Abstract
Transcription is a discontinuous process, where each nucleotide incorporation cycle offers a decision between elongation, pausing, halting, or termination. Many cis-acting regulatory RNAs, such as riboswitches, exert their influence over transcription elongation. Through such mechanisms, certain RNA elements can couple physiological or environmental signals to transcription attenuation, a process where cis-acting regulatory RNAs directly influence formation of transcription termination signals. However, through another regulatory mechanism called processive antitermination (PA), RNA polymerase can bypass termination sites over much greater distances than transcription attenuation. PA mechanisms are widespread in bacteria, although only a few classes have been discovered overall. Also, although traditional, signal-responsive riboswitches have not yet been discovered to promote PA, it is increasingly clear that small RNA elements are still oftentimes required. In some instances, small RNA elements serve as loading sites for cellular factors that promote PA. In other instances, larger, more complicated RNA elements participate in PA in unknown ways, perhaps even acting alone to trigger PA activity. These discoveries suggest that what is now needed is a systematic exploration of PA in bacteria, to determine how broadly these transcription elongation mechanisms are utilized, to reveal the diversity in their molecular mechanisms, and to understand the general logic behind their cellular applications. This review covers the known examples of PA regulatory mechanisms and speculates that they may be broadly important to bacteria.
INTRODUCTION
An extraordinarily diverse range of genetic regulatory mechanisms has been discovered in the half century since Francois Jacob and Jacques Monod first proposed the operon model of gene regulation (1). Studies based on this model identified a soluble regulator, located distally from the targeted operon, that acts to repress transcription initiation of the lac operon. This discovery led to the identification and characterization of many more repressor proteins, each acting in modestly different ways to reduce the efficiency of transcription initiation. Soon followed discoveries of other types of transcriptional regulators, including those that activate gene expression by enhancing transcription initiation. And now, in an era where bacterial genome sequences can be acquired and draft-annotated in mere days and at low cost, it is clear that all bacteria encode for dozens or hundreds of proteins that regulate transcription initiation and that this ‘layer’ of genetic regulation is both ubiquitous and profoundly important. However, perhaps because transcription initiation is so universally recognized as a key point of regulatory influence (2), later stages of transcription elongation have not yet been sufficiently analyzed for genetic regulation. While the molecular mechanisms of transcription have been, and continue to be, intensively investigated, the biological extent of post-initiation regulatory mechanisms has been incompletely analyzed. Transcription initiation is only the first stage of gene expression. The stages that follow include transcription elongation, transcription termination, translation and mRNA degradation; each of these stages can be subjected to genetic regulatory control (3).
While riboswitches, which control transcription attenuation in a signal-dependent manner, are widely used by bacteria, their initial discoveries have been significantly aided by the extensive conservation of their sequences and secondary structures (4, 5). This level of sequence conservation is not observed for many other types of transcription elongation regulatory strategies, a limitation that may have slowed discovery of the latter. How, then, may other transcription elongation-based regulatory strategies be systematically discovered if experimentalists cannot rely primarily on bioinformatics searches of highly conserved regulatory RNAs? And what kinds of transcription elongation regulatory mechanisms have not yet been found? One type of regulatory mechanism that might still be understudied, but yet has been identified through a variety of experimental approaches, is called processive antitermination (PA). These systems offer a convenient and powerful mechanism for altering the efficiency of transcription elongation (6–8).
In PA mechanisms, antitermination factors associate with a bacterial RNA polymerase (RNAP) elongation complex, leading to read-through of termination sites (6). Termination signals normally induce rapid dissociation of the transcription elongation complex (TEC) and are most often located at the ends of operons (9). However, when placed within operons, they can serve as key points of regulatory control (10). In bacteria, there are two known classes of termination signals: intrinsic and Rho-dependent terminators (9). In many bacteria intrinsic terminators consist of a GC-rich RNA hairpin followed by a poly-uridine tract. Alone (11), or enhanced by a factor such as NusA (12, 13), these RNA elements promote pausing of the TEC, followed by release of the nascent transcript and dissociation of polymerase (14). In contrast, Rho-dependent termination depends upon the adenosine triphosphate (ATP)-dependent translocase Rho associating with Rho-utilization (rut) sites on a nascent mRNA and translocating the RNA to eventually promote TEC dissociation (15, 16). Both classes of termination sites may be specifically regulated by signal-responsive riboswitches (5, 17) or trans-encoded small RNAs (18, 19). However, whereas riboswitches exert control over a single intrinsic terminator site, or a particular entry point for Rho, PA systems differ in that they modify TECs to render them generally resistant to downstream termination sites (8). PA systems, therefore, are capable of causing read-through of multiple termination sites, even over long genomic distances. While only a few classes of PA mechanisms have been discovered in the past four decades, they vary widely in the molecular mechanisms they utilize and in their biological applications. Several new examples of PA mechanisms have been discovered more recently, which appear to be broadly used by bacteria for regulation of diverse sets of genes. We extrapolate from these discoveries that many new PA mechanisms still await discovery.
PROCESSIVE ANTITERMINATION
Termination of transcription at any given location is rarely 100% complete, with some proportion of elongation complexes proceeding past the point of termination. In general, two types of mechanisms can control transcription elongation to affect the efficiency of termination: transcription attenuation and processive antitermination (PA). For the former, regulatory mechanisms determine the formation of either Rho-dependent or Rho-independent termination sites (10). Importantly, transcription attenuation-based regulatory mechanisms exert their influence on only a single, defined terminator region. In other words, a regulatory RNA that promotes transcription attenuation by definition evolved in concert with the terminator region that it targets—it does not affect other terminator regions. Riboswitches, which are signal-responsive, cis-acting regulatory RNAs, oftentimes affect gene expression via transcription attenuation-based mechanisms (20). As discussed elsewhere in this book, riboswitches are widespread in bacteria and offer localized control of transcription termination sites throughout bacterial genomes. In many instances, these transcription attenuation-based regulatory elements can be considered modular, with a signal-responsive portion followed by a portion responsible for premature transcription termination (21).
In contrast, PA mechanisms do not necessarily target a specific terminator region, but instead manipulate elongating RNAP complexes to avoid termination signals throughout an individual transcript (6). These PA strategies do not take a single form and may reduce transcript termination through a variety of direct and indirect effects. For example, some PA strategies rely on direct interference with factor-mediated termination (22). Alternatively, they can modify recruitment of transcription elongation factors, such as NusA, to affect nascent RNA behavior (23, 24). Additionally, they may alter recruitment of ribosomes in a manner that affects termination within coding regions (25). Furthermore, some PA systems have evolved to utilize multiple strategies simultaneously (23, 26).
Phage Lambda Antitermination
During lytic growth, phage λ transcription temporally progresses from one large set of genes to another (27). In order to switch from intermediate-early gene expression to delayed-early gene expression, the phage utilizes a unique protein, λN, to promote antitermination, which enables expression of downstream genes (Figure 1A) (28). λN is a small protein that is intrinsically disordered alone (29) but is stabilized by protein and RNA contacts in the final, λN antitermination complex (Figure 2A) (23). Formation of the λN antitermination complex is triggered by synthesis of a nut sequence, composed of two RNA elements. The first, boxB, is a 15-nucleotide motif that resembles a GNRA tetraloop structure (Figure 3) (30, 31) and serves as the substrate for λN binding (23, 24). In addition to binding λN, boxB also interacts with NusA. Formation of the antitermination complex occurs in steps, with initial association of λN to boxB followed by binding of NusA to the λN:boxB complex (32). This minimal λN:boxB:NusA complex is sufficient for antitermination of nut-proximal terminator sequences (6), although it is generally believed that the full antitermination complex in vivo relies on additional elongation proteins loaded at the second RNA element. This second RNA element, boxA, acts as a loading site for the NusB:NusE (S10) complex (33). Binding of the NusB:NusE (S10) complex to boxA promotes additional contacts between λN and NusA. This results in a unique complex of factors that are associated with RNAP near the RNA exit channel and remain together as a ribonucleoprotein complex (Figure 2B) (23).
Binding of λN alone to RNAP modifies transcription elongation both in vitro and in vivo, promoting antitermination by modulating RNA exit channel elements and by suppressing melting of the RNA:DNA hybrid after terminator hairpin formation or in response to Rho activity (23, 24, 34, 35). However, formation of the complex with the full complement of transcription elongation factors is thought to further stabilize the interaction of λN with RNAP and increase its duration of occupancy—and, therefore, overall processivity—of λN antitermination (36). In “standard” transcription elongation complexes, NusA binds RNA polymerase near the RNA exit channel where it can enhance intrinsic termination (37). Indeed, NusA affects transcription termination at many locations across the genome and is even required for formation of some NusA-dependent termination sites (13). However, λN is thought to counteract the direct effects of NusA on terminator hairpin folding (24). A recent high-resolution structural model of the λN antitermination complex revealed that the C-terminal RNA-binding domains of NusA are repositioned such that they redirect nascent RNA away from the RNA exit channel (Figure 2C). This is predicted to reduce formation of terminator hairpins, thereby essentially reprogramming NusA into a transcription antitermination factor (23). Formation of the λN complex also inhibits Rho-dependent termination. In “standard” elongation complexes, NusG helps recruit Rho to nascent RNA and thereby aids in Rho termination (38, 39). In contrast, the λN antitermination complex is likely to restrict NusG-mediated recruitment of Rho by instead promoting association of factors that compete for binding to NusG (e.g., S10:NusB), and also because of restricted access to the nascent RNA as it is looped out of the antitermination complex (23). Therefore, the λN complex acts as a physical roadblock to prevent Rho translocation and helps occlude access to Rho utilization (rut) sites.
Phage λ also contains a second antitermination system, which relies upon another unique protein (λQ) to promote antitermination of late-expressed genes (6, 40). However, unlike the N-antitermination system, λQ protein is a DNA-binding protein that associates with RNA polymerase within the promoter region during transcription initiation and triggers formation of an antitermination complex that is different from the N complex (41).
Ribosomal RNA Operon Antitermination
Dissociation of transcription elongation complexes by Rho helicase underlies the polarity which occurs when nonsense mutations reduce transcript abundance of downstream genes (42). Rho is capable of loading onto RNA molecules via C-rich binding sequences (rut sites), but the presence of ribosomes during coupled transcription-translation generally reduces Rho loading and translocation (43). Given that ribosomal RNA operons are not translated and are thereby not protected by ribosomes, their transcripts must be protected from Rho termination by other means. This protection may be partially explained by the extensive secondary structure of ribosomal RNAs, which acts to reduce loading of Rho at potential rut sites (44, 45). However, in Escherichia coli and many other bacteria, these operons are also subjected to an antitermination system that resembles closely the λN-antitermination mechanism (44–46). For example, the 5’ leader regions of E. coli rRNA operons contain boxA as well as a boxB-like hairpin, although only boxA appears to be essential for antitermination activity (Figure 1C) (33, 47). Binding of the NusB:NusE (S10) complex to boxA RNA occurs in a manner similar to N-mediated antitermination, ultimately promoting a conformational state that strongly disfavors association of Rho (33).
In contrast to λN antitermination, which requires N protein in addition to host Nus proteins, rRNA antitermination requires an additional host factor, SuhB (48). The complete elongation complex containing NusB:NusE, NusA, NusG, and SuhB is required not only for full rRNA antitermination activity in vitro but for correct rRNA maturation in vivo (48). In addition to regulation of rRNA transcription, boxA and Nus factors directly repress suhB translation in enterobacteria in a manner reminiscent of λN autoregulation and have been implicated in regulation of additional genes (49). Therefore, the rRNA antitermination system relies exclusively on general transcription elongation factors and their recruitment to the boxA RNA element. This system serves a dual purpose in rRNA operons, promoting both antitermination and RNA folding, and may regulate yet additional transcripts. Together, these observations suggest that N-antitermination may have arisen as a modification of the host Nus protein antitermination system, where λN protein evolved to reconfigure and further manipulate host transcription elongation factors.
RNA Elements that Promote Processive Antitermination
In addition to the role that RNA elements (boxA and boxB) play in antitermination of phage λ and rRNA operons, a few PA systems have been discovered that involve larger and more complicated RNA elements. Many if not most lambdoid phages utilize PA systems related to both N- or Q-antitermination (6). However, phage HK022 differs in that it encodes for λQ yet lacks λN, despite the fact it still requires antitermination of early-expressed genes (50). Moreover, HK022 does not utilize nut sites for antitermination. Instead, early gene antitermination is mediated directly by a larger RNA motif called put, found in regions analogous to λ nut sites (Figure 1B) (51). HK022 put forms a two-hairpin RNA element of approximately 65 nucleotides in length that is critical for antitermination activity (Figure 3) (51, 52). This element appears to directly affect RNAP elongation activity through pause suppression, potentially requiring no additional elements to promote antitermination (50). Evolution of this mechanism is likely interrelated with the evolution of a λN-like protein, Nun, which is also produced by HK022 (53, 54). Nun, found in the same relative genomic position as λN in phage λ, instead promotes Nun-termination at nut elements by binding to boxB and inhibiting RNAP translocation (55, 56). HK022 put promotes antitermination of both Rho-dependent termination and Nun-dependent transcription arrest in the HK022 early transcripts (55) as well as intrinsic terminators (57). While some mechanistic details of put-mediated antitermination are still lacking, its discovery was significant as it demonstrated proof-in-principle that PA could be driven primarily by RNA elements.
More recently, an even larger and more structurally complicated RNA element was discovered to trigger PA in bacteria. This RNA element, which is at least ~125 nucleotides in length and is constructed from an array of at least five helical elements and a characteristic pseudoknot, was discovered to be broadly conserved in Bacillales (Figure 3) (58). Coined the EAR element, for eps-associated RNA, it is almost always associated with operons that encode for biosynthesis of biofilm or capsule exopolysaccharides (Figure 1D). Either mutagenesis of conserved residues or deletion of EAR resulted in incomplete transcription of the Bacillus subtilis eps operon. Instead, transcripts were found to be prematurely truncated at the site of intrinsic terminators, located in the middle region of the eps operon. Indeed, placement of EAR directly upstream of this terminator site resulted in nearly complete read-through of the terminators in vivo, whereas, conversely, mutagenesis of conserved EAR residues resulted in termination. Moreover, placement of EAR upstream of unrelated intrinsic terminators, originating from sources other than the eps operon, still resulted in their read-through, demonstrating that EAR promotes general PA of intrinsic terminators. That EAR promoted read-through of intrinsic terminators is strikingly different than the biological utilization of the λN and rRNA PA systems, which are believed to function primarily for read-through of Rho termination. However, EAR PA has not yet been recapitulated in vitro or in a heterologous host, indicating that at least one additional factor may be required for its antitermination activity, in contrast to HK022 put. Regardless, discovery of EAR demonstrated that structurally complicated RNAs, with the size and apparent complexity resembling that of riboswitches, are sometimes used to promote PA. Moreover, the distribution of EAR PA determinants further showcases how PA mechanisms can be broadly important for biologically important functions such as biofilm formation.
SPECIALIZED NUSG PARALOGS
RfaH
Although most known PA systems are found in phage genomes or are reliant on general transcription elongation factors, some Gammaproteobacteria encode for the specialized PA and translation factor RfaH (26). RfaH is a paralog of NusG. NusG is an elongation factor generally associated with transcription elongation complexes and is an integral component of the λ and rRNA PA systems (59). RfaH, encoded by an essential gene in E. coli, is required for the expression of a regulon of virulence-related pathways—including synthesis of haemolysin, lipopolysaccharide, and the F-factor sex pilus (59, 60)—as well as additional targets involved in the production of membrane or extracellular components (61).
As a paralog of NusG, RfaH is a small protein containing two conserved domains. In general, the core domains of NusG homolog proteins exhibit strongly conserved structure (62, 63) and interface with RNAP in a similar fashion (63–65). The first domain is an N-terminal domain (NTD) unique to the NusG/Spt5 family of proteins (66). This domain is responsible for binding of RfaH to RNAP at the same site normally occupied by NusG. The C-terminal domain (CTD) contains a KOW (Kyprides, Ouzounis, Woese) motif found in several ribosomal proteins in addition to NusG (67). This characteristic CTD is shared among nearly all NusG homologs as well as several ribosomal proteins (67), and is believed to function as a tether that can interact with additional proteins (68).
While RfaH and NusG have distinct regulatory consequences, they rely on similar mechanisms to improve transcriptional processivity (65). The NTD of both proteins share highly similar sequences and structures (61, 63) and suppress pausing at many sites when added to purified transcription complexes in vitro (22, 69, 70). Both proteins are believed to suppress pausing by binding to the β’ clamp and β pincer and stabilizing the active closed conformation of RNAP (63, 71). Recently, single molecule cryo-EM studies have clarified how stabilization of RNAP structure can promote processive elongation. Certain types of transcriptional pauses are affected by a swivelling of the RNAP β’ pincer elements, resulting in an increase in pause lifetimes (72). However, binding of NusG or RfaH to RNAP disfavors this “swiveled” conformation, thereby suppressing pausing (65). Additional mechanisms for anti-pausing activity of NusG proteins have been proposed, including stabilization of the elongation complex by direct binding to non-template DNA (22, 70, 73) as well as upstream DNA (74–77). Indeed, both NusG and RfaH interact with the upstream DNA fork and promote re-annealing of the upstream DNA, although the specific effects of this activity on RNAP activity are unclear (65, 74). These mechanisms are conserved between NusG and RfaH, are are likely to be shared to varying degrees with other NusG paralogs.
RfaH is specifically recruited to the operons that comprise its regulon by a DNA element called the operon polarity suppressor, or ops (Figure 4A). Deletion of this 8-bp conserved element reduces downstream gene expression (60); correspondingly, introduction of ops to other transcripts increases their expression (59). Depletion of RfaH mirrors these results, indicating that RfaH and ops are both required for expression of target operons (78). RfaH is specifically recruited to transcription elongation complexes by binding to the non-template DNA strand of the ops-element; this occurs during the lifetime of a programmed transcriptional pause (22). The ops-element forms both a consensus pause sequence as well as a DNA hairpin loop that makes specific, direct contacts to the RfaH NTD (79). RfaH and NusG are mutually exclusive, as both homologs share the same binding site on RNAP (80, 81). Moreover, once recruited, RfaH exhibits increased affinity for RNAP relative to NusG, allowing for extended association of RfaH with TECs (82). This increased affinity may also be responsible for the more pronounced effects of RfaH NTD on RNAP as compared to NusG (65). In this way, RfaH exerts its regulatory effects specifically on those operons that include the ops element.
RfaH in solution differs from RNAP-associated RfaH. Instead of the common β-barrel fold found in most high-resolution structures of KOW domains, the CTD of free RfaH forms a dramatically different α-helical structure (80). This α-helical CTD interacts with the NTD, partially masking the RNAP-binding portion and thereby resulting in an autoinhibited form of the protein (Figure 5) (82). After a conformational change is triggered, the NTD can associate fully with the transcription complex, which in turn promotes re-folding of the CTD to the β-barrel structure found in NusG (Figure 5) (26). Because of this structural mechanism, RfaH adopts the classical NusG KOW domain structure only after the NTD has fully associated with RNA polymerase.
Though NusG and RfaH display nearly identical anti-pausing effects on transcription complexes in vitro, their overall regulatory outcomes are different. In some instances, NusG may promote pausing in vivo (83), perhaps as a result of increased affinity for certain non-template DNA strand sequences (70). More importantly, NusG is known to directly bind Rho (68). This interaction is likely to broadly promote Rho-dependent termination activity, possibly by increasing the rate at which Rho successfully binds RNA and forms a closed translocation-capable conformation (39). Ultimately, association of NusG results in Rho-dependent termination and suppression of transcription, particularly in genomic regions that feature foreign DNA (25). This activity is essential in most E. coli strains primarily due to suppression of toxic genes in prophage DNA (25). However, in addition to its interaction with Rho, the NusG CTD can associate with NusE (S10), as well as NusA (84, 85). Similar to NusG, RfaH can associate with NusE (S10); however, in contrast to NusG, RfaH is incapable of binding Rho (26, 84). Because of this, RfaH strongly discourages Rho termination within its targeted operons (86).
Finally, the remaining mechanism by which RfaH may promote antitermination is through recruitment of ribosomes to nascent transcripts. NusG proteins are thought to couple transcription and translation by facilitating macromolecular interactions between both of these machines (84). RfaH in particular has been shown to exhibit much stronger polarity effects in vivo than its effects on transcription in vitro (82). Also, genes that are known to be regulated by RfaH display particularly poor ribosome binding sites, suggesting that translational enhancement is likely to be a key feature of RfaH regulation (86). It is possible that binding of NusG or RfaH to ribosomal S10 (NusE) may assist ribosome recruitment, thereby increasing local concentration of ribosomes and promoting translation initiation on nascent RNA (26, 87). This functional interaction might also affect transcription processivity. Indeed, recent data suggest that the leading ribosome—which conducts translation immediately upstream of RNAP, and that may participate in the RNAP-ribosome “expressome” (88, 89)—improves transcription processivity by directly blocking RNAP backtracking (90) and by obstructing Rho access (84, 91).
Through these aggregate mechanisms, RfaH acts as a specialized elongation factor that exhibits anti-pausing activity, prevents NusG-mediated Rho termination, and encourages ribosome recruitment, for each of the operons that display ops elements.
Other NusG Paralogs (ActX, TaA, UpxY, LoaP)
Although RfaH is the most prominent and best studied NusG paralog, other examples have been identified, several of which have been predicted to function in transcription antitermination (92–95). All of these homologs share significant sequence similarity to NusG and RfaH and undoubtedly share conserved structural features. Moreover, for those NusG paralogs where a functional role has been demonstrated, they have inevitably been found to affect transcription of certain targeted transcripts, suggesting that NusG paralogs are broadly used by bacteria as specialized transcription regulators (93–95).
ActX and TraB proteins are most phylogenetically similar to RfaH (92, 96) and are found in a variety of conjugative plasmids conferring antibiotic resistance in Gammaproteobacteria (92, 97). Though a function has not been demonstrated for these proteins, they are often transcribed as the first open reading frame in long pilus biosynthesis operons and are suspected to be involved in the transcription of conjugation genes (98).
Myxococcus xanthus, a Gram-negative soil bacterium, produces the well-studied polyketide antibiotic TA (also called myxovirescin) (99, 100). The first open reading frame of the TA-producing gene cluster is taA, which encodes for a NusG paralog (93). Disruption of the taA gene eliminated antibiotic production, suggesting a regulatory relationship. However, the specific role of TaA in expression of the TA gene cluster is unknown, although as a NusG paralog and relative of other known NusG specialized paralogs, it has been proposed to regulate transcription elongation, perhaps through PA.
More recently, a NusG paralog called UpxY has been proposed to function as a family of regulators for complex polysaccharide pathways in Bacteroidetes (94, 101). They are widely used by these microorganisms. Indeed, many Bacteroides encode between six and nine copies of the UpxY proteins. The genes encoding these proteins, initially described in Bacteroides fragilis, are each associated with a different capsular polysaccharide gene cluster (Figure 4C) (101). These proteins have been shown to affect transcription of their associated gene cluster and it has been proposed that they participate in antitermination-based regulatory mechanisms that involve unique sequence features located within the 5’ leader regions of their respective operons. Additionally, while these regulators might be co-transcribed with the operons they affect, they can also affect gene expression when moved to a distal location, supporting the claim that they are recruited to their targeted operons, perhaps via sequence elements within the transcript leader regions. Yet, despite these observations, little is known regarding the molecular mechanisms of UpxY proteins. Adding a new wrinkle to the overall family of NusG paralogous proteins, Bacteroides fragilis also encodes a set of unique proteins (UpxZ) alongside genes encoding UpxY proteins. The UpxZ proteins can act as trans-inhibitors of UpxY proteins, and have been hypothesized to hierarchically regulate the expression of different sets of capsular polysaccharides, although the underlying mechanism of this inhibition is also unknown (94).
The most recently described NusG paralog is LoaP. Genes encoding LoaP are consistently positioned adjacent to polyketide biosynthesis pathways in Firmicutes, or near polysaccharide biosynthesis gene clusters in certain Firmicutes, Actinobacteria and Spirochaetes (95). While proximity alone is not evidence of a functional relationship, Bacillus velezensis loaP was shown to affect expression of an adjacent polyketide synthesis gene cluster. More specifically, Bacillus velezensis loaP is situated adjacent to a gene cluster (dfn) that encodes for production of the polyketide antibiotic difficidin (102). Deletion of loaP resulted in low abundance across the dfn transcript, whereas complementation of loaP from an ectopic locus restored dfn expression. Moreover, these global expression experiments revealed that the difficidin operon is not the only region affected by LoaP. LoaP is also required for transcription elongation of a second polyketide gene cluster, which encodes for production of another antibiotic, macrolactin, indicating that LoaP controls a regulon of antibiotic biosynthesis operons in B. velezensis (Figure 4B).
Upon depletion of LoaP, transcript abundance dramatically decreases at intrinsic terminator sites located within the targeted operons, whereas induction of LoaP restored read-through of these terminators (95). These initial data suggest that LoaP antitermination may function primarily on intrinsic termination, in contrast to the suppression of Rho-dependent termination known for RfaH (22, 95). However, while this might be due to the preference of Gram-positive bacteria for Rho-independent antitermination, it is also true that Rho termination has been insufficiently characterized in B. velezensis and other Firmicutes (103). Therefore, the relationship between LoaP proteins and Rho termination is still unknown. Nor have the full determinants for LoaP PA yet been described. While LoaP affects transcript abundance across the length of the targeted operon, it appears to require sequence elements located somewhere within the 5’ leader region. Indeed, when placed upstream of terminator signals and upstream of a reporter gene, the dfn leader region alone is sufficient for promoting LoaP-dependent PA activity (95). Therefore, the recruitment signals for LoaP are located fully within this leader region. Interestingly, a small UNCG-type hairpin was identified in the leader regions of both the difficidin and macrolactin operons in a sequence region required for antitermination, although its exact relationship to LoaP regulation has not yet been investigated (Figure 3) (95).
The discovery of LoaP, along with the initial description of TaA, suggests that transcription elongation may be a broad point of regulatory control for secondary metabolite gene clusters in bacteria. Therefore, it is important to study PA mechanisms in order to improve discovery and production of new natural products from bacteria.
Phylogenetic Overview of the NusG Family of Proteins
NusG paralogs putatively involved in antitermination have been identified in a variety of bacteria, including but not limited to Alpha-, Gamma- and Delta-proteobacteria, Bacteroidetes, and most recently Firmicutes, Actinobacteria, and Spirochaetes (86, 95). Of the general transcription elongation factors, only NusG is found in all three domains of life, suggesting its function is important in all organisms. Therefore, essentially all bacteria encode for a core NusG protein, while archaea and eukaryotes encode for a similar protein, Spt5 (80). As a result, all NusG family proteins share core conserved sequence and structure features (80).
Although analysis of the paralogs supports grouping them within the overall NusG family, each sub-grouping displays significant sequence diversity, with some subgroups displaying very limited overall sequence identity despite sharing remarkably conserved structural elements (95). In recent work, the phylogenetic analyses of the NusG family was extended to include as many distant homologs as are detectable by HMM-based homology modeling (95). This large-scale phylogenetic analysis utilized structural modeling to efficiently align specialized NusG paralogs with limited sequence similarity, and focused on comprehensively covering the diversity of paralog sequences without restriction to the known subgroups. The resulting phylogenetic tree (Figure 6) confirmed that each set of NusG paralogs forms its own distinct group, separate from core bacterial NusG and archaeal Spt5, while also revealing a few new candidate subgroups (82, 95). It is likely that each subgroup will be defined by specific sequence differences. Indeed, a number of characteristic differences between sequences—such as between RfaH or UpxY and core NusG—have been identified as being important for the distinct activities of those specialized paralogs (82, 101, 104).
As NusG paralogs were found in a variety of distinct genetic contexts (82), it was important to systematically identify associations between these genes and potential target pathways. Overall, they were found in diverse genomic contexts, with some positioned alone, at the beginning of complex polysaccharide or secondary metabolite gene clusters, at the end of operons, or in unique contexts (82, 95). For example, NusG paralogous sequences from Betaproteobacteria and Bacteroides are located in or near large polysaccharide pathways. TaA and LoaP sequences are generally present in or near large polyketide biosynthesis pathways, which suggests they share a broad relationship to secondary metabolites (82, 95). Indeed, there appears to be a general association of NusG specialized paralogs with polysaccharide biosynthesis gene clusters, and to a lesser extent polyketide synthase gene clusters. In fact, of all the paralog groups, only the Gammaproteobacterial RfaH and its related ActX gene sequences were not frequently identified near or in these classes of gene clusters (95).
There also appear to be several subgroups of NusG paralogs with interesting genomic association and evolutionary distribution, but that have not been characterized or named. For example, a group of sequences closely related to RfaH and found in Alpha-, Beta-, and Gamma-proteobacteria is oftentimes associated with polysaccharide gene clusters. Similarly, at least two more uncharacterized and unnamed putative groups of sequences are consistently associated with polysaccharide and polyketide biosynthesis gene clusters. From this, it can be tentatively speculated that NusG specialized paralogs evolved as regulators of these long operons (polysaccharides and secondary metabolite biosynthesis genes) and became further specialized into RfaH in Gammaproteobacteria. Finally, an additional set of paralog sequences in Alphaproteobacteria was not found in a consistent genomic context, and remains unnamed. Ultimately, the evolutionary relationship between all these different NusG paralogs remains unclear, as bootstrap support for early branches after divergence from core NusG is low, likely due to the extensive sequence divergence in this family. Elucidating the true history of this family may require different approaches, integrating more information about the structural changes and sequence insertions and deletions during evolution of the NusG paralogs. However, it is already clear that the NusG family of proteins is widely used in bacteria as specialized transcription elongation regulatory factors, and that they regulate expression of fundamentally important pathways, albeit through largely unexplored molecular mechanisms.
OUTLOOK
The past few years have uncovered a few new examples of PA mechanisms, as well as remarkable new insight into the structural basis of antitermination activity. However, it is possible that these findings still only represent a small proportion of what remains to be discovered. Therefore, what is now needed is a systematic exploration of the molecular mechanisms used by NusG paralogs, combined with new bioinformatics searches for RNA elements that promote PA. From this, an accurate portrayal of the extent of PA usage in bacteria will emerge, which will resolve whether transcription elongation is a much broader point of regulatory control than has historically been perceived. Furthermore, studying new PA systems will help uncover the diversity of their molecular mechanisms and shed important light on when and why PA mechanisms are employed by bacteria.
References
- 1.Jacob F, Monod J. 1961. Genetic regulatory mechanisms in the synthesis of proteins. J Mol Biol 3:318–356. [DOI] [PubMed] [Google Scholar]
- 2.Walsh C 2003. Regulation of Antibiotic Biosynthesis in Producer Organisms, p. 159–174. In Antibiotics: Actions, Origins, Resistance. ASM Press. [Google Scholar]
- 3.Waters LS, Storz G. 2009. Regulatory RNAs in bacteria. Cell 136:615–628. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Barrick JE, Breaker RR. 2007. The distributions, mechanisms, and structures of metabolite-binding riboswitches. Genome Biol 8:R239. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Breaker RR. 2011. Prospects for riboswitch discovery and analysis. Mol Cell 43:867–879. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Weisberg RA, Gottesman ME. 1999–1. Processive Antitermination. J Bacteriol 181:359–367. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Roberts JW, Shankar S, Filter JJ. 2008. RNA polymerase elongation factors. Annu Rev Microbiol 62:211–233. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Santangelo TJ, Artsimovitch I. 2011–5. Termination and antitermination: RNA polymerase runs a stop sign. Nat Rev Microbiol 9:319–329. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Ray-Soni A, Bellecourt MJ, Landick R. 2016. Mechanisms of Bacterial Transcription Termination: All Good Things Must End. Annu Rev Biochem 85:319–347. [DOI] [PubMed] [Google Scholar]
- 10.Merino E, Yanofsky C. 2005. Transcription attenuation: a highly conserved regulatory strategy used by bacteria. Trends Genet 21:260–264. [DOI] [PubMed] [Google Scholar]
- 11.Gusarov I, Nudler E. 1999. The mechanism of intrinsic transcription termination. Mol Cell 3:495–504. [DOI] [PubMed] [Google Scholar]
- 12.Greenblatt J, McLimont M, Hanly S. 1981. Termination of transcription by nusA gene protein of Escherichia coli. Nature 292:215–220. [DOI] [PubMed] [Google Scholar]
- 13.Mondal S, Yakhnin AV, Sebastian A, Albert I, Babitzke P. 2016. NusA-dependent transcription termination prevents misregulation of global gene expression. Nature Microbiology 1:15007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Epshtein V, Cardinale CJ, Ruckenstein AE, Borukhov S, Nudler E. 2007. An allosteric path to transcription termination. Mol Cell 28:991–1001. [DOI] [PubMed] [Google Scholar]
- 15.Holmes WM, Platt T, Rosenberg M. 1983. Termination of transcription in E. coli. Cell 32:1029–1032. [DOI] [PubMed] [Google Scholar]
- 16.Epshtein V, Dutta D, Wade J, Nudler E. 2010. An allosteric mechanism of Rho-dependent transcription termination. Nature 463:245–249. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Proshkin S, Mironov A, Nudler E. 2014. Riboswitches in regulation of Rho-dependent transcription termination. Biochim Biophys Acta 1839:974–977. [DOI] [PubMed] [Google Scholar]
- 18.DebRoy S, Gebbie M, Ramesh A, Goodson JR, Cruz MR, van Hoof A, Winkler WC, Garsin D a. 2014. A riboswitch-containing sRNA controls gene expression by sequestration of a response regulator. Science 345:937–940. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Sedlyarova N, Shamovsky I, Bharati BK, Epshtein V, Chen J, Gottesman S, Schroeder R, Nudler E. 2016. sRNA-Mediated Control of Transcription Termination in E. coli. Cell 167:111–121.e13. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Montange RK, Batey RT. 2008. Riboswitches: emerging themes in RNA structure and function. Annu Rev Biophys 37:117–33. [DOI] [PubMed] [Google Scholar]
- 21.Ceres P, Garst AD, Marcano-Velázquez JG, Batey RT. 2013. Modularity of select riboswitch expression platforms enables facile engineering of novel genetic regulatory devices. ACS Synth Biol 2:463–472. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Artsimovitch I, Landick R. 2002. The transcriptional regulator RfaH stimulates RNA chain synthesis after recruitment to elongation complexes by the exposed nontemplate DNA strand. Cell 109:193–203. [DOI] [PubMed] [Google Scholar]
- 23.Said N, Krupp F, Anedchenko E, Santos KF, Dybkov O, Huang Y-H, Lee C-T, Loll B, Behrmann E, Bürger J, Mielke T, Loerke J, Urlaub H, Spahn CMT, Weber G, Wahl MC. 2017. Structural basis for λN-dependent processive transcription antitermination. Nat Microbiol 2:17062. [DOI] [PubMed] [Google Scholar]
- 24.Gusarov I, Nudler E. 2001. Control of intrinsic transcription termination by N and NusA: the basic mechanisms. Cell 107:437–449. [DOI] [PubMed] [Google Scholar]
- 25.Cardinale CJ, Washburn RS, Tadigotla VR, Brown LM, Gottesman ME, Nudler E. 2008. Termination factor Rho and its cofactors NusA and NusG silence foreign DNA in E. coli. Science 320:935–938. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Burmann BM, Knauer SH, Sevostyanova A, Schweimer K, Mooney R a., Landick R, Artsimovitch I, Rösch P. 2012. An α helix to β barrel domain switch transforms the transcription factor RfaH into a translation factor. Cell 150:291–303. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Friedman DI. 1988. Regulation of phage gene expression by termination and antitermination of transcription. The bacteriophages 2:263–319. [Google Scholar]
- 28.Patterson TA, Zhang Z, Baker T, Johnson LL, Friedman DI, Court DL. 1994. Bacteriophage lambda N-dependent transcription antitermination. Competition for an RNA site may regulate antitermination. J Mol Biol 236:217–228. [DOI] [PubMed] [Google Scholar]
- 29.Mogridge J, Legault P, Li J, Van Oene MD, Kay LE, Greenblatt J. 1998. Independent ligand-induced folding of the RNA-binding domain and two functionally distinct antitermination regions in the phage lambda N protein. Mol Cell 1:265–275. [DOI] [PubMed] [Google Scholar]
- 30.Legault P, Li J, Mogridge J, Kay LE, Greenblatt J. 1998. NMR structure of the bacteriophage lambda N peptide/boxB RNA complex: recognition of a GNRA fold by an arginine-rich motif. Cell 93:289–299. [DOI] [PubMed] [Google Scholar]
- 31.Thapar R, Denmon AP, Nikonowicz EP. 2014. Recognition modes of RNA tetraloops and tetraloop-like motifs by RNA-binding proteins. Wiley Interdiscip Rev RNA 5:49–67. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Mogridge J, Mah T-F, Greenblatt J. 1998. Involvement of boxA Nucleotides in the Formation of a Stable Ribonucleoprotein Complex Containing the Bacteriophage λ N Protein. J Biol Chem 273:4143–4148. [DOI] [PubMed] [Google Scholar]
- 33.Nodwell JR, Greenblatt J. 1993. Recognition of boxA antiterminator RNA by the E. coli antitermination factors NusB and ribosomal protein S10. Cell 72:261–268. [DOI] [PubMed] [Google Scholar]
- 34.Mason SW, Li J, Greenblatt J. 1992. Host factor requirements for processive antitermination of transcription and suppression of pausing by the N protein of bacteriophage lambda. J Biol Chem 267:19418–19426. [PubMed] [Google Scholar]
- 35.Rees WA, Weitzel SE, Das A, von Hippel PH. 1997. Regulation of the elongation-termination decision at intrinsic terminators by antitermination protein N of phage lambda. J Mol Biol 273:797–813. [DOI] [PubMed] [Google Scholar]
- 36.Nudler E, Gottesman ME. 2002. Transcription termination and anti-termination in E. coli. Genes Cells 7:755–768. [DOI] [PubMed] [Google Scholar]
- 37.Liu K, Zhang Y, Severinov K, Das A, Hanna MM. 1996. Role of Escherichia coli RNA polymerase alpha subunit in modulation of pausing, termination and anti-termination by the transcription elongation factor NusA. EMBO J 15:150–161. [PMC free article] [PubMed] [Google Scholar]
- 38.Peters JM, Mooney RA, Grass JA, Jessen ED, Tran F, Landick R. 2012. Rho and NusG suppress pervasive antisense transcription in Escherichia coli. Genes Dev 26:2621–2633. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Valabhoju V, Agrawal S, Sen R. 2016. Molecular Basis of NusG-mediated Regulation of Rho-dependent Transcription Termination in Bacteria. J Biol Chem 291:22386–22403. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Herskowitz I, Signer ER. 1970. A site essential for expression of all late genes in bacteriophage lambda. J Mol Biol 47:545–556. [DOI] [PubMed] [Google Scholar]
- 41.Yarnell WS, Roberts JW. 1992. The phage lambda gene Q transcription antiterminator binds DNA in the late gene promoter as it modifies RNA polymerase. Cell 69:1181–1189. [DOI] [PubMed] [Google Scholar]
- 42.Lowery C, Richardson JP. 1977. Characterization of the nucleoside triphosphate phosphohydrolase (ATPase) activity of RNA synthesis termination factor p. II. Influence of synthetic RNA …. J Biol Chem. [PubMed] [Google Scholar]
- 43.Guérin M, Robichon N, Geiselmann J, Rahmouni AR. 1998. A simple polypyrimidine repeat acts as an artificial Rho-dependent terminator in vivo and in vitro. Nucleic Acids Res 26:4895–4900. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Li SC, Squires CL, Squires C. 1984. Antitermination of E. coli rRNA transcription is caused by a control region segment containing lambda nut-like sequences. Cell 38:851–860. [DOI] [PubMed] [Google Scholar]
- 45.Squires CL, Greenblatt J, Li J, Condon C, Squires CL. 1993. Ribosomal RNA antitermination in vitro: requirement for Nus factors and one or more unidentified cellular components. Proc Natl Acad Sci U S A 90:970–974. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Arnvig KB, Zeng S, Quan S, Papageorge A, Zhang N, Villapakkam AC, Squires CL. 2008. Evolutionary comparison of ribosomal operon antitermination function. J Bacteriol 190:7251–7257. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Berg KL, Squires C, Squires CL. 1989. Ribosomal RNA operon anti-termination. Function of leader and spacer region box B-box A sequences and their conservation in diverse micro-organisms. J Mol Biol 209:345–358. [DOI] [PubMed] [Google Scholar]
- 48.Singh N, Bubunenko M, Smith C, Abbott DM, Stringer AM, Shi R, Court DL, Wade JT. 2016. SuhB Associates with Nus Factors To Facilitate 30S Ribosome Biogenesis in Escherichia coli. MBio 7:e00114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Baniulyte G, Singh N, Benoit C, Johnson R, Ferguson R, Paramo M, Stringer AM, Scott A, Lapierre P, Wade JT. 2017. Identification of regulatory targets for the bacterial Nus factor complex. Nat Commun 8:2027. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Clerget M, Jin DJ, Weisberg RA. 1995. A zinc-binding region in the beta’ subunit of RNA polymerase is involved in antitermination of early transcription of phage HK022. J Mol Biol 248:768–780. [DOI] [PubMed] [Google Scholar]
- 51.King RA, Banik-Maiti S, Jin DJ, Weisberg RA. 1996. Transcripts that increase the processivity and elongation rate of RNA polymerase. Cell 87:893–903. [DOI] [PubMed] [Google Scholar]
- 52.Banik-Maiti S, King RA, Weisberg RA. 1997. The antiterminator RNA of phage HK022. J Mol Biol 272:677–687. [DOI] [PubMed] [Google Scholar]
- 53.Robert J, Sloan SB, Weisberg RA, Gottesman ME, Robledo R, Harbrecht D. 1987. The remarkable specificity of a new transcription termination factor suggests that the mechanisms of termination and antitermination are similar. Cell 51:483–492. [DOI] [PubMed] [Google Scholar]
- 54.Hung SC, Gottesman ME. 1997. The Nun protein of bacteriophage HK022 inhibits translocation of Escherichia coli RNA polymerase without abolishing its catalytic activities. Genes Dev 11:2670–2678. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.King RA, Weisberg RA. 2003. Suppression of factor-dependent transcription termination by antiterminator RNA. J Bacteriol 185:7085–7091. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Vitiello CL, Kireeva ML, Lubkowska L, Kashlev M, Gottesman M. 2014. Coliphage HK022 Nun protein inhibits RNA polymerase translocation. Proc Natl Acad Sci U S A 111:E2368–75. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Oberto J, Clerget M, Ditto M, Cam K, Weisberg RA. 1993. Antitermination of early transcription in phage HK022. Absence of a phage-encoded antitermination factor. J Mol Biol 229:368–381. [DOI] [PubMed] [Google Scholar]
- 58.Irnov I, Winkler WC. 2010. A regulatory RNA required for antitermination of biofilm and capsular polysaccharide operons in Bacillales. Mol Microbiol 76:559–575. [DOI] [PubMed] [Google Scholar]
- 59.Bailey MJ, Hughes C, Koronakis V. 1996. Increased distal gene transcription by the elongation factor RfaH, a specialized homologue of NusG. Mol Microbiol 22:729–737. [DOI] [PubMed] [Google Scholar]
- 60.Bailey MJ, Koronakis V, Schmoll T, Hughes C. 1992. Escherichia coli HlyT protein, a transcriptional activator of haemolysin synthesis and secretion, is encoded by the rfaH (sfrB) locus required for expression of sex factor and lipopolysaccharide genes. Mol Microbiol 6:1003–1012. [DOI] [PubMed] [Google Scholar]
- 61.Bailey MJ, Hughes C, Koronakis V. 1997. RfaH and the ops element, components of a novel system controlling bacterial transcription elongation. Mol Microbiol 26:845–851. [DOI] [PubMed] [Google Scholar]
- 62.Reay P, Yamasaki K, Terada T, Kuramitsu S, Shirouzu M, Yokoyama S. 2004. Structural and sequence comparisons arising from the solution structure of the transcription elongation factor NusG from Thermus thermophilus. Proteins 56:40–51. [DOI] [PubMed] [Google Scholar]
- 63.Martinez-Rucobo FW, Sainsbury S, Cheung ACM, Cramer P. 2011. Architecture of the RNA polymerase-Spt4/5 complex and basis of universal transcription processivity. EMBO J 30:1302–1310. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Liu B, Steitz TA. 2017. Structural insights into NusG regulating transcription elongation. Nucleic Acids Res 45:968–974. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Kang JY, Mooney RA, Nedialkov Y, Saba J, Mishanina TV, Artsimovitch I, Landick R, Darst SA. 2018. Structural Basis for Transcript Elongation Control by NusG Family Universal Regulators. Cell 173:1650–1662.e14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Ponting CP. 2002. Novel domains and orthologues of eukaryotic transcription elongation factors. Nucleic Acids Res 30:3643–3652. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Kyrpides NC, Woese CR, Ouzounis CA. 1996. KOW: a novel motif linking a bacterial transcription factor with ribosomal proteins. Trends Biochem Sci 21:425–426. [DOI] [PubMed] [Google Scholar]
- 68.Mooney RA, Schweimer K, Roesch P, Gottesman M, Landick R. 2009. Two structurally independent domains of E. coli NusG create regulatory plasticity via distinct interactions with RNA polymerase and regulators. J Mol Biol 391:341–358. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Burova E, Hung SC, Sagitov V, Stitt BL, Gottesman ME. 1995. Escherichia coli NusG protein stimulates transcription elongation rates in vivo and in vitro. J Bacteriol 177:1388–1392. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Yakhnin AV, Murakami KS, Babitzke P. 2016. NusG is a Sequence-specific RNA Polymerase Pause Factor that Binds to the Non-template DNA Within the Paused Transcription Bubble. J Biol Chem. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Weixlbaumer A, Leon K, Landick R, Darst SA. 2013. Structural basis of transcriptional pausing in bacteria. Cell 152:431–441. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Kang JY, Mishanina TV, Bellecourt MJ, Mooney RA, Darst SA, Landick R. 2018. RNA Polymerase Accommodates a Pause RNA Hairpin by Global Conformational Rearrangements that Prolong Pausing. Mol Cell 69:802–815.e1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Crickard JB, Fu J, Reese JC. 2016. Biochemical Analysis of Yeast Suppressor of Ty 4/5 (Spt4/5) Reveals the Importance of Nucleic Acid Interactions in the Prevention of RNA Polymerase II Arrest. J Biol Chem 291:9853–9870. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Nedialkov Y, Svetlov D, Belogurov GA, Artsimovitch I. 2018. Locking the non-template DNA to control transcription. Mol Microbiol. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Guo G, Gao Y, Zhu Z, Zhao D, Liu Z, Zhou H, Niu L, Teng M. 2015. Structural and biochemical insights into the DNA-binding mode of MjSpt4p:Spt5 complex at the exit tunnel of RNAPII. J Struct Biol 192:418–425. [DOI] [PubMed] [Google Scholar]
- 76.Ehara H, Yokoyama T, Shigematsu H, Yokoyama S, Shirouzu M, Sekine S-I. 2017. Structure of the complete elongation complex of RNA polymerase II with basal factors. Science 357:921–924. [DOI] [PubMed] [Google Scholar]
- 77.Turtola M, Belogurov GA. 2016. NusG inhibits RNA polymerase backtracking by stabilizing the minimal transcription bubble. Elife 5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Beutin L, Manning PA, Achtman M, Willetts N. 1981. sfrA and sfrB products of Escherichia coli K-12 are transcriptional control factors. J Bacteriol 145:840–844. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Zuber PK, Artsimovitch I, NandyMazumdar M, Liu Z, Nedialkov Y, Schweimer K, Rösch P, Knauer SH. 2018. The universally-conserved transcription factor RfaH is recruited to a hairpin structure of the non-template DNA strand. Elife 7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Belogurov GA, Vassylyeva MN, Svetlov V, Klyuyev S, Grishin NV, Vassylyev DG, Artsimovitch I. 2007. Structural basis for converting a general transcription factor into an operon-specific virulence regulator. Mol Cell 26:117–129. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Yakhnin AV, Yakhnin H, Babitzke P. 2008. Function of the Bacillus subtilis transcription elongation factor NusG in hairpin-dependent RNA polymerase pausing in the trp leader. Proc Natl Acad Sci U S A 105:16131–16136. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Belogurov G a., Mooney R a., Svetlov V, Landick R, Artsimovitch I. 2009. Functional specialization of transcription elongation factors. EMBO J 28:112–22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Yakhnin AV, Babitzke P. 2014. NusG/Spt5: are there common functions of this ubiquitous transcription elongation factor? Curr Opin Microbiol 18:68–71. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Burmann BM, Schweimer K, Luo X, Wahl MC, Stitt BL, Gottesman ME, Rösch P. 2010. A NusE:NusG complex links transcription and translation. Science 328:501–504. [DOI] [PubMed] [Google Scholar]
- 85.Strauß M, Vitiello C, Schweimer K, Gottesman M, Rösch P, Knauer SH. 2016. Transcription is regulated by NusA:NusG interaction. Nucleic Acids Res 44:5971–5982. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Tomar SK, Artsimovitch I. 2013. NusG-Spt5 proteins-Universal tools for transcription modification and communication. Chem Rev 113:8604–8619. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Saxena S, Myka KK, Washburn R, Costantino N, Court DL, Gottesman ME. 2018. Escherichia coli transcription factor NusG binds to 70S ribosomes. Mol Microbiol 108:495–504. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Kohler R, Mooney RA, Mills DJ, Landick R, Cramer P. 2017. Architecture of a transcribing-translating expressome. Science 356:194–197. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Demo G, Rasouly A, Vasilyev N, Svetlov V, Loveland AB, Diaz-Avalos R, Grigorieff N, Nudler E, Korostelev AA. 2017. Structure of RNA polymerase bound to ribosomal 30S subunit. Elife 6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Proshkin S, Rahmouni AR, Mironov A, Nudler E. 2010. Cooperation between translating ribosomes and RNA polymerase in transcription elongation. Science 328:504–508. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Banerjee S, Chalissery J, Bandey I, Sen R. 2006. Rho-dependent transcription termination: more questions than answers. J Microbiol 44:11–22. [PMC free article] [PubMed] [Google Scholar]
- 92.Núñez B, Avila P, de la Cruz F. 1997. Genes involved in conjugative DNA processing of plasmid R6K. Mol Microbiol 24:1157–1168. [DOI] [PubMed] [Google Scholar]
- 93.Paitan Y, Orr E, Ron EZ, Rosenberg E. 1999. A NusG-like transcription anti-terminator is involved in the biosynthesis of the polyketide antibiotic TA of Myxococcus xanthus. FEMS Microbiol Lett 170:221–7. [DOI] [PubMed] [Google Scholar]
- 94.Chatzidaki-Livanis M, Weinacht KG, Comstock LE. 2010. Trans locus inhibitors limit concomitant polysaccharide synthesis in the human gut symbiont Bacteroides fragilis. Proc Natl Acad Sci U S A 107:11976–11980. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Goodson JR, Klupt S, Zhang C, Straight P, Winkler WC. 2017. LoaP is a broadly conserved antiterminator protein that regulates antibiotic gene clusters in Bacillus amyloliquefaciens. Nat Microbiol 2:17003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Arutyunov D, Arenson B, Manchak J, Frost LS. 2010. F plasmid TraF and TraH are components of an outer membrane complex involved in conjugation. J Bacteriol 192:1730–1734. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Jones CS, Osborne DJ, Stanley J. 1993. Molecular comparison of the IncX plasmids allows division into IncX1 and IncX2 subgroups. J Gen Microbiol 139:735–741. [DOI] [PubMed] [Google Scholar]
- 98.NandyMazumdar M, Artsimovitch I. 2015. Ubiquitous transcription factors display structural plasticity and diverse functions: NusG proteins - Shifting shapes and paradigms. Bioessays 37:324–334. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Varon M, Fuchs N, Monosov M, Tolchinsky S, Rosenberg E. 1992. Mutation and mapping of genes involved in production of the antibiotic TA in Myxococcus xanthus. Antimicrob Agents Chemother 36:2316–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Simunovic V, Zapp J, Rachid S, Krug D, Meiser P, Müller R. 2006. Myxovirescin A biosynthesis is directed by hybrid polyketide synthases/nonribosomal peptide synthetase, 3-hydroxy-3-methylglutaryl-CoA synthases, and trans-acting acyltransferases. Chembiochem 7:1206–20. [DOI] [PubMed] [Google Scholar]
- 101.Chatzidaki-Livanis M, Coyne MJ, Comstock LE. 2009. A family of transcriptional antitermination factors necessary for synthesis of the capsular polysaccharides of Bacteroides fragilis. J Bacteriol 191:7288–7295. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Chen X-H, Vater J, Piel J, Franke P, Scholz R, Schneider K, Koumoutsi A, Hitzeroth G, Grammel N, Strittmatter AW, Gottschalk G, Süssmuth RD, Borriss R. 2006. Structural and functional characterization of three polyketide synthase gene clusters in Bacillus amyloliquefaciens FZB 42. J Bacteriol 188:4024–36. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Mitra P, Ghosh G, Hafeezunnisa M, Sen R. 2017. Rho Protein: Roles and Mechanisms. Annu Rev Microbiol 71:687–709. [DOI] [PubMed] [Google Scholar]
- 104.Shi D, Svetlov D, Abagyan R, Artsimovitch I. 2017. Flipping states: a few key residues decide the winning conformation of the only universally conserved transcription factor. Nucleic Acids Res 45:8835–8843. [DOI] [PMC free article] [PubMed] [Google Scholar]