a single tilt, gold foil grids can be used to minimise beam-induced movement during imaging [99,100]. Notably, bacterial RNAP transcription complexes suffer from severe preferential orientation [101], however, this is routinely combated by the addition of the zwitterionic detergent CHAPSO during sample preparation [22,23,39,40]. Other detergents, such as β-octyl glucoside have also been used to promote random particle distribution and discourage complex dissociation [27,102,103]. Recently, molecular goniometers have been constructed using DNA origami, which enable the DNA-binding protein to bind and be precisely oriented via a sequence-specific DNA stage [104]. As proof-of-concept, this nanoscale technology was utilised to resolve the 82 kDa DNA-binding protein, BurrH [104]. Moving forward, this concept can be adaptable to other small (<100 kDa) or asymmetric protein–DNA complexes.

Structure determination

3D reconstruction and model building represent the final hurdle towards determining a structure within the cryo-EM workflow. As previously discussed, protein–DNA complexes are driven by functionally relevant conformational and compositional changes as part of their dynamic modes of action [18,83]. While this intrinsic feature poses challenges for structure determination by cryo-EM, recent innovations make it possible to study these dynamic assemblies and gain unique insights into their molecular mechanism. In this section, we highlight these key innovations, which include algorithms to visualise molecular motions and identify interacting components, along with tools for flexible fitting in cryo-EM maps by molecular dynamics.

During data acquisition, millions of snapshots across a conformational landscape are captured for the molecule of interest. This structural heterogeneity has typically been approached using various ‘3D classification’ or ‘heterogeneous refinement’ tools, implemented in cryo-EM software packages such as RELION [105], cryoSPARC [106] or cisTEM [107]. These tools effectively divide the data into a small number of independent and discrete states, each of which are assumed to be structurally homogeneous. However, in scenarios where macromolecular complexes exhibit continuous conformational transitions of single domains or motions across multiple domains, these discrete classification algorithms are ineffective for heterogeneous reconstruction as they often omit functionally relevant or transient states [83]. As a result, more focused approaches have evolved to deal with continuous flexibility in cryo-EM data. These include, multi-body refinement, which uses a discrete number of independently moving, rigid bodies to model the dynamics within a protein complex and improve density maps of flexible regions [65,108]. Implemented in RELION, this multi-body approach has been utilised to characterise the TraR-induced structural changes in E. coli RNAP to regulate transcription. An analogous strategy, masked 3D refinement, can also employed to combat structural heterogeneity by applying a mask that excludes contents outside a region of interest, local resolution can be improved [26]. To avoid introducing artefacts or overfitting, generated mask are often low-pass filtered with soft edges [83]. 3D Variability Analysis (3DVA), available in cryoSPARC, is an algorithm that fits a linear subspace model to visualise molecular motions of macromolecules at high resolution [66]. 3DVA has since been utilised to resolve the dynamic interaction between RNAP and the transcription factor, HelD from Bacillus subtilis [42]. A similar tool, cryoDRGN (http://cryodrgn.csail.mit.edu), utilises deep neural networks to reconstruct density maps that model both discrete compositional heterogeneity and continuous conformational changes [67]. While these methods primarily combat motions across multiple domains, they can provide a potential trajectory to additionally refine single domain motions.

Following 3D reconstruction, atomic models provide the basis to structurally and functionally interpret the cryo-EM density maps. Most commonly, this process involves the input of existing X-ray or NMR structures from the PDB [22,26–29], however in their absence, the neural network-based, structure prediction programs, AlphaFold [109] or RoseTTA [110] now offer an alternative. To guide this initial structure into the target density map, various flexible fitting tools, such as cryo_fit within phenix [70], ISOLDE [71] or iMODFIT [72] in Chimera, Namdinator [73] among others [29,111] can be used to accommodate conformational heterogeneity by utilising molecular dynamics simulations or normal mode analysis [112]. In scenarios where it is difficult to annotate protein and DNA in the density maps, the recent tools Haruspex [68] and Emap2sec+ [69] can be employed to detect these structures in high-resolution maps (>4 Å) and lower resolution maps 5–10 Å, respectively using neural networks. Aside from flexible fitting, atomic models can also be constructed de novo when the resolution is better than 3.5 Å [113]. Although the Rosetta refinement strategy can be implemented as a de novo model-building approach for cryo-EM maps at 3–5 Å [114]. For tools that permit further refinement and validation of these models, we direct readers to a more detailed review [115].

The advent of cryo-EM, driven by advances in hardware and data processing, has revolutionised our understanding of transcription regulation in bacteria. This is evidenced by the rapidly expanding structural repertoire of bacterial transcription complexes over the past two years, with >65 resolved by cryo-EM to date. As illustrated within this review, these structures have shed light on unique conformational changes not seen in previous crystallographic studies, which include: promoter remodelling to stabilise intermediate complexes of transcription initiation [22,23,37]; insights into the conformational plasticity during the transition from transcription initiation to elongation [22,28]; the cooperative assembly of the transcriptional repressor, NanR [26]; the effector-induced reconfiguration of BusR to bind a bipartite DNA motif [27]; and the interaction of Crl and WhiB7 with RNAP to tether the σ-factors that they regulate through protein–protein or protein–DNA interactions, respectively [32,34]. However, despite these advancements reviewed here, gaps in our knowledge remain.

To yield a comprehensive model of transcription regulation for a given bacterial system, researchers must employ an integrative structural, biophysical and biochemical approach [116]. While cryo-EM is powerful at resolving large, conformationally dynamic assemblies, which are difficult to be captured by crystallography due to crystal packing, there are still limitations in size and resolution. In contrast, crystallography is better suited to yield atomic coordinates of macromolecules under 100–200 kDa and can attain higher resolutions (<2 Å) as molecules are constrained to a crystal lattice [117]. It is now routine to dock these smaller, more ordered structures into cryo-EM maps using flexible fitting to accommodate minor conformational differences and allow a more in-depth interpretation of the model [112]. Hence, crystallography remains a useful tool in structural biology to complement cryo-EM studies.

In keeping with the theme of an integrative approach, small angle X-ray scattering (SAXS) can ‘observe’ the conformational landscape of transcription complexes in solution and compare this with the cryo-EM model to evaluate biological relevance [118]. Likewise, the stoichiometry and molecular mass of protein–DNA complexes can be determined in solution, using an emerging, label-free analytical ultracentrifugation method that features multi-wavelength detection to deconvolute the spectral signals of protein and DNA based on their unique optical properties [26,119,120]. Single molecule mass photometry can also be employed as a tool to determine molecular mass of complexes in solution [121]. Other tools for characterising protein–DNA interactions (reviewed in [122]), include cryo-electron tomography, NMR, mass spectrometry/cross linking, Förster resonance energy transfer (FRET) and hydrogen–deuterium exchange.

Moving forward, future structural endeavours using cryo-EM will no doubt build on the contributions highlighted here to deconvolute the complicated and often multifaceted molecular mechanisms of transcriptional regulation in bacteria. Alongside regular hardware and software improvements, we anticipate machine learning methods, such as cryoDRGN [67], along with the prospect of time-resolved cryo-EM [123,124] will enable researchers to explore more transient intermediate complexes and thus gain a deeper understanding of the molecular choreography that drives these regulatory mechanisms.

  • Transcription complexes are dynamic assembles whose function is often intertwined with their many structural configurations. The precise choreography and nature of these motions remains incompletely understood. This knowledge is essential to understand the molecular mechanisms of transcription regulation in bacteria.

  • Fuelled by the ‘resolution revolution’, cryo-EM has emerged to provide researchers a means of probing these larger and structurally heterogeneous macromolecules, which are sensitive to crystallisation. To date, these studies have contributed >65 complex structures and provided unprecedented insights into bacterial transcription regulation.

  • The prospect of temporally linking the dynamic nature of transcription complexes remains of immense interest. Excitingly, the evolution of machine learning and time-resolved cryo-EM applications represent a future avenue to explore these transient intermediates.

The authors declare that there are no competing interests associated with the manuscript.

R.C.J.D. acknowledges the following for funding support, in part: 1) the New Zealand Royal Society Marsden Fund (contract UOC1506); 2) a Ministry of Business, Innovation and Employment Smart Ideas grant (contract UOCX1706); and 3) the Biomolecular Interactions Centre (University of Canterbury).

Open access for this article was enabled by the participation of University of Melbourne in an all-inclusive Read & Publish pilot with Portland Press and the Biochemical Society under a transformative agreement with CAUL.

D.M.W., R.C.J.D., and C.R.H. reviewed the literature, wrote the manuscript and created the figures.

We thank the editors from the Biochemical Society for the invitation to contribute.

3DVA

3D Variability Analysis

CAP

cyclic AMP (cAMP) receptor protein

CryoDRGN

Cryo-Deep Reconstructing Generative Networks

Cryo-EM

Cryo-electron microscopy

CTD/NTD

C-terminal/ N-terminal domain

DNA

Deoxyribonucleic acid

EMDB

Electron Microscopy Data Bank

kDa

Kilodalton

PDB

Protein Data Bank

RNAP

RNA polymerase

RPc

RNA polymerase closed complex

RPo

RNA polymerase open complex

σ-factor

Sigma factor

1
Burgess
,
R.R.
(
1969
)
Separation and characterization of the subunits of ribonucleic acid polymerase
.
J. Biol. Chem.
244
,
6168
6176
2
Vassylyev
,
D.G.
,
Sekine
,
S.I.
,
Laptenko
,
O.
,
Lee
,
J.
,
Vassylyeva
,
M.N.
,
Borukhov
,
S.
et al (
2002
)
Crystal structure of a bacterial RNA polymerase holoenzyme at 2.6 Å resolution
.
Nature
417
,
712
719
3
Murakami
,
K.S.
(
2002
)
Structural basis of transcription initiation: an RNA polymerase holoenzyme-DNA complex
.
Science
296
,
1285
1290
4
Feklístov
,
A.
,
Sharon
,
B.D.
,
Darst
,
S.A.
and
Gross
,
C.A.
(
2014
)
Bacterial sigma factors: a historical, structural, and genomic perspective
.
Annu. Rev. Microbiol.
68
,
357
376
5
Campbell
,
E.A.
,
Muzzin
,
O.
,
Chlenov
,
M.
,
Sun
,
J.L.
,
Olson
,
C.A.
,
Weinman
,
O.
et al (
2002
)
Structure of the bacterial RNA polymerase promoter specificity σ subunit
.
Mol. Cell
9
,
527
539
6
Barne
,
K.A.
,
Bown
,
J.A.
,
Busby
,
S.J.W.
and
Minchin
,
S.D.
(
1997
)
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.
16
,
4034
4040
7
Feklistov
,
A.
,
Barinova
,
N.
,
Sevostyanova
,
A.
,
Heyduk
,
E.
,
Bass
,
I.
,
Vvedenskaya
,
I.
et al (
2006
)
A basal promoter element recognized by free RNA polymerase σ subunit determines promoter recognition by RNA polymerase holoenzyme
.
Mol. Cell
23
,
97
107
8
Haugen
,
S.P.
,
Berkmen
,
M.B.
,
Ross
,
W.
,
Gaal
,
T.
,
Ward
,
C.
and
Gourse
,
R.L.
(
2006
)
rRNA promoter regulation by nonoptimal binding of σ region 1.2: an additional recognition element for RNA polymerase
.
Cell
125
,
1069
1082
9
Estrem
,
S.T.
,
Ross
,
W.
,
Gaal
,
T.
,
Chen
,
Z.W.
,
Niu
,
W.
,
Ebright
,
R.H.
et al (
1999
)
Bacterial promoter architecture: subsite structure of UP elements and interactions with the carboxy-terminal domain of the RNA polymerase alpha subunit
.
Genes Dev.
13
,
2134
2147
10
Browning
,
D.F.
and
Busby
,
S.J.W.
(
2016
)
Local and global regulation of transcription initiation in bacteria
.
Nat. Rev. Microbiol.
14
,
638
650
11
Browning
,
D.F.
and
Busby
,
S.J.
(
2004
)
The regulation of bacterial transcription initiation
.
Nat. Rev. Microbiol.
2
,
57
65
12
Seshasayee
,
A.S.N.
,
Sivaraman
,
K.
and
Luscombe
,
N.M.
(
2011
) An Overview of Prokaryotic Transcription Factors. In
A Handbook of Transcription Factors
(
Hughes
,
T.R.
, ed.), pp.
7
23
,
Springer Netherlands
,
Dordrecht
13
Perez-Rueda
,
E.
,
Hernandez-Guerrero
,
R.
,
Martinez-Nuñez
,
M.A.
,
Armenta-Medina
,
D.
,
Sanchez
,
I.
and
Ibarra
,
J.A.
(
2018
)
Abundance, diversity and domain architecture variability in prokaryotic DNA-binding transcription factors
.
PLoS ONE
13
,
e0195332
14
Weisberg
,
R.A.
(
2004
)
A genetic switch: phage lambda revisited. Third Edition. By Mark Ptashne
.
Q. Rev. Biol.
79
,
427
428
15
Ruff
,
E.
,
Record
,
M.
and
Artsimovitch
,
I.
(
2015
)
Initial events in bacterial transcription initiation
.
Biomolecules
5
,
1035
1062
16
Lee
,
D.J.
,
Minchin
,
S.D.
and
Busby
,
S.J.
(
2012
)
Activating transcription in bacteria
.
Annu. Rev. Microbiol.
66
,
125
152
17
Helmann
,
J.D.
(
2009
)
RNA polymerase: a nexus of gene regulation
.
Methods (San Diego, Calif)
47
,
1
5
18
Hanske
,
J.
,
Sadian
,
Y.
and
Müller
,
C.W.
(
2018
)
The cryo-EM resolution revolution and transcription complexes
.
Curr. Opin. Struct. Biol.
52
,
8
15
19
Nogales
,
E.
and
Scheres
,
S.H.
(
2015
)
Cryo-EM: a unique tool for the visualization of macromolecular complexity
.
Mol. Cell
58
,
677
689
20
Nogales
,
E.
,
Louder
,
R.K.
and
He
,
Y.
(
2016
)
Cryo-EM in the study of challenging systems: the human transcription pre-initiation complex
.
Curr. Opin. Struct. Biol.
40
,
120
127
21
Benjin
,
X.
and
Ling
,
L.
(
2020
)
Developments, applications, and prospects of cryo-electron microscopy
.
Protein Sci.
29
,
872
882
22
Yang
,
Y.
,
Liu
,
C.
,
Zhou
,
W.
,
Shi
,
W.
,
Chen
,
M.
,
Zhang
,
B.
et al (
2021
)
Structural visualization of transcription activated by a multidrug-sensing MerR family regulator
.
Nat. Commun.
12
,
2702
23
Fang
,
C.
,
Li
,
L.
,
Zhao
,
Y.
,
Wu
,
X.
,
Philips
,
S.J.
,
You
,
L.
et al (
2020
)
The bacterial multidrug resistance regulator BmrR distorts promoter DNA to activate transcription
.
Nat. Commun.
11
,
6284
24
Shi
,
W.
,
Zhang
,
B.
,
Jiang
,
Y.
,
Liu
,
C.
,
Zhou
,
W.
,
Chen
,
M.
et al (
2021
)
Structural basis of copper-efflux-regulator-dependent transcription activation
.
iScience
24
,
102449
25
Fang
,
C.
,
Philips
,
S.J.
,
Wu
,
X.
,
Chen
,
K.
,
Shi
,
J.
,
Shen
,
L.
et al (
2021
)
Cuer activates transcription through a DNA distortion mechanism
.
Nat. Chem. Biol.
17
,
57
64
26
Horne
,
C.R.
,
Venugopal
,
H.
,
Panjikar
,
S.
,
Wood
,
D.M.
,
Henrickson
,
A.
,
Brookes
,
E.
et al (
2021
)
Mechanism of NanR gene repression and allosteric induction of bacterial sialic acid metabolism
.
Nat. Commun.
12
,
1988
27
Bandera
,
A.M.
,
Bartho
,
J.
,
Lammens
,
K.
,
Drexler
,
D.J.
,
Kleinschwärzer
,
J.
,
Hopfner
,
K.P.
et al (
2021
)
Busr senses bipartite DNA binding motifs by a unique molecular ruler architecture
.
Nucleic Acids Res.
49
,
10166
10177
28
Chen
,
J.
,
Gopalkrishnan
,
S.
,
Chiu
,
C.
,
Chen
,
A.Y.
,
Campbell
,
E.A.
,
Gourse
,
R.L.
et al (
2019
)
E. coli TraR allosterically regulates transcription initiation by altering RNA polymerase conformation
.
eLife
8
,
e49375
29
Zhou
,
S.
,
Bhukya
,
H.
,
Malet
,
N.
,
Harrison
,
P.J.
,
Rea
,
D.
,
Belousoff
,
M.J.
et al (
2021
)
Molecular basis for control of antibiotic production by a bacterial hormone
.
Nature
590
,
463
467
30
Shi
,
W.
,
Jiang
,
Y.
,
Deng
,
Y.
,
Dong
,
Z.
and
Liu
,
B.
(
2020
)
Visualization of two architectures in class-II CAP-dependent transcription activation
.
PLoS Biol.
18
,
e3000706
31
Liu
,
B.
,
Hong
,
C.
,
Huang
,
R.K.
,
Yu
,
Z.
and
Steitz
,
T.A.
(
2017
)
Structural basis of bacterial transcription activation
.
Science
358
,
947
951
32
Cartagena
,
A.J.
,
Banta
,
A.B.
,
Sathyan
,
N.
,
Ross
,
W.
,
Gourse
,
R.L.
,
Campbell
,
E.A.
et al (
2019
)
Structural basis for transcription activation by Crl through tethering of σ(S) and RNA polymerase
.
Proc. Natl Acad. Sci. U.S.A.
116
,
18923
7
33
Shi
,
J.
,
Li
,
F.
,
Wen
,
A.
,
Yu
,
L.
,
Wang
,
L.
,
Wang
,
F.
et al (
2021
)
Structural basis of transcription activation by the global regulator Spx
.
Nucleic Acids Res.
49
,
10756
10769
34
Lilic
,
M.
,
Darst
,
S.A.
and
Campbell
,
E.A.
(
2021
)
Structural basis of transcriptional activation by the Mycobacterium tuberculosis intrinsic antibiotic-resistance transcription factor WhiB7
.
Mol. Cell
81
,
2875
2886.e5
35
Capodagli
,
G.C.
,
Tylor
,
K.M.
,
Kaelber
,
J.T.
,
Petrou
,
V.I.
,
Federle
,
M.J.
and
Neiditch
,
M.B.
(
2020
)
Structure-function studies of Rgg binding to pheromones and target promoters reveal a model of transcription factor interplay
.
Proc. Natl Acad. Sci. U.S.A.
117
,
24494
24502
36
Abdelkareem
,
M.
,
Saint-André
,
C.
,
Takacs
,
M.
,
Papai
,
G.
,
Crucifix
,
C.
,
Guo
,
X.
et al (
2019
)
Structural basis of transcription: RNA polymerase backtracking and Its reactivation
.
Mol. Cell
75
,
298
309.e4
37
Boyaci
,
H.
,
Chen
,
J.
,
Jansen
,
R.
,
Darst
,
S.A.
and
Campbell
,
E.A.
(
2019
)
Structures of an RNA polymerase promoter melting intermediate elucidate DNA unwinding
.
Nature
565
,
382
385
38
Lilic
,
M.
,
Chen
,
J.
,
Boyaci
,
H.
,
Braffman
,
N.
,
Hubin
,
E.A.
,
Herrmann
,
J.
et al (
2020
)
The antibiotic sorangicin A inhibits promoter DNA unwinding in a Mycobacterium tuberculosis rifampicin-resistant RNA polymerase
.
Proc. Natl Acad. Sci. U.S.A.
117
,
30423
30432
39
Wang
,
F.
,
Shi
,
J.
,
He
,
D.
,
Tong
,
B.
,
Zhang
,
C.
,
Wen
,
A.
et al (
2020
)
Structural basis for transcription inhibition by E. coli SspA
.
Nucleic Acids Res.
48
,
9931
9942
40
Shin
,
Y.
,
Qayyum
,
M.Z.
,
Pupov
,
D.
,
Esyunina
,
D.
,
Kulbachinskiy
,
A.
and
Murakami
,
K.S.
(
2021
)
Structural basis of ribosomal RNA transcription regulation
.
Nat. Commun.
12
,
528
41
Kang
,
J.Y.
,
Mooney
,
R.A.
,
Nedialkov
,
Y.
,
Saba
,
J.
,
Mishanina
,
T.V.
,
Artsimovitch
,
I.
et al (
2018
)
Structural basis for transcript elongation control by nusG family universal regulators
.
Cell
173
,
1650
1662.e14
42
Newing
,
T.P.
,
Oakley
,
A.J.
,
Miller
,
M.
,
Dawson
,
C.J.
,
Brown
,
S.H.J.
,
Bouwer
,
J.C.
et al (
2020
)
Molecular basis for RNA polymerase-dependent transcription complex recycling by the helicase-like motor protein HelD
.
Nat. Commun.
11
,
6420
43
Kouba
,
T.
,
Koval’
,
T.
,
Sudzinová
,
P.
,
Pospíšil
,
J.
,
Brezovská
,
B.
,
Hnilicová
,
J.
et al (
2020
)
Mycobacterial HelD is a nucleic acids-clearing factor for RNA polymerase
.
Nat. Commun.
11
,
6419
44
Klein
,
B.J.
,
Bose
,
D.
,
Baker
,
K.J.
,
Yusoff
,
Z.M.
,
Zhang
,
X.
and
Murakami
,
K.S.
(
2010
)
RNA polymerase and transcription elongation factor Spt4/5 complex structure
.
Proc. Natl Acad. Sci. U.S.A.
108
,
546
550
45
Shi
,
J.
,
Wen
,
A.
,
Zhao
,
M.
,
Jin
,
S.
,
You
,
L.
,
Shi
,
Y.
et al (
2020
)
Structural basis of Mfd-dependent transcription termination
.
Nucleic Acids Res.
48
,
11762
11772
46
Kang
,
J.Y.
,
Llewellyn
,
E.
,
Chen
,
J.
,
Olinares
,
P.D.B.
,
Brewer
,
J.
,
Chait
,
B.T.
et al (
2021
)
Structural basis for transcription complex disruption by the Mfd translocase
.
eLife
10
,
e62117
47
Glyde
,
R.
,
Ye
,
F.
,
Jovanovic
,
M.
,
Kotta-Loizou
,
I.
,
Buck
,
M.
and
Zhang
,
X.
(
2018
)
Structures of bacterial RNA polymerase complexes reveal the mechanism of DNA loading and transcription initiation
.
Mol. Cell
70
,
1111
1120.e3
48
Hobman
,
J.L.
(
2007
)
Merr family transcription activators: similar designs, different specificities
.
Mol. Microbiol.
63
,
1275
1278
49
Julian
,
D.J.
,
Kershaw
,
C.J.
,
Brown
,
N.L.
and
Hobman
,
J.L.
(
2009
)
Transcriptional activation of MerR family promoters in Cupriavidus metallidurans CH34
.
Antonie Van Leeuwenhoek
96
,
149
159
50
Heldwein
,
E.E.
and
Brennan
,
R.G.
(
2001
)
Crystal structure of the transcription activator BmrR bound to DNA and a drug
.
Nature
409
,
378
382
51
Philips
,
S.J.
,
Canalizo-Hernandez
,
M.
,
Yildirim
,
I.
,
Schatz
,
G.C.
,
Mondragón
,
A.
and
O'Halloran
,
T.V.
(
2015
)
TRANSCRIPTION. Allosteric transcriptional regulation via changes in the overall topology of the core promoter
.
Science
349
,
877
881
52
Schultz
,
S.C.
,
Shields
,
G.C.
and
Steitz
,
T.A.
(
1990
)
Crystallization of Escherichia coli catabolite gene activator protein with its DNA binding site
.
J. Mol. Biol.
213
,
159
166
53
Benoff
,
B.
,
Yang
,
H.
,
Lawson
,
C.L.
,
Parkinson
,
G.
,
Liu
,
J.
,
Blatter
,
E.
et al (
2002
)
Structural basis of transcription activation: the CAP-alpha CTD-DNA complex
.
Science
297
,
1562
1566
54
Hubin
,
E.A.
,
Fay
,
A.
,
Xu
,
C.
,
Bean
,
J.M.
,
Saecker
,
R.M.
,
Glickman
,
M.S.
et al (
2017
)
Structure and function of the mycobacterial transcription initiation complex with the essential regulator rbpA
.
eLife
6
,
e22520
55
Liu
,
B.
,
Zuo
,
Y.
and
Steitz
,
T.A.
(
2016
)
Structures of E. coli σS-transcription initiation complexes provide new insights into polymerase mechanism
.
Proc. Natl Acad. Sci. U.S.A.
113
,
4051
4056
56
Morris
,
R.P.
,
Nguyen
,
L.
,
Gatfield
,
J.
,
Visconti
,
K.
,
Nguyen
,
K.
,
Schnappinger
,
D.
et al (
2005
)
Ancestral antibiotic resistance in Mycobacterium tuberculosis
.
Proc. Natl Acad. Sci. U.S.A.
102
,
12200
12205
57
Coombes
,
D.
,
Davies
,
J.S.
,
Newton-Vesty
,
M.C.
,
Horne
,
C.R.
,
Setty
,
T.G.
,
Subramanian
,
R.
et al (
2020
)
The basis for non-canonical ROK family function in the N-acetylmannosamine kinase from the pathogen Staphylococcus aureus
.
J. Biol. Chem.
295
,
3301
3315
58
North
,
R.A.
,
Horne
,
C.R.
,
Davies
,
J.S.
,
Remus
,
D.M.
,
Muscroft-Taylor
,
A.C.
,
Goyal
,
P.
et al (
2018
)
“Just a spoonful of sugar…”: import of sialic acid across bacterial cell membranes
.
Biophys. Rev.
10
,
219
227
59
Currie
,
M.J.
,
Manjunath
,
L.
,
Horne
,
C.R.
,
Rendle
,
P.M.
,
Subramanian
,
R.
,
Friemann
,
R.
et al (
2021
)
N-Acetylmannosamine-6-phosphate 2-epimerase uses a novel substrate-assisted mechanism to catalyze amino sugar epimerization
.
J. Biol. Chem.
297
,
101113
60
Horne
,
C.R.
,
Kind
,
L.
,
Davies
,
J.S.
and
Dobson
,
R.C.J.
(
2020
)
On the structure and function of Escherichia coli YjhC: an oxidoreductase involved in bacterial sialic acid metabolism
.
Proteins
88
,
654
668
61
de Lencastre
,
H.
,
Dengler
,
V.
,
McCallum
,
N.
,
Kiefer
,
P.
,
Christen
,
P.
,
Patrignani
,
A.
et al (
2013
)
Mutation in the C-Di-AMP cyclase dacA affects fitness and resistance of methicillin resistant Staphylococcus aureus
.
PLoS ONE.
8
,
e73512
62
Chen
,
J.
,
Chiu
,
C.
,
Gopalkrishnan
,
S.
,
Chen
,
A.Y.
,
Olinares
,
P.D.B.
,
Saecker
,
R.M.
et al (
2020
)
Stepwise promoter melting by bacterial RNA polymerase
.
Mol. Cell
78
,
275
288.e6
63
Lyumkis
,
D.
(
2019
)
Challenges and opportunities in cryo-EM single-particle analysis
.
J. Biol. Chem.
294
,
5181
5197
64
Costa
,
T.R.D.
,
Ignatiou
,
A.
and
Orlova
,
E.V.
(
2017
)
Structural analysis of protein complexes by cryo electron microscopy
.
Methods Mol. Biol.
1615
,
377
413
65
Nakane
,
T.
,
Kimanius
,
D.
,
Lindahl
,
E.
and
Scheres
,
S.H.
(
2018
)
Characterisation of molecular motions in cryo-EM single-particle data by multi-body refinement in RELION
.
eLife
7
,
e36861
66
Punjani
,
A.
and
Fleet
,
D.J.
(
2021
)
3D variability analysis: resolving continuous flexibility and discrete heterogeneity from single particle cryo-EM
.
J. Struct. Biol.
213
,
107702
67
Zhong
,
E.D.
,
Bepler
,
T.
,
Berger
,
B.
and
Davis
,
J.H.
(
2021
)
CryoDRGN: reconstruction of heterogeneous cryo-EM structures using neural networks
.
Nat. Methods
18
,
176
185
68
Mostosi
,
P.
,
Schindelin
,
H.
,
Kollmannsberger
,
P.
and
Thorn
,
A.
(
2020
)
Haruspex: a neural network for the automatic identification of oligonucleotides and protein secondary structure in cryo-electron microscopy maps
.
Angew. Chem. Int. Ed. Engl.
59
,
14788
14795
69
Wang
,
X.
,
Alnabati
,
E.
,
Aderinwale
,
T.W.
,
Maddhuri Venkata Subramaniya
,
S.R.
,
Terashi
,
G.
and
Kihara
,
D.
(
2021
)
Detecting protein and DNA/RNA structures in cryo-EM maps of intermediate resolution using deep learning
.
Nat. Commun.
12
,
2302
70
Kim
,
D.N.
,
Moriarty
,
N.W.
,
Kirmizialtin
,
S.
,
Afonine
,
P.V.
,
Poon
,
B.
,
Sobolev
,
O.V.
et al (
2019
)
Cryo_fit: democratization of flexible fitting for cryo-EM
.
J. Struct. Biol.
208
,
1
6
71
Croll
,
T.I.
(
2018
)
ISOLDE: a physically realistic environment for model building into low-resolution electron-density maps
.
Acta Crystallogr. D Struct. Biol.
74
,
519
530
72
Lopéz-Blanco
,
J.R.
and
Chacón
,
P.
(
2013
)
iMODFIT: efficient and robust flexible fitting based on vibrational analysis in internal coordinates
.
J. Struct. Biol.
184
,
261
270
73
Kidmose
,
R.T.
,
Juhl
,
J.
,
Nissen
,
P.
,
Boesen
,
T.
,
Karlsen
,
J.L.
and
Pedersen
,
B.P.
(
2019
)
Namdinator - automatic molecular dynamics flexible fitting of structural models into cryo-EM and crystallography experimental maps
.
IUCrJ.
6
,
526
531
74
Weissenberger
,
G.
,
Henderikx
,
R.J.M.
and
Peters
,
P.J.
(
2021
)
Understanding the invisible hands of sample preparation for cryo-EM
.
Nat. Methods
18
,
463
471
75
Cheng
,
Y.
,
Grigorieff
,
N.
,
Penczek
,
P.A.
and
Walz
,
T.
(
2015
)
A primer to single-particle cryo-electron microscopy
.
Cell
161
,
438
449
76
Stark
,
H.
and
Chari
,
A.
(
2016
)
Sample preparation of biological macromolecular assemblies for the determination of high-resolution structures by cryo-electron microscopy
.
Microscopy (Oxf)
65
,
23
34
77
Seabrook
,
S.A.
and
Newman
,
J.
(
2013
)
High-throughput thermal scanning for protein stability: making a good technique more robust
.
ACS Comb. Sci.
15
,
387
392
78
Chari
,
A.
,
Haselbach
,
D.
,
Kirves
,
J.M.
,
Ohmer
,
J.
,
Paknia
,
E.
,
Fischer
,
N.
et al (
2015
)
Proteoplex: stability optimization of macromolecular complexes by sparse-matrix screening of chemical space
.
Nat. Methods
12
,
859
865
79
Drulyte
,
I.
,
Johnson
,
R.M.
,
Hesketh
,
E.L.
,
Hurdiss
,
D.L.
,
Scarff
,
C.A.
,
Porav
,
S.A.
et al (
2018
)
Approaches to altering particle distributions in cryo-electron microscopy sample preparation
.
Acta Crystallogr. D Struct. Biol.
74
,
560
571
80
Thompson
,
R.F.
,
Walker
,
M.
,
Siebert
,
C.A.
,
Muench
,
S.P.
and
Ranson
,
N.A.
(
2016
)
An introduction to sample preparation and imaging by cryo-electron microscopy for structural biology
.
Methods
100
,
3
15
81
Basanta
,
B.
,
Hirschi
,
M.M.
,
Grotjahn
,
D.A.
and
Lander
,
G.C.
(
2021
)
A case for glycerol as an acceptable additive for single particle cryoEM samples
.
bioRxiv
1
10
82
Nwanochie
,
E.
and
Uversky
,
V.N.
(
2019
)
Structure determination by single-particle cryo-electron microscopy: only the Sky (and intrinsic disorder) is the limit
.
Int. J. Mol. Sci.
20
,
4186
83
Serna
,
M.
(
2019
)
Hands on methods for high resolution cryo-Electron microscopy structures of heterogeneous macromolecular complexes
.
Front. Mol. Biosci.
6
,
33
84
De Wijngaert
,
B.
,
Sultana
,
S.
,
Singh
,
A.
,
Dharia
,
C.
,
Vanbuel
,
H.
,
Shen
,
J.
et al (
2021
)
Cryo-EM structures reveal transcription initiation steps by yeast mitochondrial RNA polymerase
.
Mol. Cell
81
,
268
280.e5
85
Kastner
,
B.
,
Fischer
,
N.
,
Golas
,
M.M.
,
Sander
,
B.
,
Dube
,
P.
,
Boehringer
,
D.
et al (
2008
)
Grafix: sample preparation for single-particle electron cryomicroscopy
.
Nat. Methods
5
,
53
55
86
Narayanan
,
A.
,
Vago
,
F.S.
,
Li
,
K.
,
Qayyum
,
M.Z.
,
Yernool
,
D.
,
Jiang
,
W.
et al (
2018
)
Cryo-EM structure of Escherichia coli σ(70) RNA polymerase and promoter DNA complex revealed a role of σ non-conserved region during the open complex formation
.
J. Biol. Chem.
293
,
7367
7375
87
Stark
,
H.
(
2010
)
Grafix: stabilization of fragile macromolecular complexes for single particle cryo-EM
.
Methods Enzymol.
481
,
109
126
88
Adamus
,
K.
,
Le
,
S.N.
,
Elmlund
,
H.
,
Boudes
,
M.
and
Elmlund
,
D.
(
2019
)
Agarfix: simple and accessible stabilization of challenging single-particle cryo-EM specimens through crosslinking in a matrix of agar
.
J. Struct. Biol.
207
,
327
331
89
Goswami
,
P.
,
Locke
,
J.
and
Costa
,
A.
(
2018
)
Preparing frozen-hydrated protein-Nucleic acid assemblies for high-resolution cryo-EM imaging
.
Methods Mol. Biol.
1814
,
287
296
90
Noble
,
A.J.
,
Wei
,
H.
,
Dandey
,
V.P.
,
Zhang
,
Z.
,
Tan
,
Y.Z.
,
Potter
,
C.S.
et al (
2018
)
Reducing effects of particle adsorption to the air-water interface in cryo-EM
.
Nat. Methods
15
,
793
795
91
Glaeser
,
R.M.
and
Han
,
B.G.
(
2017
)
Opinion: hazards faced by macromolecules when confined to thin aqueous films
.
Biophys. Rep.
3
,
1
7
92
D'Imprima
,
E.
,
Floris
,
D.
,
Joppe
,
M.
,
Sánchez
,
R.
,
Grininger
,
M.
and
Kühlbrandt
,
W.
(
2019
)
Protein denaturation at the air-water interface and how to prevent it
.
eLife
8
,
e42747
93
Ashtiani
,
D.
,
Venugopal
,
H.
,
Belousoff
,
M.
,
Spicer
,
B.
,
Mak
,
J.
,
Neild
,
A.
et al (
2018
)
Delivery of femtolitre droplets using surface acoustic wave based atomisation for cryo-EM grid preparation
.
J. Struct. Biol.
203
,
94
101
94
Dandey
,
V.P.
,
Wei
,
H.
,
Zhang
,
Z.
,
Tan
,
Y.Z.
,
Acharya
,
P.
,
Eng
,
E.T.
et al (
2018
)
Spotiton: new features and applications
.
J. Struct. Biol.
202
,
161
169
95
Liu
,
N.
,
Zhang
,
J.
,
Chen
,
Y.
,
Liu
,
C.
,
Zhang
,
X.
,
Xu
,
K.
et al (
2019
)
Bioactive functionalized monolayer graphene for high-resolution cryo-Electron microscopy
.
J. Am. Chem. Soc.
141
,
4016
4025
96
Wang
,
F.
,
Liu
,
Y.
,
Yu
,
Z.
,
Li
,
S.
,
Feng
,
S.
,
Cheng
,
Y.
et al (
2020
)
General and robust covalently linked graphene oxide affinity grids for high-resolution cryo-EM
.
Proc. Natl Acad. Sci. U.S.A.
117
,
24269
24273
97
Glaeser
,
R.M.
,
Han
,
B.-G.
,
Watson
,
Z.
,
Ward
,
F.
and
Cate
,
J.H.D.
(
2018
)
Streptavidin affinity grids for cryo-EM
.
Biophys. J.
114
,
163a
98
Tan
,
Y.Z.
,
Baldwin
,
P.R.
,
Davis
,
J.H.
,
Williamson
,
J.R.
,
Potter
,
C.S.
,
Carragher
,
B.
et al (
2017
)
Addressing preferred specimen orientation in single-particle cryo-EM through tilting
.
Nat. Methods
14
,
793
796
99
Russo
,
C.J.
and
Passmore
,
L.A.
(
2014
)
Electron microscopy: ultrastable gold substrates for electron cryomicroscopy
.
Science
346
,
1377
1380
100
Naydenova
,
K.
,
Peet
,
M.J.
and
Russo
,
C.J.
(
2019
)
Multifunctional graphene supports for electron cryomicroscopy
.
Proc. Natl Acad. Sci. U.S.A.
116
,
11718
11724
101
Chen
,
J.
,
Noble
,
A.J.
,
Kang
,
J.Y.
and
Darst
,
S.A.
(
2019
)
Eliminating effects of particle adsorption to the air/water interface in single-particle cryo-electron microscopy: bacterial RNA polymerase and CHAPSO
.
J. Struct. Biol. X.
1
,
100005
102
Ha
,
B.
,
Larsen
,
K.P.
,
Zhang
,
J.
,
Fu
,
Z.
,
Montabana
,
E.
,
Jackson
,
L.N.
et al (
2021
)
High-resolution view of HIV-1 reverse transcriptase initiation complexes and inhibition by NNRTI drugs
.
Nat. Commun.
12
,
2500
103
Larsen
,
K.P.
,
Mathiharan
,
Y.K.
,
Kappel
,
K.
,
Coey
,
A.T.
,
Chen
,
D.H.
,
Barrero
,
D.
et al (
2018
)
Architecture of an HIV-1 reverse transcriptase initiation complex
.
Nature
557
,
118
122
104
Aksel
,
T.
,
Yu
,
Z.
,
Cheng
,
Y.
and
Douglas
,
S.M.
(
2021
)
Molecular goniometers for single-particle cryo-electron microscopy of DNA-binding proteins
.
Nat. Biotechnol.
39
,
378
386
105
Scheres
,
S.H.
(
2012
)
RELION: implementation of a Bayesian approach to cryo-EM structure determination
.
J. Struct. Biol.
180
,
519
530
106
Punjani
,
A.
,
Rubinstein
,
J.L.
,
Fleet
,
D.J.
and
Brubaker
,
M.A.
(
2017
)
cryoSPARC: algorithms for rapid unsupervised cryo-EM structure determination
.
Nat. Methods
14
,
290
296
107
Grant
,
T.
,
Rohou
,
A.
and
Grigorieff
,
N.
(
2018
)
cisTEM, user-friendly software for single-particle image processing
.
eLife
7
,
e35383
108
Nakane
,
T.
and
Scheres
,
S.H.W.
(
2021
)
Multi-body refinement of cryo-EM images in RELION
.
Methods Mol. Biol.
2215
,
145
160
109
Jumper
,
J.
,
Evans
,
R.
,
Pritzel
,
A.
,
Green
,
T.
,
Figurnov
,
M.
,
Ronneberger
,
O.
et al (
2021
)
Highly accurate protein structure prediction with AlphaFold
.
Nature
596
,
583
589
110
Bradley
,
P.
,
Misura
,
K.M.
and
Baker
,
D.
(
2005
)
Toward high-resolution de novo structure prediction for small proteins
.
Science
309
,
1868
1871
111
Kulik
,
M.
,
Mori
,
T.
and
Sugita
,
Y.
(
2021
)
Multi-scale flexible fitting of proteins to cryo-EM density maps at medium resolution
.
Front. Mol. Biosci.
8
,
631854
112
Trabuco
,
L.G.
,
Villa
,
E.
,
Mitra
,
K.
,
Frank
,
J.
and
Schulten
,
K.
(
2008
)
Flexible fitting of atomic structures into electron microscopy maps using molecular dynamics
.
Structure
16
,
673
683
113
Alnabati
,
E.
and
Kihara
,
D.
(
2020
)
Advances in structure modeling methods for cryo-electron microscopy maps
.
Molecules
25
, 82
114
Wang
,
R.Y.-R.
,
Kudryashev
,
M.
,
Li
,
X.
,
Egelman
,
E.H.
,
Basler
,
M.
,
Cheng
,
Y.
et al (
2015
)
De novo protein structure determination from near-atomic-resolution cryo-EM maps
.
Nat. Methods
12
,
335
338
115
Beckers
,
M.
,
Mann
,
D.
and
Sachse
,
C.
(
2021
)
Structural interpretation of cryo-EM image reconstructions
.
Prog. Biophys. Mol. Biol.
160
,
26
36
116
Rout
,
M.P.
and
Sali
,
A.
(
2019
)
Principles for integrative structural biology studies
.
Cell
177
,
1384
1403
117
Shoemaker
,
S.C.
and
Ando
,
N.
(
2018
)
X-rays in the cryo-electron microscopy Era: structural biology's dynamic future
.
Biochemistry
57
,
277
285
118
Mertens
,
H.D.
and
Svergun
,
D.I.
(
2010
)
Structural characterization of proteins and complexes using small-angle X-ray solution scattering
.
J. Struct. Biol.
172
,
128
141
119
Gorbet
,
G.E.
,
Pearson
,
J.Z.
,
Demeler
,
A.K.
,
Cölfen
,
H.
and
Demeler
,
B.
(
2015
)
Next-Generation AUC: analysis of multiwavelength analytical ultracentrifugation data
.
Methods Enzymol.
562
,
27
47
120
Horne
,
C.R.
,
Henrickson
,
A.
,
Demeler
,
B.
and
Dobson
,
R.C.J.
(
2020
)
Multi-wavelength analytical ultracentrifugation as a tool to characterise protein-DNA interactions in solution
.
Eur. Biophys. J.
49
,
819
827
121
Li
,
Y.
,
Struwe
,
W.B.
and
Kukura
,
P.
(
2020
)
Single molecule mass photometry of nucleic acids
.
Nucleic Acids Res.
48
,
e97
122
Ziegler
,
S.J.
,
Mallinson
,
S.J.B.
,
St John
,
P.C.
and
Bomble
,
Y.J.
(
2021
)
Advances in integrative structural biology: towards understanding protein complexes in their cellular context
.
Comput. Struct. Biotechnol. J.
19
,
214
225
123
Dandey
,
V.P.
,
Budell
,
W.C.
,
Wei
,
H.
,
Bobe
,
D.
,
Maruthi
,
K.
,
Kopylov
,
M.
et al (
2020
)
Time-resolved cryo-EM using spotiton
.
Nat. Methods
17
,
897
900
124
Frank
,
J.
(
2017
)
Time-resolved cryo-electron microscopy: recent progress
.
J. Struct. Biol.
200
,
303
306
This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY-NC-ND). Open access for this article was enabled by the participation of University of Melbourne in an all-inclusive Read & Publish pilot with Portland Press and the Biochemical Society under a transformative agreement with CAUL.