ABSTRACT
Bacteria exhibit a rich repertoire of RNA molecules that intricately regulate gene expression at multiple hierarchical levels, including small RNAs (sRNAs), riboswitches, and antisense RNAs. Notably, the majority of these regulatory RNAs lack or have limited protein-coding capacity but play pivotal roles in orchestrating gene expression by modulating transcription, post-transcription or translation processes. Leveraging and redesigning these regulatory RNA elements have emerged as pivotal strategies in the domains of metabolic engineering and synthetic biology. While previous investigations predominantly focused on delineating the roles of regulatory RNA in Gram-negative bacterial models such as Escherichia coli and Salmonella enterica, this review aims to summarize the mechanisms and functionalities of endogenous regulatory RNAs inherent to typical Gram-positive bacteria, notably Bacillus subtilis. Furthermore, we explore the engineering and practical applications of these regulatory RNA elements in the arena of synthetic biology, employing B. subtilis as a foundational chassis.
ARTICLE INFO
Keywords:
Regulatory RNA devices Small RNAs Riboswitches Synthetic biology Bacillus subtilis
1. Introduction
Regulatory RNAs are recognized as ubiquitous and functionally diversified post-transcriptional regulator of gene expression in both prokaryotes and eukaryotes [1]. They participate in many cellular physiological processes, such as biofilm formation, ion homeostasis, metabolism regulation, anti-toxification, pathogenesis [1,2]. Regulatory RNAs found in prokaryotes are categorized as riboswitches, small non-coding RNAs (sRNAs), antisense sRNAs etc [2]. The regulatory RNAs vary in length and function through distinct mechanisms.
Riboswitches are typical regulatory RNAs in 5' or 3' untranslated region (UTR) of mRNA (messenger RNA). It could bind to specific small molecules (ligands) and regulate gene expression through changes in secondary structure of the mRNAs and thus the binding of the ribosomes [3,4]. sRNAs are typically trans-encoded regulatory RNAs with an average length of 50-300 nt [5] and interact with multiple target mRNAs by imperfect base pairing, causing mRNA degradation or translation blocking. Antisense RNAs engage in extensive base-pairing interactions with the complementary mRNA, as they are transcribed from the DNA strand opposite to that encoding the mRNA and can span from ten to thousands of nucleotides in length [6].
In addition to regulating their natural targets, regulatory RNAs such as riboswitches and sRNAs have been structurally redesigned and developed as efficient and independent regulatory tools in prokaryotes to regulate non-natural target gene expression via the canonical or noncanonical regulation mechanism [7-16]. CRISPR RNAs (clustered regularly interspaced short palindromic repeat RNAs), essential components of the bacterial innate immune system against bacteriophage, guide Cas proteins (CRISPR-associated proteins) to targeted DNA or RNA sequences. They were endowed with regulatory function in the CRISPRi (CRISPR interference) scenario [17-20]. The interference of CRISPRi is mediated by both deactivated Cas protein and non-coding RNA, controlling the mRNA generation of the target genes [21]. Considering only a short transcript is required, the sRNA regulation systems could be quickly constructed. Comparing to CRISPRi, some engineered or de novo designed sRNA have less polar effect on regulating polycistronic mRNA expressions [8,22]. Besides, the expression of the deactivated Cas protein may impose a larger metabolic burden than sRNA.
Till now most research on regulatory RNA primarily focused on Gram-negative bacteria [23], such as Escherichia coli and Salmonella enterica. Compared to their counterparts in Gram-negative bacteria, regulatory RNAs in B. subtilis, which is an important chassis in synthetic biology, are less explored. Here we focus on riboswitches and sRNAs in B. subtilis, summarizing endogenous RNA regulatory mechanism, analyzing design methods of artificial RNA devices and predicting future development of artificial RNA tools for application in the field of synthetic biology and metabolic engineering. Although CRISPR RNAs/gRNAs are involved in the regulation of gene expression when applied to CRISPR interference, they do not fall into the category of classic regulatory RNAs and thus are not within the scope of discussion for this review.
2. Riboswitch in B. subtilis
2.1. Regulation mechanism of riboswitches
Riboswitches are predominantly located in 5' UTR of mRNA and act in cis, consisting of an aptamer domain to sense and bind target ligands and an expression platform that modulates the activation or repression of downstream genes [24]. Generally, riboswitches primarily regulate transcription in Gram-positive bacteria but translation in Gram-negative microorganisms [25,26]. Some riboswitches in Gram-negative bacteria lack intrinsic terminators; their termination is assisted by Rho [25]. Transcriptionally, riboswitches could mediate transcription anti-termination or termination after ligand binding. Transcription termination is categorized into Rho-dependent termination and Rho-independent termination. Rho-dependent termination requires binding of Rho (recognizes C-rich residue and unstructured RNA) to rut site (Rho utilization site) on mRNA [26]. Rho moves along mRNA until it meets and interacts with RNA polymerase (RNAP), leading to the dissociation of the transcription elongation complex and subsequent termination of transcription. In certain situations, riboswitch-mediated translational control and transcriptional control are coupled (Fig. 1A). When ligands bind to riboswitches, rut site would be exposed and RBS (ribosome binding site) is simultaneously sequestered. This dual action results in Rho binding and ribosome detachment [27]. Rho-independent termination requires an intrinsic transcription terminator consisting of a strong hairpin structure followed by poly-uridine residues [28]. Riboswitch-mediated transcription termination leads to formation of terminator and release of RNAP from DNA template and RNA transcript (Fig. IB). Conversely, in Rho-independent transcription antitermination, ligand-bound riboswitch sequesters the terminator, allowing RNAP to continue elongating through the DNA-RNA complex (Fig. 1C) [29]. Translationally, RBS could be obscured or exposed by secondary structure change of riboswitches in 5UTR, dynamically controlling ribosome binding or detachment (Fig. ID and E). For instance, the ribozyme-riboswitch glmS in B. subtilis, located in the 5'UTR, activates its self-cleavage activity upon binding to GlcN6P, leading to the degradation of mRNA [30] (Fig. IF). In B. subtilis, all identified riboswitches regulate through Rho-independent transcriptional termination or anti-termination, translation inhibition and self-cleavage. Riboswitches could also exert control via Rho-dependent transcription termination or translation initiation in other prokaryotes [31-33].
2.2. Ligands of B. subtilis riboswitches
According to types of ligands, riboswitches are classified into metabolite riboswitches, tRNA riboswitches and protein riboswitches. Details of endogenous riboswitches from B. subtilis are shown in Table 1. Metabolites are the main source of riboswitch ligands. Elucidated metabolite riboswitch ligands in B. subtilis include Flavin mononucleotide (FMN) [34], S-adenosylmethionine (SAM) [35-37], adenine [38-40], ATP [41], c-di-AMP [42-46], lysine [47] and glucosamine-6-phosphate (G6P) [30,48]. T-box or tRNA riboswitches that selectively bind to a cognate tRNA regulate gene expression through transcription antitermination by binding uncharged tRNA (to which its cognate amino acid is not chemically bonded). In B. subtilis, glyQS [49-51] and tyrS [52,53] are regulated by tRNA. For some riboswitches their ligands are the proteins encoded by the downstream coding regions. For instance, the expression of some ribosomal protein genes are autoregulated through protein riboswitches [54], including L10(L12)4 [55], L13-S9 [56], L19 [56], L20 [57,58], S4 [58], S10 [59], and S15 [58]. These regulatory elements primarily serve to limit the accumulation of excessive unbound ribosomal proteins by suppressing the transcription or translation of downstream genes [55,58]. Tryptophan synthesis is also regulated by riboswitches responding to trp mRNA binding attenuation protein (TRAP, synthesis of TRAP is controlled by tryptophan) [60]. Utilization of alternative sugars is also regulated by riboswitches [61], the ligands of which belong to BglG family, including GlcT [62], SacT [63], SacY [64], and LicT [65]. There are other three riboswitches, PyrR [66], GlpP [67], and HutP [68] responsible for biosynthesis and uptake of nucleotides, glycerol-3-phosphate, and histidine utilization respectively.
3. Engineering of riboswitches in B. subtilis and their applications
3.1. Monitoring DNA mutation
Under harsh environmental pressure, B. subtilis sporulate to aid their survival. In biological production, sporulated cells have the potential to serve as time-delayed chassis for expression at specific time. Recently, sporulated cells have also been applied as vessels for DNA storage [69]. However, after DNA replication during late growth phase before sporulation or after continuous subculture, DNA mutation would happen (Fig. 2A). Denis Tamiev et al. created a DNA mutation monitor on plasmid based on theophylline riboswitch [70]. By monitoring fluorescence of riboswitch-controlled RFP (red fluorescence protein), DNA point mutation could be indirectly monitored (Fig. 2A). The length of riboswitch is always much shorter compared to coding gene, thereby exerting little metabolic pressure. This approach is significant for monitoring mutations of critical genes in industrial production, since mutation is a major issue in industrial culture.
3.2. Transforming constitutive gene expression into inducible gene expression
Inducible systems, capable of turning genes on and off, are essential for biochemical expression in microbial production. But in B. subtilis, inducible systems are widely used, while the high price of inducers like IPTG or xylose limit their application in industry [71,72]. In that sense, alternative induction systems with low basal level, high induction rate and low-cost inducer need to be constructed. By inserting riboswitches in 5UTR, constitutive systems could be reformed into inducible systems. Phan et al. discovered gcv operon controlled by glycine riboswitch in B. subtilis can be converted into an inducible expression-secretion system [72]. Furthermore, riboswitches could also strengthen inducible systems. For instance, inserting lysine or theophylline riboswitches downstream of an inducible promoter can increase the induction fold change up to a hundredfold [73].
3.3. Orthogonal regulation
Orthogonal regulation in synthetic biology refers to independent regulation of genes without interference with each other or the biological environment. This approach enables the design of more complex metabolic networks. There are two main strategies for achieving orthogonal regulation based on riboswitch.
The first strategy is designing analogues of riboswitch ligands. Regulation by analogues of natural ligands could mitigate influences on other metabolic reaction in vivo. Artificial analogues, synthesized through chemical reactions [74], could be designed to bind more tightly to aptamers [75,76]. While synthesizing new ligands is challenging, altering the nucleotide sequence of aptamers is comparably less complicated. Additionally, the inherent modularity of riboswitches lays the foundation for riboswitch engineering, allowing the exchange of aptamer domains and expression platforms between different riboswitches [77].
The second strategy is reforming riboswitches to recognizing new ligands, encompassing three kinds of approaches: 1) introducing mutation on aptamers to generate derivatives of riboswitches [78,79]; 2) utilizing exogenous riboswitches from other strains [79,80]; 3) designing chimera riboswitches hybridized by more than two different riboswitches [77,81,82], which could also alter regulation mechanism (Fig. 2B). These approaches facilitate interactions between different riboswitches and ligands. Such strategies are instrumental in creating complex gene circuits, enabling multiple simultaneous regulations or multistep cascading regulations.
4. sRNA in B. subtilis
4.1. Regulation mechanism of sRNAs
Most of the sRNAs discovered act post-transcriptionally or translationally by base-pairing with target mRNA. With the burgeoning discovery of the sRNA mechanisms in prokaryotes, sRNAs are found to work almost in all levels of gene expression.
Transcriptionally, sRNAs could mediate gene expression through transcription read-through (transcription could not stop normally at terminators). In bacteria, there are two kinds of transcription readthrough. The first type of transcription read-through involves abnormal transcription termination of sRNA itself. This phenomenon, observed in both SR6 and SR7, is speculated as the result of inefficient transcription terminator [83,84]. The correlation between environmental stress and sRNA read-through has not been fully explained. After sRNA read-through, the range of target genes may be expanded. Another form of sRNA-mediated transcription read-through involves competition between sRNA and Rho of Rho utilization (rut) site (Fig. ЗА). DsrA, ArcZ, and RprA in E. coli [85] and SraL in S. enterica [86] have been demonstrated to compete with Rho for a specific mRNA rut site. Rho-dependent termination was most discovered in enteric Gram-negative bacteria and relatively less known in Gram-positive bacteria [87]. But recent research have increasingly shown importance of Rho-dependent termination in Gram-positive bacteria [88]. However, competition between sRNA and Rho of rut site has not been clarified in B. subtilis yet.
Post-transcription regulation mediated by sRNAs is widespread in bacteria. Most of the post-transcription regulation are mediated by ribonuclease. However, due to the different ribonuclease repertoire among species, the mechanism of post-transcription differs [89-91]. In B. subtilis, sRNAs from type I TA systems (toxin-antitoxin systems) commonly regulate through post-transcriptional degradation. The degradation is mainly mediated by RNase III and assisted by RNase Y and RNase JI (Fig. 3B). Until now, only four type I TA systems have been fully investigated in B. subtilis; txp/RatA [92], bsrG/SR4 [93,94], bsrE/SR5 [95,96], and yonT-yoyJ/SR.6 [84] (Fig. 4). Those sRNAs from TA systems act as small antitoxin molecules and base-pair with toxin mRNA. With RNase III recognizing base pair region, the toxin-antitoxin complex would be cleaved by RNase III and further digested by RNase Y and RNase JI [97] (Fig. 3B). Exceptionally, SR6 regulates yoyJ through translation inhibition [84]. Post-transcription protection is another way of regulation opposite to degradation. In B. subtilis, for example, when sRNA RoxS binds to 5' end of yflS mRNA, it prevents RNase JI from degrading the mRNA (Fig. 3B) [98].
Translationally, sRNAs could also mediate translational inhibition by binding RBS on mRNA or by binding ribosome proteins (Fig. 3C). Basepairing of sRNA and RBS is widely seen in bacteria, yoyJ/SR6 in B. subtilis [84], for example. Even though sRNAs are not found in B. subtilis to bind ribosomes, SprFl, a ribosome-binding sRNA, is recently found in Staphylococcus aureus to block binding from ribosome to RBS, thus inhibiting translation [99] (Fig. 3C). Translational activation usually acts through opening up secondary structure near RBS. In B. subtilis, RosX binds to 5'UTR of yflS mRNA, protecting it from RNase JI and stimulating 30S (ribosomal subunit) binding to RBS [98].
Post-translationally, sRNAs in type III TA system act as antitoxins by binding to toxin proteins and sequestering them by forming protein-RNA complexes [100] (Fig. 3D).
4.2. Classification of sRNAs in B. subtilis
The types and mechanisms of plenty endogenous sRNAs in B. subtilis are thoroughly characterized [101]. Here we provide a new perspective based on the role of sRNAs in metabolism and physiological process (Fig. 5). Some characterized sRNAs in B. subtilis directly participate in physiological processes like transportation and sporulation. Others are actively involved in metabolic pathways regulations including arginine metabolism, iron metabolism and control of NAD+/NADH balance. Additionally, some sRNAs are part of the immune system of B. subtilis, TA system, for example.
FsrA plays an important part in the tricarboxylic acid cycle (TCA cycle), down-regulating aconitase (citB) and succinate dehydrogenase (sdhCAB) post-transcriptionally [102]. FsrA also mediates repression of glutamate synthase, which serves as a vital link between central carbon metabolism and nitrogen metabolism. FsrA also represses dicarboxylate transporter (DctP), important for increasing TCA cycle intermediates. RoxS [98] is another important sRNA in TCA cycle. RoxS activates yflS by binding to the 5' end of the yflS mRNA with the C-rich region CRR3, protecting yflS (encoding a malate transporter) from RNase JI and stimulating 80S binding to the RBS. RoxS also down-regulates several genes post-transcriptionally, including ppnkB (encoding NAD-kinase) and TCA components sucC (encoding succinate dehydrogenase) and citZ (encoding citrate synthase). Another sRNA, corroborated by electrophoretic mobility shift assays (EMSA) to interact with RoxS and FsrA, is called RosA [103,104]. The length of RosA varies in different species (225, 193, 128, or 92 nt), demonstrated to be the result of activity of endo- and exo-ribonucleases [103].
Arginine metabolism is regulated by at least two transcriptional regulators RocR [105,106] and AhrC. SRI down regulates ahrC post-transcriptionally by targeting ahrC mRNA, which encodes a transcriptional activator of rocABC and rocDEF operon in arginine metabolism. Toeprinting studies and secondary structure probing of the ahrC/SRl complex indicated that SRI inhibits translation initiation by inducing structural change downstream from the RBS of ahrC. The interaction between SRI and ahrC mRNA, facilitated by Hfq, has more than 7 base-pairing regions as predicted by computational analysis.
The iron-sparing response is regulated by sRNA FsrA in conjunction with three small basic proteins, FbpABC. FsrA, working alongside FbpABC, represses many "low-priority" iron-containing enzymes. The lactate-inducible lutABC operon encodes iron sulfur-containing enzymes required for growth on lactate. FsrA, together with FbpB, also represses the synthesis of the LutABC lactate oxidase enzymes [107].
In the process of sporulation, it has been demonstrated that SRI targets kinA (a sporulation-specific ATP-dependent histidine kinase) mRNA. The deletion of srl accelerates sporulation but results in lower spore quality [108]. SRI inhibits the translation of kinA mRNA in vivo, but does not affect its stability. Research by Mars RA et al. has predicted that sRNAs S25, S31, S37, S526, S547, S623, S661, S1009, S1083, SI279, SI388, SI445 and SI559 could also be related to sporulation [54]. According to PhD thesis of Holly Hall, promoters of S357, S547, S612 and S849 are active during the early stages of the sporulation process [109]. These findings suggest that sporulation in B. subtilis could be significantly influenced by the activity of many sRNAs. The overexpression or knock-out of sporulation related sRNAs could also influence sporulation efficiency [109]. This understanding could be crucial for scaling-up production in industrial settings, potentially reducing the negative impact on the dormancy of spores.
In B. subtilis, sRNAs play a role in type I toxin-antitoxin (TA) systems as part of its immune mechanisms. The txpA/RatA system consists of the antitoxin sRNA RatA and toxin TxpA (59 aa). TxpA could lead to cell lysis in the absence of RatA [92]. The bsrG/SR4 system includes the antitoxin sRNA SR4 and toxin BsrG (38 aa). BsrG causes cell wall defects, membrane invaginations, and altered cells shape in the absence of SR4 [93,94]. The bsrE/SR5 system includes the antitoxin sRNA SR5 and toxin BsrE (30 aa), where BsrE is less toxic than the other type I TA system toxins [95,96]. The yonT-yoyJ/SR6 system encodes antitoxin sRNA SR6 and two toxins YonT and YoyJ. YonT causes cell lysis, while YoyJ is weaker than YonT but is still detrimental in the absence of SR6 [84].
From the perspective of endogenous sRNAs' function in the physiological process, sRNAs play important roles in transportation, metabolism, sporulation, and the immune system to bacteriophage. According to function of base-pairing genes, predicted sRNAs in B. subtilis may also be involved in the replication of genes, cell wall biogenesis, ribosome synthesis, tRNA synthesis, two-component systems and so on [54]. Given the diverse regulatory mechanisms of sRNAs and the comprehensive understanding of their regulatory pathways. sRNAs are being engineered and utilized as efficient toolboxes in synthetic biology, enabling the fine-tuning of specific gene expression.
5. Artificial sRNA design and application
5.1. Transcription regulation with riboswitch-targeting sRNAs
Synthetic sRNAs could also play the role as transcription activator. Small transcription activating RNAs (STARs) have been previously designed to target transcription attenuators and riboswitches in E. coli [9,110]. Those sRNAs could pair with premature terminator and activate gene expression. This work achieved artificial sRNA-mediated transcription activation in vivo first time [9]. Lins et al. firstly implemented STARs in B. subtilis, called riboswitch-targeting sRNAs (rtRNA) [111]. By targeting at terminator stem-loop, rtRNAs activate gene expression by turning riboswitches into ON state (Fig. 6A). rtRNAs could both work in vitro and in vivo, increasing gene expression up to 103-fold [111]. This work effectively engineered a natural RNA transcriptional repressor as well as the ability to convert intrinsic terminators into transcription-on regulators. This also achieved sRNA-based metabolic regulation and RNA-only genetic networks in vivo. The simplicity of rtRNAs suggests that sRNA based transcription activation may be a natural mechanism of gene regulation waiting to be discovered in B. subtilis [9].
5.2. Post-transcription regulation with redesigned bsrG/SR4
According to the way of inhibition mechanism of base-pair, TA systems could be constructed into gene regulation tools without manipulating any protein. Post-transcriptionally, we have modified the TA system bsrG/SR4 in B. subtilis to a useful genetic tool, named as modulation via the small RNA-dependent operation system (MS-DOS) [112]. Operation region, a part of the toxin bsrG coding region, is required to insert after the stop codon of target genes. Base-pairing between operation region and SR4 triggers RNase III degradation of a complex of the target gene and SR4 (Fig. 6B),achieving post-transcription inhibition over targeted genes. MS-DOS was verified by inhibiting ftsZ in B. subtilis, the cell of which was lengthened greatly because of abnormal cell division [112]. MS-DOS was also applied to regulate crucial genes in hyaluronan biosynthesis. Down-regulation of pfkA resulted in the highest hyaluronan titer (1.52 g/L) which was 1.6-fold of the parental strain [112]. Inhibition by MS-DOS could be more stable than sRNA regulation translationally, because MS-DOS introduce RNase III cleavage site. So orthogonal regulation of multiple genes could be easier using MS-DOS.
5.3. Translation regulation with redesigned yonT-yoyJ/SR6
Endogenous sRNAs could be modified and designed as useful metabolic regulation tool, requiring no genome-editing process. But not all endogenous sRNAs have the potential to be designed as robust tools to regulate target genes without destroying core scaffold [113]. There were a lot of examples of artificial sRNA systems in E. coli, finding out more native scaffold [114,115] or designing novel scaffold [116,117]. The sRNA system MicC-Hfq from E. coli was also successfully transplanted into C. glutamicum [118], but such translation regulation sRNA system was not established in B. subtilis. We have modified TA system yoyJ/SR6 in B. subtilis into regulatory tool acting translationally by base-pairing 24 nucleotides with mRNA starting from N-terminal coding sequence AUG (Fig. 6C). With minimized structure, SR6 was proven to maintain a strong repression activity of 83%. This artificial sRNA system was also applied in E. coli, demonstrated to have repression efficiency above 80%, which could function without Hfq, causing lower metabolic burden [119]. sRNAs with arbitrary sequences and fixed secondary structures were also designed by a de novo sRNA design program to match any gene of interest, which was demonstrated to be pretty efficient in down-regulating expression of comER and ftsZ and functioned well in acetoin biosynthesis regulation [119]. Arbitrary sequence gets rid of traditional fixed sRNA scaffold, making gene regulation by sRNA more customized. To regulate different genes under different situations, sRNAs with different inhibition efficiency could be chosen. Because of its convenience to construct, this method could also be applied to screen genes of interest through high-throughput screening [119].
For sRNA knock-down tools in diverse bacteria, Cho et al. designed broad-host-range sRNA system (BHR-sRNA system) base-on sRNA scaffold of RoxS from B. subtilis [120]. Translationally, BHR-sRNA system achieved knockdown of reporter genes in 12 strains out of 16 strains, with slight modification in each strain. This demonstrated that sRNA system with same mechanism could be applied in multiple species, which is pretty meaningful in trans-bacteria regulation.
6. Conclusions and outlook
6.1. Create riboswitches for detection of small molecules
Due to its versality and designability, riboswitches have potential to bind various small molecules. To develop aptamers recognizing small molecules, SELEX (Systematic Evolution of Ligands by Exponential enrichment) is a prominent method, not only selecting for binding but also structure changes on binding in aptamers. SELEX could be divided into in vitro SELEX [121,122] and in vivo SELEX [123-125]. However, after being selected under in vitro conditions, aptamers may lose activity under in vivo conditions [124]. Another problem is immobilization of both aptamers and targets, inevitably changing function of target compound [123,126].
With the progress of bioinformatics, in silico design of riboswitches is developing with a high speed and applicable in various kinds of strains. Designing RNA aptamers recognizing versatile molecules have immense practical importance. Riboswitches could be applied as small molecule detectors monitoring metabolism (Fig. 7A), since de novo riboswitch responsive to specific ligand could be designed according to thermodynamic and kinetic analysis [76,127].
6.2. Rapid identification of metabolic targets with sRNA library
As gene regulation tools, sRNAs and their libraries could be constructed more easily, offering higher efficiency and non-polar regulatory advantages when compared to CRISPRi or CRISPRa [8]. For example, we have constructed single-stem loop small non-coding RNAs (ssl-sRNA) library with predictable and programmable activities and applied to screen out gene candidate in complex metabolic pathway in E. coli [8]. Knocking-down of some genes in a specific metabolic pathway, related transporter genes or related regulators could increase production [8]. This method could be improved for rapid identification of metabolic targets. After constructing a sRNA library with one-pot PCR using multiple primers and selecting out transformants with high production, next-generation DNA sequencing would be carried out to find out which sRNA plays the role and then the down-regulated gene would also be found out [8] (Fig. 7B). Rapid identification could quickly select out critical genes and hugely shorten the time of metabolic engineering. This method could be also applied in B. subtilis and other strains.
6.3. Designing of toehold switch based on regulatory RNA
Toehold switches are de novo RNA engineering elements, consisted of two strand, a triggering RNA strand and a toehold-hairpin strand with regulated gene (Fig. 7C) [128]. After base-paring with trigger RNA, toehold-hairpin would be opened and RBS would be exposed, enabling ribosome binding. Compared with other regulatory RNA devices, toehold switches are highly modular, orthogonal and programmable [128,129]. Toehold switches could not only regulate gene expression [13,128], but also detect mRNAs, such as virus RNAs [130]. To date, toehold switches have primarily been developed and applied in E. coli. There is a likelihood that they could also be adapted for use in B. subtilis synthetic biology and even in diagnostics [131]. Furthermore, enhancing existing regulatory RNA devices in B. subtilis might involve incorporating novel features inspired by the engineering of other regulatory RNAs.
Declaration of competing interest
There are no conflicts of interest to declare.
Acknowledgements
This work was supported by the National Natural Science Foundation of China (31970085 and 32000058) and the National Key Research and Development Program of China (2021YFC2100800).
https://doi.org/10.1016/j.synbio.2024.01.013
Received 10 December 2023; Received in revised form 15 January 2024; Accepted 31 January 2024
Available online 10 February 2024
References
[1] Morris KV, Mattick JS. The rise of regulatory RNA. Nat Rev Genet 2014;15(6): 423-37.
[2] Waters LS, Storz G. Regulatory RNAs in bacteria. Cell 2009;136(4):615-28.
[3] Breaker RR. Riboswitches and the RNA world. Cold Spring Harbor Perspect Biol 2012;4(2):a003566.
[4] Serganov A, Nudler E. A decade of riboswitches. Cell 2013;152(1-2):17-24.
[5] Storz G, Vogel J, Wassarman KM. Regulation by small RNAs in bacteria: expanding frontiers. Mol Cell 2011;43(6):880-91.
[6] Thomason MK, Storz G. Bacterial antisense RNAs: how many are there, and what are they doing? Annu Rev Genet 2010;44:167-88.
[7] Na D, Yoo SM, Chung H, Park H, Park JH, Lee SY. Metabolic engineering of Escherichia coli using synthetic small regulatory RNAs. Nat Biotechnol 2013;31 (2):170-4.
[8] Wang Y, Yin G, Weng H, Zhang L, Du G, Chen J, Kang Z. Gene knockdown by structure defined single-stem loop small non-coding RNAs with programmable regulatory activities. Synth Syst Biotechnol 2023;8(1):86-96.
[9] Chappell J, Takahashi MK, Lucks JB. Creating small transcription activating RNAs. Nat Chem Biol 2015;11(3):214-20.
[10] Yarra SS, Ashok G, Mohan U. Toehold switches; a foothold for synthetic biology. Biotechnol Bioeng 2023;120(4):932-52.
[11] Choi S, Lee G, Kim J. Cellular computational logic using toehold switches. Int J Mol Sci 2022;23(8):4265.
[12] Hong S, Kim J, Kim J. Multilevel gene regulation using switchable transcription terminator and toehold switch in Escherichia coli. Appl Sci 2021;11(10):4532.
[13] Falgenhauer E, Muckl A, Schwarz-Schilling M, Simmel FC. Transcriptional interference in toehold switch-based RNA circuits. ACS Synth Biol 2022;11(5): 1735-45.
[14] Hong KQ, Zhang J, Jin B, Chen T, Wang ZW. Development and characterization of a glycine biosensor system for fine-tuned metabolic regulation in Escherichia coli. Microb Cell Factories 2022;21(1):56.
[15] Vikram, Mishra V, Rana A, Ahire JJ. Riboswitch-mediated regulation of riboflavin biosynthesis genes in prokaryotes. 3 Biotech 2022;12(10):278.
[16] Irla M, Hakvag S, Brautaset T. Developing a riboswitch-mediated regulatory system for metabolic flux control in thermophilic Bacillus methanolicus. Int J Mol Sci 2021;22(9):4686.
[17] Wu Y, Liu Y, Lv X, Li J, Du G, Liu L, Camers-B. CRISPR/Cpf1 assisted multiplegenes editing and regulation system for Bacillus subtilis. Biotechnol Bioeng 2020; 117(6):1817-25.
[18] Wu Y, Chen T, Liu Y, Tian R, Lv X, Li J, Du G, Chen J, Ledesma-Amaro R, Liu L. Design of a programmable biosensor-CRISPRi genetic circuits for dynamic and autonomous dual-control of metabolic flux in Bacillus subtilis. Nucleic Acids Res 2020;48(2):996-1009.
[19] Wu Y, Chen T, Liu Y, Lv X, Li J, Du G, Ledesma-Amaro R, Liu L. CRISPRi allows optimal temporal control of N-acetylglucosamine bioproduction by a dynamic coordination of glucose and xylose metabolism in Bacillus subtilis. Metab Eng 2018;49:232-41.
[20] Yu W, Jin K, Wu Y, Zhang Q, Liu Y, Li J, Du G, Chen J, Lv X, Ledesma-Amaro R, Liu L. A pathway independent multi-modular ordered control system based on thermosensors and CRISPRi improves bioproduction in Bacillus subtilis. Nucleic Acids Res 2022;50(11):6587-600.
[21] Mitarai N BJ, Krishna S, Semsey S, Csiszovszki Z, Massé E, Sneppen K. Dynamic features of gene expression control by small regulatory RNAs. Proc Natl Acad Sci USA 2009;106(26):10655-9.
[22] Rock JM, Hopkins FF, Chavez A, Diallo M, Chase MR, Gerrick ER, Pritchard JR, Church GM, Rubin EJ, Sassetti CM, Schnappinger D, Fortune SM. Programmable transcriptional repression in mycobacteria using an orthogonal CRISPR interference platform. Nature Microbiology 2017;2(4).
[23] Irnov I, Sharma CM, Vogel J, Winkler WC. Identification of regulatory RNAs in Bacillus subtilis. Nucleic Acids Res 2010;38(19):6637-51.
[24] Husser C, Dentz N, Ryckelynck M. Structure-switching RNAs: from gene expression regulation to small molecule detection. Small Structures 2021;2(4): 2000132.
[25] Ariza-Mateos A, Nuthanakanti A, Serganov A. Riboswitch mechanisms: new tricks for an old dog. Biochemistry (Moscow) 2021;86(8):962-75.
[26] Hollands K, Proshkin S, Sklyarova S, Epshtein V, Mironov A, Nudler E, Groisman EA. Riboswitch control of Rho-dependent transcription termination. Proc Natl Acad Sci U S A 2012;109(14):5376-81.
[27] Bastet L, Chauvier A, Singh N, Lussier A, Lamontagne A-M, Prévost K, Massé E, Wade JT, Lafontaine DA. Translational control and Rho-dependent transcription termination are intimately linked in riboswitch regulation. Nucleic Acids Res 2017;45(12):7474-86.
[28] Yarnell Ws RJ. Mechanism of intrinsic transcription termination and antitermination. Science 1999;284(5414):611-5.
[29] Grundy Fj HT. tRNA as a positive regulator of transcription antitermination in B. subtilis. Cell 1993;74:475-82.
[30] Lau MW, Ferre-D'Amare AR. In vitro evolution of coenzyme-independent variants from the glmS ribozyme structural scaffold. Methods 2016;106:76-81.
[31] Howe JA, Xiao L, Fischmann TO, Wang H, Tang H, Villafania A, Zhang R, Barbieri CM, Roemer T. Atomic resolution mechanistic studies of ribocil: a highly selective unnatural ligand mimic of the E. coli FMN riboswitch. RNA Biol 2016;13 (10):946-54.
[32] Wang H, Mann PA, Xiao L, Gill C, Galgoci AM, Howe JA, Villafania A, Barbieri CM, Malinverni JC, Sher X, Mayhood T, McCurry MD, Murgolo N, Flattery A, Mack M, Roemer T. Dual-targeting small-molecule inhibitors of the Staphylococcus aureus FMN riboswitch disrupt riboflavin homeostasis in an infectious setting. Cell Chem Biol 2017;24(5):576-88.
[33] Caron MP, Bastet L, Lussier A, Simoneau-Roy M, Masse E, Lafontaine DA. Dualacting riboswitch control of translation initiation and mRNA decay. Proc Natl Acad Sci U S A 2012;109(50):E3444-53.
[34] Winkler WC. An mRNA structure that controls gene expression by binding FMN. Proc Natl Acad Sci USA 2022;99(25):15908-13.
[35] Manz C, Kobitski AY, Samanta A, Nienhaus K, Jäschke A, Nienhaus GU. Exploring the energy landscape of a SAM-I riboswitch. J Biol Phys 2021;47(4):371-86.
[36] Price IR, Grigg JC, Ke A. Common themes and differences in SAM recognition among SAM riboswitches. Biochim Biophys Acta 2014;1839(10):931-8.
[37] Grundy FJ, Henkin TM. The S box regulon: a new global transcription termination control system for methionine and cysteine biosynthesis genes in Gram-positive bacteria. Mol Microbiol 1998;30(4):737-49.
[38] Zakataeva NP, Gronskiy SV, Sheremet AS, Kutukova EA, Novikova AE, Livshits VA. A new function for the Bacillus PbuE purine base efflux pump: efflux of purine nucleosides. Res Microbiol 2007;158(8-9):659-65.
[39] Gong S, Wang Y, Zhang W. Kinetic regulation mechanism of pbuE riboswitch. J Chem Phys 2015;142(1):015103.
[40] Seif E, Altman S. RNase P cleaves the adenine riboswitch and stabilizes pbuE mRNA in Bacillus subtilis. RNA 2008;14(6):1237-43.
[41] Watson PY, Fedor MJ. The ydaO motif is an ATP-sensing riboswitch in Bacillus subtilis. Nat Chem Biol 2012;8(12):963-5.
[42] Nelson JW, Sudarsan N, Furukawa K, Weinberg Z, Wang JX, Breaker RR. Riboswitches in eubacteria sense the second messenger c-di-AMP. Nat Chem Biol 2013;9(12):834-9.
[43] Gundlach J, Kruger L, Herzberg C, Turdiev A, Poehlein A, Tascon I, Weiss M, Hertel D, Daniel R, Hanelt I, Lee VT, Stulke J. Sustained sensing in potassium homeostasis: cyclic di-AMP controls potassium uptake by KimA at the levels of expression and activity. J Biol Chem 2019;294(24):9605-14.
[44] Gundlach J. Control of potassium homeostasis is an essential function of the second messenger cyclic di-AMP in Bacillus subtilis. Sci Signal 2017;10.
[45] Ren A, Patel DJ. c-di-AMP binds the ydaO riboswitch in two pseudo-symmetryrelated pockets. Nat Chem Biol 2014;10(9):780-6.
[46] Gao A, Serganov A. Structural insights into recognition of c-di-AMP by the ydaO riboswitch. Nat Chem Biol 2014;10(9):787-92.
[47] Sun Y, Wang Y, Tan Z-J, Zhang W. Regulation mechanism of lysC riboswitch in gram-positive bacterium Bacillus subtilis. J Biomol Struct Dyn 2019;38(9): 2784-91.
[48] Collins JA, Irnov I, Baker S, Winkler WC. Mechanism of mRNA destabilization by the glmS ribozyme. Genes Dev 2007;21(24):3356-68.
[49] Grundy FJ, Henkin TM. Kinetic analysis of tRNA-directed transcription antitermination of the Bacillus subtilis glyQS gene in vitro. J Bacteriol 2004;186 (16):5392-9.
[50] Grundy FJ, Yousef MR, Henkin TM. Monitoring uncharged tRNA during transcription of the Bacillus subtilis glyQS gene. J Mol Biol 2005;346(1):73-81.
[51] Yousef MR, Grundy FJ, Henkin TM. Structural transitions induced by the interaction between tRNA(Gly) and the Bacillus subtilis glyQS T box leader RNA. J Mol Biol 2005;349(2):273-87.
[52] Grundy Fj HS, Rollins SM, Henkin TM. Specificity of tRNA-mRNA interactions in Bacillus subtilis tyrS antitermination. J Bacteriol 1997;179(8):2587-94.
[53] Gerdeman MS. Invitrostructure-function studies of the Bacillus subtilis tyrS mRNA antiterminator: evidence for factorindependent tRNA acceptor stem binding specificity. Nucleic Acids Res 2002;30:1065-72.
[54] Mars RA, Nicolas P, Denham EL, van Dijl JM. Regulatory RNAs in Bacillus subtilis: a gram-positive perspective on bacterial RNA-mediated regulation of gene expression. Microbiol Mol Biol Rev 2016;80(4):1029-57.
[55] Yakhnin H, Yakhnin AV, Babitzke P. Ribosomal protein L10(L12)4 autoregulates expression of the Bacillus subtilis rplJL operon by a transcription attenuation mechanism. Nucleic Acids Res 2015;43(14):7032-43.
[56] Yao Z, Barrick J, Weinberg Z, Neph S, Breaker R, Tompa M, Ruzzo WL. A computational pipeline for high-throughput discovery of cis-regulatory noncoding RNA in prokaryotes. PLoS Comput Biol 2007;3(7):e126.
[57] Choonee N, Even S, Zig L, Putzer H. Ribosomal protein L20 controls expression of the Bacillus subtilis infC operon via a transcription attenuation mechanism. Nucleic Acids Res 2007;35(5):1578-88.
[58] Deiorio-Haggar K, Anthony J, Meyer MM. RNA structures regulating ribosomal protein biosynthesis in bacilli. RNA Biol 2013;10(7):1180-4.
[59] Todd PS, Allen D, Samsel Leigh A, Liu Raymond, Lindahl Lasse, Zengel Janice M. Phylogenetic analysis of L4-mediated autogenous control of the S10 ribosomal protein operon. J Bacteriol 1999;181:6124-32.
[60] Babitzke P. Regulation of transcription attenuation and translation initiation by allosteric control of an RNA-binding protein: the Bacillus subtilis TRAP protein. Curr Opin Microbiol 2004;7(2):132-9.
[61] Hubner S, Declerck N, Diethmaier C, Le Coq D, Aymerich S, Stulke J. Prevention of cross-talk in conserved regulatory systems: identification of specificity determinants in RNA-binding anti-termination proteins of the BglG family. Nucleic Acids Res 2011;39(10):4360-72.
[62] Langbein I, Bachem S, Stülke J. Specific interaction of the RNA-binding domain of the Bacillus subtilis transcriptional antiterminator GlcT with its RNA target, RAT. J Mol Biol 1999;293(4):795-805.
[63] Debarbouille M, Arnaud M, Fouet A, Klier A, Rapoport G. The sacT gene regulating the sacPA operon in Bacillus subtilis shares strong homology with transcriptional antiterminators. J Bacteriol 1990;172(7):3966-73.
[64] Idelson A-CO, SacY M. A transcriptional antiterminator from Bacillus subtilis, is regulated by phosphorylation in vivo. J Bacteriol 1998;180(3):660-6.
[65] Schnetz K SJ, Gertz S, Krüger S, Krieg M, Hecker M, LicT Rak B. LicT, a Bacillus subtilis transcriptional antiterminator protein of the BglG family. J Bacteriol 1996; 178(7):1971-9.
[66] Lu Y TR, Switzer RL. Function of RNA secondary structures in transcriptional attenuation of the Bacillus subtilis pyr operon. Proc Natl Acad Sci USA 1996: 14462-7.
[67] Elisabeth Glatz R-P r N. Lars Rutberg and Blanka Rutberg. A dual role for the Bacillus subtilis glpD leader and the GlpP protein in the regulated expression of glpD: antitermination and control of mRNA stability. Mol Microbiol 1996;19(2): 319-28.
[68] Thirumananseri Kumarevel HM, Kumar Penmetcha KR. Structural basis of HutPmediated anti-termination and roles of the Mg21 ion and L-histidine ligand. Nature 2005;434(7030):183-91.
[69] Feng Liu JL, Zhang Tongzhou, Chen Jun, Ho Chun Loong. Engineered sporeforming Bacillus as a microbial vessel for long-term DNA data storage. ACS Synth Biol 2022;11(11):3583-91.
[70] Tamiev D, Lantz A, Vezeau G, Salis H, Reuel NF. Controlling heterogeneity and increasing titer from riboswitch-regulated Bacillus subtilis spores for time-delayed protein expression applications. ACS Synth Biol 2019;8(10):2336-46.
[71] Guan C, Cui W, Cheng J, Zhou L, Liu Z, Zhou Z. Development of an efficient autoinducible expression system by promoter engineering in Bacillus subtilis. Microb Cell Factories 2016;15:66.
[72] Phan TT, Schumann W. Development of a glycine-inducible expression system for Bacillus subtilis. J Biotechnol 2007;128(3):486-99.
[73] Fu G, Yue J, Li D, Li Y, Lee SY, Zhang D. An operator-based expression toolkit for Bacillus subtilis enables fine-tuning of gene expression and biosynthetic pathway regulation. Proc Natl Acad Sci U S A 2022;119(11):e2119980119.
[74] Wang G-N, Lau PS, Li Y, Ye X-S. Synthesis and evaluation of glucosamine-6- phosphate analogues as activators of glmS riboswitch. Tetrahedron 2012;68(46): 9405-12.
[75] Kim JN, Blount KF, Puskarz I, Lim J, Link KH, Breaker RR. Design and antimicrobial action of purine analogues that bind guanine riboswitches. ACS Chem Biol 2009;4(11):915-27.
[76] Findeiß S, Hammer S, Wolfinger MT, Kühnl F, Flamm C, Hofacker IL. In silico design of ligand triggered RNA switches. Methods 2018;143:90-101.
[77] Ceres P, Trausch JJ, Batey RT. Engineering modular 'ON' RNA switches using biological components. Nucleic Acids Res 2013;41(22):10449-61.
[78] Ogawa A, Inoue H, Itoh Y, Takahashi H. Facile expansion of the variety of orthogonal ligand/aptamer pairs for artificial riboswitches. ACS Synth Biol 2023; 12(1):35-42.
[79] Dixon N, Duncan JN, Geerlings T, Dunstan MS, McCarthy JEG, Leys D. J. Micklefield. Reengineering orthogonally selective riboswitches. Proc Natl Acad Sci USA 2010;107(7):2830-5.
[80] Kent R, Dixon N. Systematic evaluation of genetic and environmental factors affecting performance of translational riboswitches. ACS Synth Biol 2019;8(4): 884-901.
[81] Robinson CJ, Vincent HA, Wu MC, Lowe PT, Dunstan MS, Leys D, Micklefield J. Modular riboswitch toolsets for synthetic genetic control in diverse bacterial species. J Am Chem Soc 2014;136(30):10615-24.
[82] Dixon N, Robinson CJ, Geerlings T, Duncan JN, Drummond SP, Micklefield J. Orthogonal riboswitches for tuneable coexpression in bacteria. Angew Chem Int Ed Engl 2012;51(15):3620-4.
[83] Ul Haq I, Muller P, Brantl S. SR7 - a dual-function antisense RNA from Bacillus subtilis. RNA Biol 2021;18(1):104-17.
[84] Celine Reif CLaSB. Bacillus subtilis type I antitoxin SR6 promotes degradation of toxin yonT mRNA and is required to prevent toxic yoyJ overexpression. Toxins 2018;10(2):74.
[85] Sedlyarova N, Shamovsky I, Bharati BK, Epshtein V, Chen J, Gottesman S, Schroeder R, Nudler E. sRNA-mediated control of transcription termination in E. coli. Cell 2016;167(1):111-21.
[86] Silva IJ, Barahona S, Eyraud A, Lalaouna D, Figueroa-Bossi N, Masse E, Arraiano CM. SraL sRNA interaction regulates the terminator by preventing premature transcription termination of rho mRNA. Proc Natl Acad Sci U S A 2019;116(8):3042-51.
[87] Ciampi MS. Rho-dependent terminators and transcription termination. Microbiology 2006;152(9):2515-28.
[88] Bossi L, Figueroa-Bossi N, Bouloc P, Boudvillain M. Regulatory interplay between small RNAs and transcription termination factor Rho. Biochimica et Biophysica Acta (BBA) - Gene Regulatory Mechanisms 2020;1863(7).
[89] Trinquier A, Durand S, Braun F, Condon C. Regulation of RNA processing and degradation in bacteria. Biochim Biophys Acta Gene Regul Mech 2020;1863(5): 194505.
[90] Durand S, Condon C. RNases and helicases in gram-positive bacteria. Microbiol Spectr 2018;6(2).
[91] Bechhofer Dh DM. Bacterial ribonucleases and their roles in RNA metabolism. Crit Rev Biochem Mol Biol 2019;5(3):242-300.
[92] Silvaggi JM, Perkins JB, Losick R. Small untranslated RNA antitoxin in Bacillus subtilis. J Bacteriol 2005;187(19):6641-50.
[93] Jahn N, Brantl S. One antitoxin-two functions: SR4 controls toxin mRNA decay and translation. Nucleic Acids Res 2013;41(21):9870-80.
[94] Jahn N, Preis H, Wiedemann C, Brantl S. BsrG/SR4 from Bacillus subtilis-the first temperature-dependent type I toxin-antitoxin system. Mol Microbiol 2012;83(3): 579-98.
[95] Muller P, Jahn N, Ring C, Maiwald C, Neubert R, Meissner C, Brantl S. A multistress responsive type I toxin-antitoxin system: bsrE/SR5 from the B. subtilis chromosome. RNA Biol 2016;13(5):511-23.
[96] Meissner C, Jahn N, Brantl S. In vitro characterization of the type I toxin-antitoxin system bsrE/SR5 from Bacillus subtilis. J Biol Chem 2016;291(2):560-71.
[97] Jahn N, Brantl S. Heat-shock-induced refolding entails rapid degradation of bsrG toxin mRNA by RNases Y and J1. Microbiology (Read) 2016;162(3):590-9.
[98] Durand S, Braun F, Helfer AC, Romby P, Condon C. sRNA-mediated activation of gene expression by inhibition of 5'-3' exonucleolytic mRNA degradation. Elife 2017;6:e23602.
[99] Pinel-Marie ML, Brielle R, Riffaud C, Germain-Amiot N, Polacek N, Felden B. RNA antitoxin SprF1 binds ribosomes to attenuate translation and promote persister cell formation in Staphylococcus aureus. Nat Microbiol 2021;6(2):209-20.
[100] Kang SM, Kim DH, Jin C, Lee BJ. A systematic overview of type II and III toxinantitoxin systems with a focus on druggability. Toxins 2018;10(12):515.
[101] Brantl S, Muller P. Cis- and trans-encoded small regulatory RNAs in Bacillus subtilis. Microorganisms 2021;9(9):1865.
[102] Smaldone GT, Revelles O, Gaballa A, Sauer U, Antelmann H, Helmann JD. A global investigation of the Bacillus subtilis iron-sparing response identifies major changes in metabolism. J Bacteriol 2012;194(10):2594-605.
[103] Durand S, Callan-Sidat A, McKeown J, Li S, Kostova G, Hernandez-Fernaud JR, Alam MT, Millard A, Allouche D, Constantinidou C, Condon C, Denham EL. Identification of an RNA sponge that controls the RoxS riboregulator of central metabolism in Bacillus subtilis. Nucleic Acids Res 2021;49(11):6399-419.
[104] McKellar SW, Ivanova I, Arede P, Zapf RL, Mercier N, Chu LC, Mediati DG, Pickering AC, Briaud P, Foster RG, Kudla G, Fitzgerald JR, Caldelari I, Carroll RK, Tree JJ, Granneman S. RNase III CLASH in MRSA uncovers sRNA regulatory networks coupling metabolism to toxin expression. Nat Commun 2022;13(1): 3560.
[105] Calogero S, Gardan R. A novel regulatory protein controlling arginine utilization in Bacillus subtilis, belongs to the NtrC/NifA family of transcriptional activators. J Bacteriol 1994;176(5):1234-41.
[106] Rozenn Gardan GR, Débarbouillé Michel. Role of the transcriptional activator RocR in the arginine-degradation pathway of Bacillus subtilis. Mol Microbiol 1997; 24(4):825-37.
[107] Smaldone GT, Antelmann H, Gaballa A, Helmann JD. The FsrA sRNA and FbpB protein mediate the iron-dependent induction of the Bacillus subtilis lutABC ironsulfur- containing oxidases. J Bacteriol 2012;194(10):2586-93.
[108] Ul Haq I, Brantl S, Muller P. A new role for SR1 from Bacillus subtilis: regulation of sporulation by inhibition of kinA translation. Nucleic Acids Res 2021;49(18): 10589-603.
[109] Hall H. The sporulation-specific small regulatory RNAs of Bacillus subtilis. University of Warwick; 2017.
[110] Meyer S, Chappell J, Sankar S, Chew R, Lucks JB. Improving fold activation of small transcription activating RNAs (STARs) with rational RNA engineering strategies. Biotechnol Bioeng 2015;113(1):216-25.
[111] Lins M, Amorim L, Correa GG, Picao BW, Mack M, Cerri MO, Pedrolli DB. Targeting riboswitches with synthetic small RNAs for metabolic engineering. Metab Eng 2021;68:59-67.
[112] Yang S, Wang Y, Wei C, Liu Q, Jin X, Du G, Chen J, Kang Z. A new sRNA-mediated posttranscriptional regulation system for Bacillus subtilis. Biotechnol Bioeng 2018; 115(12):2986-95.
[113] Man S, Cheng R, Miao C, Gong Q, Gu Y, Lu X, Han F, Yu W. Artificial transencoded small non-coding RNAs specifically silence the selected gene expression in bacteria. Nucleic Acids Res 2011;39(8):e50.
[114] Kang Z, Wang X, Li Y, Wang Q, Qi Q. Small RNA RyhB as a potential tool used for metabolic engineering in Escherichia coli. Biotechnol Lett 2012;34(3):527-31.
[115] Wadler VC. A dual function for a bacterial small RNA: SgrS performs base pairingdependent regulation and encodes a functional polypeptide. Proc Natl Acad Sci U S A 2007;104(51):20454-9.
[116] Mutalik VK, Qi L, Guimaraes JC, Lucks JB, Arkin AP. Rationally designed families of orthogonal RNA regulators of translation. Nat Chem Biol 2012;8(5):447-54.
[117] Rodrigo G, Landrain TE, Jaramillo A. De novo automated design of small RNA circuits for engineering synthetic riboregulation in living cells. Proc Natl Acad Sci U S A 2012;109(38):15271-6.
[118] Sun D, Chen J, Wang Y, Li M, Rao D, Guo Y, Chen N, Zheng P, Sun J, Ma Y. Metabolic engineering of Corynebacterium glutamicum by synthetic small regulatory RNAs. J Ind Microbiol Biotechnol 2019;46(2):203-8.
[119] Yin G, Peng A, Zhang L, Wang Y, Du G, Chen J, Kang Z. Design of artificial small regulatory trans-RNA for gene knockdown in Bacillus subtilis. Synth. Syst. Biotechnol. 2022;8:61-8.
[120] Cho JS, Yang D, Prabowo CPS, Ghiffary MR, Han T, Choi KR, Moon CW, Zhou H, Ryu JY, Kim HU, Lee SY. Targeted and high-throughput gene knockdown in diverse bacteria using synthetic sRNAs. Nat Commun 2023;14(1):2359.
[121] Boussebayle A, Torka D, Ollivaud S, Braun J, Bofill-Bosch C, Dombrowski M, Groher F, Hamacher K, Suess B. Next-level riboswitch developmentimplementation of Capture-SELEX facilitates identification of a new synthetic riboswitch. Nucleic Acids Res 2019;47(9):4883-95.
[122] Hoetzel J, Suess B. Structural changes in aptamers are essential for synthetic riboswitch engineering. J Mol Biol 2022;434(18).
[123] Kraus L, Duchardt-Ferner E, Bräuchle E, Fürbacher S, Kelvin D, Marx H, Boussebayle A, Maurer L-M, Bofill-Bosch C, Wöhnert J, Suess B. Development of a novel tobramycin dependent riboswitch. Nucleic Acids Res 2023;51(20): 11375-85.
[124] Xiu Y, Jang S, Jones JA, Zill NA, Linhardt RJ, Yuan Q, Jung GY, Koffas MAG. Naringenin-responsive riboswitch-based fluorescent biosensor module for Escherichia coli co-cultures. Biotechnol Bioeng 2017;114(10):2235-44.
[125] Kaiser C, Schneider J, Groher F, Suess B, Wachtveitl J. What defines a synthetic riboswitch? - Conformational dynamics of ciprofloxacin aptamers with similar binding affinities but varying regulatory potentials. Nucleic Acids Res 2021;49(7): 3661-71.
[126] Kohlberger M. SELEX: critical factors and optimization strategies for successful aptamer selection. Biotechnol Appl Biochem 2022;69(5):1771-92.
[127] Findeiß S, Etzel M, Will S, Mörl M, Stadler P. Design of artificial riboswitches as biosensors. Sensors 2017;17(9):1990.
[128] Alexander A, Green Pamela A. Toehold switches: de-novo-designed regulators of gene expression. Cell 2014;159(4):925-39.
[129] To AC-Y, Chu DH-T, Wang AR, Li FC-Y, Chiu AW-O, Gao DY, Choi CHJ, Kong S-K, Chan T-F, Chan K-M, Yip KY, Wren J. A comprehensive web tool for toehold switch design. Bioinformatics 2018;34(16):2862-4.
[130] Heo T, Kang H, Choi S, Kim J. Detection of pks island mRNAs using toehold sensors in Escherichia coli. Life 2021;11(11).
[131] Giakountis A, Stylianidou Z, Zaka A, Pappa S, Papa A, Hadjichristodoulou C, Mathiopoulos KD. Development of toehold switches as a novel ribodiagnostic method for west nile virus. Genes 2023;14(1).
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
© 2024. This work is published under http://creativecommons.org/licenses/by-nc-nd/4.0/ (the “License”). Notwithstanding the ProQuest Terms and Conditions, you may use this content in accordance with the terms of the License.
Abstract
Bacteria exhibit a rich repertoire of RNA molecules that intricately regulate gene expression at multiple hierarchical levels, including small RNAs (sRNAs), riboswitches, and antisense RNAs. Notably, the majority of these regulatory RNAs lack or have limited protein-coding capacity but play pivotal roles in orchestrating gene expression by modulating transcription, post-transcription or translation processes. Leveraging and redesigning these regulatory RNA elements have emerged as pivotal strategies in the domains of metabolic engineering and synthetic biology. While previous investigations predominantly focused on delineating the roles of regulatory RNA in Gram-negative bacterial models such as Escherichia coli and Salmonella enterica, this review aims to summarize the mechanisms and functionalities of endogenous regulatory RNAs inherent to typical Gram-positive bacteria, notably Bacillus subtilis. Furthermore, we explore the engineering and practical applications of these regulatory RNA elements in the arena of synthetic biology, employing B. subtilis as a foundational chassis.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
Details
1 The Science Center for Future Foods, Jiangnan University, Wuxi, 214122, China