Abstract

LysR-type transcriptional regulators (LTTRs) are emerging as a promising group of macromolecules for the field of biosensors. As the largest family of bacterial transcription factors, the LTTRs represent a vast and mostly untapped repertoire of sensor proteins. To fully harness these regulators for transcription factor-based biosensor development, it is crucial to understand their underlying mechanisms and functionalities. In the first part, this Review discusses the established model and features of LTTRs. As dual-function regulators, these inducible transcription factors exude precise control over their regulatory targets. In the second part of this Review, an overview is given of the exceptions to the “classic” LTTR model. While a general regulatory mechanism has helped elucidate the intricate regulation performed by LTTRs, it is essential to recognize the variations within the family. By combining this knowledge, characterization of new regulators can be done more efficiently and accurately, accelerating the expansion of transcriptional sensors for biosensor development. Unlocking the pool of LTTRs would significantly expand the currently limited range of detectable molecules and regulatory functions available for the implementation of novel synthetic genetic circuitry.
Keywords: transcription factors, LysR-type transcriptional regulators, biosensors, synthetic biology, genetic circuitry, prokaryotes
Introduction
The first synthetic genetic circuits was created at the turn of the millennium1,2 and brought forth the field of synthetic biology.3 The rapid progress of this field is largely attributed to integration of knowledge from diverse disciplines, including microbiology, genetics, genomics, chemistry, physics, information technology, along with general biology and engineering.3 As the field rapidly transitions from its nascent stages to a more advanced scientific discipline, an array of new applications and tools has come forth.4−8 Among these synthetic biology achievements, biosensors have received a great deal of attention over the last years. By leveraging the extensive sensory apparatuses from microorganisms, biosensors enable monitoring of intracellular processes and environmental signals, as well as regulating gene expression in a precise and tunable way.
Such biosensors are composed of input, processing and output parts, as shown in Figure 1A. The input can be a wide range of different sensor molecules, e.g., riboswitches, ribozymes, two-component systems and transcription factors. This review focuses on transcriptional biosensors, more specifically on allosteric transcription factors (TFs). Although two-component systems have demonstrated their utility in biosensing extracellular signals,9−11 one-component TFs have proven to be highly useful and efficient in a diverse set of synthetic genetic circuits.12−15 In addition, the bacterial TF landscape is vast and diverse, with a higher abundance of regulators in comparison to two-component systems,16 where they exhibit important physiological functions via an array of regulation mechanisms. This large variety makes TFs an ideal source of input sensor parts.
Figure 1.
(A) Schematic overview of transcription factor-based biosensors and their composition. For simplicity, the illustration is limited to inducible activators. The input is comprised of a transcription factor that can sense a ligand of interest. Signal processing occurs during DNA–protein interaction, where the transcription factor alters the expression of a target promoter depending on the input ligand concentration. This target promoter is responsible for the expression of an output gene that is chosen based on the envisioned biosensor application. (B, C) The two defining characteristics of these transcription factor-based biosensors are the response curve and ligand specificity. The response curve describes how biosensors link sensed ligand concentration to the output of choice. The ligand specificity determines how this response curve varies for different ligands. The response curves are given for a range of ligands, wherein the transcription factor exhibits low specificity (gray), as opposed to the high specificity observed for the ligand of interest (orange). Genetic circuit parts are given according to SBOL conventions.41,42 TF = transcription factor, TFBS = transcription factor binding site, FP = fluorescent protein, LOQ = limit of quantification, ULOQ = upper limit of quantification.
For TF-based biosensors, processing takes place at the protein–DNA interface, i.e., the transcription factor binding sites (TFBS), as shown in the middle panel of Figure 1A. Here, the regulator alters transcription of an output of choice, based on the presence of an input signal in the form of a ligand. By changing the output signal, biosensors can be used for different applications, which have been extensively reviewed elsewhere, including high-throughput screening of intracellular metabolites,17−21 biomedical diagnostics,22,23 environmental pollutant monitoring,22−25 adaptive laboratory evolution26−31 and dynamic pathway control.17,32−40
The two main characteristics that define biosensor circuits are the response curve (see Figure 1B) and ligand specificity (see Figure 1C), and are highly dependent on the combination of input, processing and output parts. The response curve of a biosensor describes the relation between a sensed input and the resulting output.20,43−45 Within a limited range of ligand concentrations, known as the operational range (see highlighted x-axis range in Figure 1B), a biosensor’s response undergoes significant changes, constituting its dynamic range (see highlighted y-axis signal range in Figure 1B).44 The lowest limit of quantification (LOQ) is preceded by the “OFF”-state of the biosensor, where any observed output is solely caused by leaky expression of the target promoter, rather than ligand recognition. The “ON” state is reached at the upper limit of concentration quantification (ULOQ). Higher ligand concentrations will not result in a further increase of the biosensor output due to saturation of the system.20,44 The rate at which the output varies with increasing inducer concentrations characterizes the biosensor’s sensitivity, as shown as the slope in Figure 1B.46,47 It is strongly influenced by the effect of multiple ligand molecules binding to the transcriptional regulator, which is referred to as cooperativity.45 Finally, specificity of a TF can be defined as the affinity of this regulator for a ligand compared to other small molecules, resulting in a more pronounced change in output signal for this ligand.46,47 This is shown in Figure 1C, where only one ligand results in a significant change in output within the relevant concentration range.
Despite the rich diversity in transcriptional regulators available in nature, and the advancements made on the customization of their characteristics toward applications within the biosensor context, a true repertoire of input sensors is still missing, both in terms of ligand specificity and regulatory mechanism.17,48,49 To address this challenge, the mostly untapped potential of LysR-type transcriptional regulators (LTTRs), the largest family of bacterial TFs, is being explored. While they show interesting features for biosensor development, including a broad range of ligands and tight gene regulation, research and application of LTTRs has been relatively limited. This is partially ascribed to the complexity of their regulatory systems and the scattered available information. In the following sections, an overview of LTTR family characteristics is provided, shedding light on their intricate mechanisms to illustrate their utility in synthetic genetic circuits, particularly in the context of biosensors.
The LysR-type Transcriptional Regulators as Key Regulators for Bacterial Life
The one-component TFs are further classified in protein families, of which the LTTRs constitutes the largest collection of bacterial transcriptional regulators known to date.50−52 The family was established in 1988, when nine transcriptional regulators where grouped due to similarities in their DNA-binding domain (DBD),53,54 and named after the most well studied TF of the group at the time, i.e., lysine biosynthesis regulator LysR.55,56 Since then, the group has grown tremendously, with over 852,000 predicted LTTR sequences reported on the InterPro database (IPR000847 ‘Transcription regulator HTH, LysR’). Similarly to other TF families, newly discovered regulators are assigned as LTTRs due to sequence similarities, mainly in the highly conserved DBD, where all LTTRs show a winged helix-turn-helix (wHTH) motif. The family is ubiquitous in both Gram-positive and Gram-negative bacteria,57,58 and functional orthologues are found in archaea and algal chloroplasts.57,59−63 In addition to their widespread occurrence, LTTRs can represent up to 20% of the regulators within one microorganism.51,64 As an example, Pseudomonas aeruginosa PA01 is reported to contain over 100 different LTTRs.65 The group most likely emerged from a common ancestor at an early evolutionary stage, given the widespread distribution of the regulators over extensive genetic distances. The evolutionary complexity of the group is further underscored by its ability to recognize a wide array of molecules, which requires a significant number of evolutionary steps to be achieved.66,67
The diversity and abundance of this TF family resembles their involvement in all aspects of the microbial life.68−70 LTTRs play an important role in the regulation of key processes often revolving around the cellular response to environmental changes. This also becomes apparent from Table 1, which gives an overview of the currently characterized LTTRs and their biological relevance. It is hypothesized that the evolution of LTTRs was driven by the necessity to respond to these environmental stimuli, hence their ubiquitous presence, especially in free-living microorganisms. Due to continuous exposure to fluctuating environments, such microorganisms require constant adaptations, fueling the evolution of regulatory molecules such as LTTRs.67 Examples of key pathways regulated by members of the LTTR family include stress response,57,71 motility and attachment,56,57,66,71,72 quorum sensing and metabolic signaling,57,71 antibiotic resistance,58,73 fixation of both CO2 and N2,74 biosynthesis of amino acids,75,76 catabolism of aromatic compounds,56,57,66,71,72 transport and secretion,56,57,66,71,72 virulence factor expression,77 initiation of nodulation78 and cell division.79 Taken together, the family’s substantial size and involvement in a wide diversity of pathways indicates a large potential for biotechnological applications, as it encompasses sensors responsive to an extensive range of compounds. This is further substantiated by the large amount of LTTRs in Actinobacteria, Proteobacteria, and Firmicutes, three phyla that are known for the synthesis of industrially relevant bioproducts.80 Examples of such interesting molecules sensed by members of the LTTR family are given in Table 1.
Table 1. Overview of the Currently Characterized LysR-type Transcriptional Regulators and the Different Elements Required for Converting Them into Transcription Factor-Based Biosensorsa.
| identifier | origin | biological relevance | target ligand | regulatory target | reference |
|---|---|---|---|---|---|
| AalR | Acinetobacter baylyi ADP1 | aspartate metabolism | l-aspartate | PaspA, PaspT, PaspY, PracD | (81) |
| AceR | Acinetobacter baumannii | multidrug efflux pump regulation | chlorhexidine | PaceI | (82, 83) |
| AdmX | Serratia plymuthica A153 | antibiotic synthesis regulation | indole-3-acetic acid | PadmV (operon) | (84−86) |
| AllS | Escherichia coli | anaerobic allantoin metabolism | allantoin | PallD | (87−89) |
| AlsR | Bacillus subtilis | acetoin synthesis | acetate | PalsS | (73, 90−92) |
| AmpR | Citrobacter freundii | β-lactamase synthesis | 1,6-anhydro-MurNAc-peptides | PampC | (73, 93−98) |
| AphB | Vibrio cholerae | virulence | unknown | PtcpPH | (99−103) |
| AphT | Comamonas testosteroni TA441 | aromatic compound degradation | 2-hydroxymuconic semialdehyde | PaphCEFGHJI | (104) |
| ArgP | Escherichia coli | arginine transport | arginine | PargK | (105−107) |
| AtzR | Pseudomonas spp. | cyanuric acid catabolism | cyanuric acid | PatzDEF | (25, 108, 109) |
| BauR | Pseudomonas aeruginosa | polyamine utilization | β-alanine | PbauABCD | (110) |
| BenM | Acinetobacter spp. | aromatic compound degradation | cis,cis-muconate or benzoate | PbenABCDE, PbenPK | (111−118) |
| BlaA | Streptomyces spp. | β-lactamase synthesis | β-lactam compounds | PblaL | (119) |
| CatM | Acinetobacter calcolaceticus | catechol catabolism | cis,cis-muconate | PcatBCIJFD, PcatA, PbenPK | (71, 112, 120, 121) |
| CatR | Pseudomonas putida | catechol catabolism | cis,cis-muconate | PcatBCA, PpheBA | (71, 122−126) |
| CbbR | Xanthobacter flavus | carbon dioxide fixing | ribulose 1,5-bisphosphate | PcbbLSXFPTAE | (74, 127, 128) |
| Cbl | Escherichia coli | sulfur acquisition and reduction | thiosulfate | PcysP, PcysT, PtauABCD, PssuEADCB | (129−131) |
| CbnR | Ralstonia spp. | chlorocatechol catabolism | 2-chloro-cis,cis-muconate | PcbnABCD | (17, 71, 132−134) |
| CcpC | Bacillus subtilis | carbon catabolite control | citrate | PcitB, PcitZ | (135, 136) |
| CcpE | Staphylococcus aureus | carbon catabolite control | unknown | PcitE | (30, 137) |
| CfxR | Alcaligenes eutrophus | carbon dioxide fixing | unknown | PcfxLSXYEFP | (138) |
| ChiR | Serratia marcescens | chitin degradation | unknown | PCBP21 | (139) |
| CidR | Staphylococcus aureus | murein hydrolase regulation | acetic acid | PcidABC | (140) |
| CitR | Bacillus subtilis | citrate metabolism | unknown | PcitA | (141) |
| ClcR | Pseudomonas putida | chlorocatechol catabolism | 2-chloro-cis,cis-Muconate | PclcABD | (71, 123, 142, 143) |
| CmpR | Synechococcus elongatus | CO2 fixing | ribulose 1,5-bisphosphate | PcmpABCD | (144−146) |
| CrgA | Neisseria meningitidis | pili/capsule synthesis and oxidative stress response | α-methylene-γ-butyrolactone | PmdaB | (147−149) |
| CynR | Escherichia coli | cyanate detoxification | cyanate | PcynTSX | (150) |
| CysB | Salmonella enterica serovar Typhimurium | cysteine biosynthesis | N-Acetylserine | PcysI | (130, 151−155) |
| CysL | Bacillus subtilis | sulphite reductase | sulfate, sulfite, thiosulfate | PcysJI | (156) |
| DarR | Acinetobacter baylyi ADP1 | aspartate metabolism | d-aspartate | PaspA, PaspT, PaspY, PracD | (157) |
| DbdR | Thauera aromatica AR-1 | aromatic compound degradation | 3,5-dihydroxybenzoate | PdbhL | (158) |
| DntR | Burkholeria spp. | aromatic compound degradation | 2,4-dinitrotoluene, salicylate | PdntA | (50, 159−161) |
| FdeR | Herbaspirillum seropedicae | aromatic compound degradation | naringenin | PfdeA | (162, 163) |
| FinR | Pseudomonas putida | ferredoxin-NADP+ reductase regulation | oxidative and osmotic stresses | Pfpr | (164) |
| GcvA | Escherichia coli | glycine metabolism | glycine and adenine | PgcvTHP, PgcvB | (165, 166) |
| GigC | Acinetobacter baumannii | cysteine metabolism and virulence | l-cysteine | PcysI | (167) |
| GltC | Bacillus subtilis | glutamate synthase | α-ketoglutarate | PgltA | (168−170) |
| HexA | Photorhabdus luminescens | phenotypic heterogeneity, pathogen–symbiont transition | unknown | PpcfABCDEF | (171) |
| HsdR | Comamonas testosteroni | aromatic compound degradation | testosterone | PhsdA | (172, 173) |
| HupR | Vibrio vulnificus | haem uptake | haemin | PhupA | (174) |
| HvrB | Rhodobacter capsulatus | S-adenosyl-l-homocysteine hydrolase expression | light sensitivity | PahcY | (175) |
| HypT | Salmonella enterica serovar Typhimurium | HOCl response | HOCl | PfhuA | (176) |
| IlvR | Caulobacter crescentus | isoleucine/valine biosynthesis | unknown | PilvD | (177) |
| IlvY | Escherichia coli | isoleucine/valine biosynthesis | α-acetolactate, α-acetohydroxy-butyrate | PilvC | (178) |
| IrgB | Vibrio cholerae | iron-regulated virulence factor | unknown | PirgA | (179) |
| KaeR | Lactobacillus brevis | flavonoid response | kaempferol | PLVIS1988 | (180) |
| LeuO | Salmonella enterica serovar Typhimurium | bacterial stringent response | unknown | PompS1, PompS2, PassT | (70, 181−183) |
| LinR | Sphingomonas spp. | aromatic compound degradation | 2,5-dichlorohydroquinone, 2,6-dichlorohydroquinone, chlorohydroquinone | PlinE | (71, 184, 185) |
| LrhA | Escherichia coli | flagella, motility and chemotaxis | unknown | PflhDC | (186) |
| LttR | Lactobacillus plantarum | conjugated linoleic acid production | linoleic acids | Pcla-dh (operon) | (187) |
| LysG | Corynebacterium glutamicum | amino acid export | l-lysine and other amino acids | PlysE | (188, 189) |
| LysR | Escherichia coli | lysine biosynthesis | diaminopimelate | PlysA | (55) |
| MdcR | Klebsiella pneumoniae | malonate catabolism | malonate | PmdcABCDEGHLM | (190, 191) |
| MetR | Streptococcus spp. | methionine and cysteine transport/biosynthesis | homocysteine | PmetE | (192−198) |
| MleR | Streptococcus mutans | malolactic fermentation | l-malate | PmleSP | (199) |
| MvfR (= PqsR) | Pseudomonas aeruginosa | pathogenicity regulator | 4-hydroxy-2-heptylquinone | PpqsA | (200−205) |
| NagR | Ralstonia eutropha | naphthalene catabolism | salicylate | PnagA | (206, 207) |
| NahR | Pseudomonas putida | naphthalene/salicylate catabolism | salicylate | PnahABCFDE, PsalGHINL | (208−210) |
| NdhR (= CcmR) | Synechocystis spp. | CO2-concentrating mechanism | 2-phosphoglycolate, 2-oxoglutarate | Pndh3 (operon) | (72, 211) |
| NhaR | Escherichia coli | Na+/H+ antiporter regulation | Na+ | PnhaA | (212) |
| NocR | Agrobacterium tumefaciens | nopaline catabolism | octopine | PnocPTQM | (213) |
| NodD | Rhizobium spp. | nitrogen fixation/symbiosis | flavonoids | PnodABCIJ, PnodFEL, PnodMNT, PnodO | (78, 214−217) |
| NtdR | Acidovorax spp. | aromatic compound degradation | 2,4-dinitrotoluene, 2,6-dinitrotoluene, salicylate, anthranilate | PntdA | (71) |
| OccR | Agrobacterium tumefaciens | octopine catabolism | octopine | PoccQMPT | (218−222) |
| OdcR | Pigmentiphaga spp. | 3,5-dibromo-4-hydroxybenozate catabolism | 3-bromo-4-hydroxybenzoate, 3,5-dibromo-4-hydroxybenozate | PodcA, PodcB, PodcC | (223) |
| OxyR | Escherichia coli | oxidative stress response | redox changes | PkatG, PaphCF, PoxyS, PgorA, PgrxA | (224−231) |
| PA2206 | Pseudomonas aeruginosa | oxidative stress response | redox changes | PpvdS | (232) |
| PcaQ | Agrobacterium spp. | protocatechuate catabolism | β-carboxy-cis,cis-muconate, γ-carboxymuconate | PpcaDCHGB | (71) |
| PecT | Dickeya dadantii | virulence | temperature changes | PpelB, PpelC, PpelD, PpelE | (233, 234) |
| PhcA | Pseudomonas solanacearum | virulence | unknown | PT3SS | (235, 236) |
| PltR | Pseudomonas protegens | pyoluteorin synthesis | pyoluteorin | PpltLABCDEFG | (237, 238) |
| QuiR | Listeria monocytogenes | protocatechuate biosynthesis | shikimate | Plmo0489 (operon) | (239) |
| RbcR | Chromatium vinosum | carbon dioxide fixing | unknown | PrbcAB | (240) |
| RipR | Salmonella enterica | itaconic acid resistance | isocitrate | PripCBA | (241,242) |
| RovM | Yersinia pseudotuberculosis | invasion/motility, virulence | unknown | ProvA | (243, 244) |
| S cmR | Burkholderia thailandensis | secundary metabolism regulation | unknown | Phmq (operon) | (68, 69) |
| SpvR | Salmonella typhimurium | virulence | unknown | PspvABCD | (245, 246) |
| SyrM | Rhizobium meliloti | exopolysaccharide synthesis | unknown | PsyrA | (247) |
| TcbR | Pseudomonas spp. | chlorocatechol metabolism | 2-chloro-cis,cis-muconate | PtcbCDEF | (71, 248) |
| TfdR | Ralstonia spp. | chlorophenol catabolism | 2-chloro-cis,cis-muconate | PtfdCDEFB, PtfdA | (71) |
| TfdT | Burkholderia spp. | aromatic compound degradation | 3-chlorocatechol, 4-chlorocatechol, 2-chlorobenzoate, 3-chlorobenzoate | PtfdC | (71) |
| ThnR | Sphingopyxis granuli | aromatic compound degradation | tetralin | PthnB, PthnC | (249, 250) |
| ToxR | Burkholderia glumae | toxoflavin biosynthesis | toxoflavin | PtoxABCDE, PtoxFGHI | (251) |
| TsaR | Comamonas testosteroni | aromatic compound degradation | p-tolunesulfonate | PtsaMBCD | (252−254) |
| TtuA | Agrobacterium vitis | tartrate metabolism | l-tartrate | PttuB | (255) |
| VirR | Rhodococcus equi | virulence | unknown | Porf5 (virR operon) | (190) |
| VqsA | Vibrio alginolyticus | virulence | unknown | PaphA, PluxR | (256) |
| VtlR (= LsrB) | Agrobacterium tumefaciens | regulation of motility and biofilm formation | unknown | PabcR2 | (257−259) |
| YbeF | Escherichia coli | flagella biosynthesis, putative citrate utilization related 2 | unknown | unknown | (260) |
| YbdO | Escherichia coli | citrate utilization related, flagella biosynthesis | unknown | PybdO | (260) |
| YbhD | Escherichia coli | l-malate utilization related | unknown | PybhH, PybhI | (260) |
| YcaN | Escherichia coli | unknown | unknown | PycaC, PycaD | (260) |
| YeiE | Cronobacter sakazakii | sulfite resistance | sulfite | PcysJI | (261) |
| YgfI | Escherichia coli | dihydroxyacetone, glycerol or Thr utilization | unknown | PdhaK PyjiT, PpflB, PadhE, PhycBCDEF, PnarZ | (260) |
| YiaU | Escherichia coli | membrane modification and LPS biosynthesis | unknown | PwaaPSBOJYZU PyjiT, PadeP, PyiaT, PgltD | (260) |
| YneJ | Escherichia coli | putrescine utilization related | unknown | Psad,PfnrS | (260) |
| YofA | Bacillus subtilis | cell division | unknown | PftsW | (79) |
| YtxR | Yersinia entercolitica | ADP-ribosyl-transferase toxin | unknown | PytxAB | (262) |
For each regulator, the organism of origin is given, as well as the biological context in which it is active. When available, a ligand recognized by the regulator is provided, as well as target promoters. As these transcription factors often function as global regulators, the “regulatory target” column does not represent all target promoters.
The large variety in bacterial sensor molecules within the LTTR family, shown in Table 1, is highly advantageous for biosensor development. It results in a broad range of input molecules that are readily sensed by transcriptional regulators. Moreover, it makes the LTTRs an interesting starting point for engineering new specificities, as this is done more efficiently when starting from similar compounds. For instance, this approach was utilized in developing biosensors for various flavonoids,163 or in engineering the specificity of BenM toward adipic acid.263 Beyond their diversity, LTTRs are known for stringent regulation over their target genes by binding promoters in both induced and uninduced states.57,264−266 Next to their role as transcriptional activators upon ligand interaction, continuous association with target promoters causes repression in the uninduced state due to steric hindrance on key promoter elements. Additionally, this causes negative autoregulation, given that most LTTRs are colocalized with a member of their regulon.57 This mechanism of transcriptional regulation results in low leaky expression and a fast response to inducing ligands,266 making them ideal inputs for biosensor development. The optimization of naringenin production in E. coli showed the importance if this precise gene control. The concentration of precursor malonyl-COA required careful balancing to prevent toxicity,267 which was achieved by embedding FdeR, a naringenin responsive LTTR, into a multilayer dynamic control system.40
Despite the progress made in development of LTTR-based biosensors, it is apparent that the amount of characterized LTTRs and their applications is disproportional to the overall size of this TF family. This is largely due to the limited and scattered knowledge on these transcriptional regulators. Without a clear understanding of their functionalities and properties, applying them in specific contexts is challenging. Moreover, most applications necessitate optimizing the response curve to transform natural TF systems into effective sensors for the intended purpose. Therefore, comprehending the mechanisms underlying transcriptional regulation by LTTRs is essential for unlocking the full potential of this TF family.
A general mechanism for LTTR regulation has been proposed, but many exceptions to this rule exist, underscoring its limitations. Given the extensive size of the LTTR family, it is plausible that the currently ascribed mechanism is only applicable to a subset of the regulators. Therefore, it becomes necessary to reclassify the LTTR family into distinct subfamilies, which will facilitate a more precise allocation of structure and function. The subsequent sections will explore the ’classic’ LTTR regulatory mechanism, starting with an overview of the characteristic ’sliding dimer’ mechanisms. This will be followed by a more detailed discussion on the architecture of regulatory targets of LTTRs, as well as the structural properties of these proteins. Finally, an overview of the currently identified exceptions to this ’classic’ model will be presented, underscoring the need to re-evaluate this extensive transcription factor family.
Overview of the “Sliding Dimer” Mechanism
The regulatory mechanism of the LTTRs is referred to as “sliding dimer” and is visualized in Figure 2. When expressed, LTTR proteins undergo oligomerization, forming dimers that serve as the functional unit for DNA interaction. The targeted DNA sequence for this protein–DNA interaction comprises three binding sites, as shown in Figure 2A. In the absence of the inducing ligand, LTTR dimers interact with the regulatory site (RS) and activation site 1 (AS1). RS serves as an anchor due to highly stable protein–DNA interactions. In contrast, dimers bound to AS1 sites form stable but loose DNA interactions, which is important for the regulatory mechanism.57,253 A dimer can only bind to AS1 while simultaneously dimerizing with the RS-bound dimer, a process known as cooperative binding.108,127,253,268 The tetrameric protein complex covers key promoter regions, as AS1 is situated near the −10 box of the target gene promoter. In addition, the TF-DNA interaction induces DNA bending between 50° and 100° due to the distance between the AS1 site and the RS site, the latter situated around −65 nucleotides from the transcription start site. Both the binding of key sequences and DNA bending cause inaccessibility of the target promoter, preventing RNA polymerase (RNAP) from interacting with it,116,127,265 as illustrated in Figure 2B.
Figure 2.

The regulatory mechanism of the LysR-type transcriptional regulators is referred to as the “sliding dimer” mechanism, owing to the movement of transcription factor dimers upon ligand interaction. (A) As most LysR-type transcriptional regulators (blue) are expressed from a bidirectional promoter region with one of their target genes (gray), this genetic architecture is given in more detail. (B) In the uninduced state, the tetrameric protein complex binds regulatory site (RS) and activation site 1 (AS1), causing the DNA to bend and preventing the RNA polymerase (RNAP) complex to recognize the promoter sequence. In addition, the location of the transcription factor binding sites overlap with key promoter regions (−10 and −35 boxes) of the target gene, further preventing RNAP complex–DNA interactions via steric hindrance. Due to overlap in the expression systems of the regulator and target gene, the transcription factor also causes negative autoregulation. Upon recognition of the appropriate ligand, the protein complex undergoes conformational changes that forces the dimer bound at AS1 to move to activation site 2 (AS2). These conformational changes are illustrated in a frontal and top view, showing the movement of the monomers within the tetrameric structure upon ligand interaction. As a result, the DNA bending is reduced and RNAP interaction sites become exposed, enabling activation of the target promoter transcription. On the surface of the ligand-bound regulator, RNAP recruitment sites appear after the conformational changes, further facilitating and guiding the interaction between RNAP and the target promoter. Genetic circuit parts are given according to SBOL conventions.41,42 TF = transcription factor, RS = regulatory site, AS = activation site, RNAP complex = RNA polymerase with appropriate sigma factor, LTTR = LysR-type transcriptional regulator.
Upon ligand interaction, the tetrameric structure tightens due to interactions among residues within the ligand-binding domain (LBD) and the ligand. These conformational modifications within the protein structure are transduced to the DBD,116 and forces the dimer bound to AS1, to transition to the AS2 site.50,116,127 The latter is located upstream, proximate to the −35 box, but beyond the crucial promoter regions. By moving to AS2, the protein complex exposes the promoter for RNAP, and reduces the DNA bending to a range of 9–50°. Furthermore, the induced tetramer actively recruits RNAP. The conformational changes cause a migration of internal protein sequences to the tetramer surface, which interact with the α-subunits of RNAP.116,127,253,268,269
Transcription Factor Binding Sites and Gene Architecture
The regulatory mechanism of LTTRs involves three different sites for TF interactions, namely the high affinity regulatory binding site RS and lower affinity activating binding sites AS1 and AS2, with the latter two usually found together and often overlapping in sequence.249 This is visualized in Figure 3, which shows the genetic architecture of target promoter regions of LTTRs. The differences in affinity between binding sites arise from variations in the nucleotide sequence,116,223 as discussed below and shown in Figure 3. The RS site, usually situated between −80 to −50 bp from the target gene transcription initiation site, is seen as an anchor for the TF complex.66,127,253,268 Its high affinity allows the interacting dimer to form a stable interaction, subsequently enabling a second dimer to bind to the lower affinity AS1 sequence. The latter is situated downstream of the RS site and overlaps key transcription control elements of the targeted promoter, such as the −35 and −10 boxes. While binding of an LTTR dimer onto the AS1 site can occur, it is believed to be unstable until the dimer oligomerizes with the RS bound dimer.56 This loose interaction between the regulator complex and AS1 enables the “sliding dimer” mechanism upon ligand induction, as shown in Figure 2. In addition to the differences in sequence within one promoter, TFBS sequences can also vary over different target promoters. This allows divergent regulation of LTTRs over their regulon, adding an extra layer to the transcriptional regulation.
Figure 3.

Overview on the transcription factor binding sites of LysR-type transcriptional regulators. The target promoters of these regulators show three different binding sites, namely the regulatory site (RS) and two activation sites (AS1 and AS2). These regulatory sites follow the characteristic T-N11-A sequence, with symmetry in the nucleotides surrounding the thymine and adenine of this sequence. During DNA binding, the RS site is first bound by a dimer of regulator molecules. The high degree of symmetry in this site, as shown in bold for regulators BenM, DntR, and OccR, allows for a stable and strong interaction between the dimer and the DNA. A second dimer binds AS1, which shows a much lower degree of dyad symmetry. Hence, binding at this site is only possible while simultaneously forming the full tetrameric structure together with the RS-bound dimer. Due to the distance between RS and AS1, the tetramerization and DNA binding is accompanied by DNA bending. Similarly to the variation in AS1 sequence, AS2 also shows imperfect symmetry. The AS1 and AS2 sites cover important promoter regions, such as the −10 and −35 box. The location of these boxes, as well as the transcription start site, are given per promoter for both the target promoter system (gray) and the transcription factor expression system (blue). Genetic circuit parts are given according to SBOL conventions.41,42 TF = transcription factor, RS = regulatory site, AS = activation site.
In general, TFBS sequences of LTTRs have a characteristic T-N11-A conserved sequence that shows imperfect dyad symmetry.66,265,270 The degree of symmetry within the recognition region determines the strength of TF binding,66,127,253,268,271 with a certain degree of asymmetry required for effective regulation.272 The T-N11-A motive is important for guiding the TF to the correct DNA sequence, but true recognition and specificity of TF-DNA interactions is caused by the surrounding nucleotides.66 For example, the recognition logo for BenM was expanded to ATAC-N7-GTAT due to the importance of these neighboring nucleotides,116 which is depicted in Figure 3. To enable the sliding of a dimer from AS1 to AS2, the interaction at these sites should be stable but weaker than the RS-dimer interaction. To facilitate this, AS sequences tend to show less dyad symmetry than the RS sequence, as seen for BenM in Figure 3, where both AS sites show one nucleotide difference with the RS site. These variations can be more drastically different, as is seen for regulator DntR. The AS sites of PdntA, which are given in Figure 3, could only be found via experimental methods, as they do not resemble the classic LTTR binding site sequence. In contrast, the RS shows a high dyad symmetry T-N11-A sequence.50 Similarly, the discovery of AS sites of OccR was impaired due to overlap of these sites.213,218,221 Nevertheless, this information is important for optimization of LTTR-based biosensors, as different parameters of the response curve, discussed in Figure 1B, can be significantly altered by changing the expression level of either the TF or target promoter, or altering the TFBS sequences.20,162,273
Initially, due to their co-occurrence with target genes, as illustrated in Figure 3, LTTRs were described as cis-acting regulators. Further research into the regulons of various LTTR members has unveiled that a significant number of them function as global regulators.158,265 Much of their transcriptional regulation is thus done as trans-acting regulators, but their colocalization with a specific target gene remains a characterizing feature of the family.57 This genomic architecture results in the TF divergently transcribed from a target gene, surrounding a bidirectional promoter region with overlapping operator sequences, as shown in Figure 3. This is a common architecture in bacteria, which promotes efficient regulation of multiple genes due to the TFBSs spanning both promoters.178,274 Hence, upon binding the target promoter, the LTTR simultaneously represses its own expression, also referred to as negative autoregulation (see Figure 2B). The mechanism behind the autoregulation is based on steric hindrance when the TF is bound to the bidirectional promoter,66,172,178,222 as the RS of the target gene overlaps with key regulatory elements of the TF promoter.116 This is postulated to prevent overproduction of TFs, which could put an unnecessary burden upon the cell.60,127,219,221,225,253,274 It has also been shown that negative autoregulation in combination with weak expression systems controlling TF synthesis, results in an increased sensitivity, while simultaneously reducing response time and lowering leaky expression.266 Both parameters are essential for a swift reaction upon ligand sensing, which is of great importance for the vital biological functions regulated by LTTRs.
In summary, the ’sliding dimer’ mechanism of LTTRs involves interactions at high-affinity regulatory binding sites and lower-affinity activating sites. These sites typically feature a T-N11-A motif with varying dyad symmetry affecting binding strength and regulatory effectiveness. Next to the DNA–protein interactions, transcriptional regulation by LTTRs also depends on ligand-protein and protein–protein interactions, as discussed below.
Protein Structure and Oligomerization of LysR-type Transcriptional Regulators
At the TFBS, multiple monomers oligomerize into a tetrameric protein structure that forms the interface between the input ligands and output gene expression. This is achieved by a complex interplay between the different protein domains that make up LTTRs. For example, recent work on the connecting region between the DBD and LBD, referred to as the linker helix, showed that this connector has big influences on both DNA- and ligand recognition and binding.275 Furthermore, there are several regions on the LTTRs that are essential for their regulatory mechanism, such as the regions for oligomerization115,117,172,265 and RNAP-recruitment.123,165,193,208,276
While this complexity hinders rational engineering of LTTRs,263 insights into the structure of both monomeric and multimeric forms facilitate the engineering and understanding of mutant regulators. For example, altering the specificity of LTTR regulators (see Figure 1C) is accelerated by knowledge on the ligand-binding pocket and the way these residues interact with the ligand of interest.277,278 More recently, LTTRs have been subjected to domain swapping to combine DNA- and ligand binding properties of interest.163 Here, structural information is vital to accurately define domains, design domain swapping chimera and to circumvent hurdles related to structural incompatibility.
Determining the protein structure of LTTRs has seen many hurdles.115,117 Due to the structural flexibility of DBDs, initial attempts at obtaining crystal structures were limited to LBDs.151,224,244 While providing valuable insights, the isolated domains lack information regarding overall protein structure and domain interactions.253 In addition to protein flexibility, oligomerization of these regulators further complicates structural studies due to the formation of aggregates during crystallization.116 Because oligomerization influences the structure of the monomers,253 higher order oligomers are needed to obtain valuable information on protein structure and quaternary interactions. However, tetramerization only occurs in the presence of DNA, and crystallization of LTTRs in interaction with DNA has proven difficult. The main problem is the high salt concentrations needed for stabilizing LTTR protein structures, which is a difficult environment for the crystallization of DNA complexes.132 As a result, only 16 full structures have been elucidated of this TF family. The available structures are given in Table 2.
Table 2. Overview of the LysR-type Transcriptional Regulators for Which the Full Structure Is Experimentally Determineda279.
| regulator | identifier | origin | ligand | reference |
|---|---|---|---|---|
| AphB | 3SZP | Vibrio cholerae | pH, O2 | (102) |
| ArgP | 3ISP | Mycobacterium tuberculosis | l-arginine, l-lysine | (268) |
| BenM | 3K1N | Acinetobacter baylyi | cis,cis-muconate, benzoate | (115) |
| CbnR | 7D98 | Cupriavidus necator | cis,cis-muconate | (132, 133) |
| CrgA | 3HHG | Neisseria meningitidis | unknown | (149) |
| DarR | 7DWO | Vibrio fischeri | d-aspartate | (280) |
| DntR | 5AE5 | Burkholderia cepacia | salicylate | (50) |
| Hink | 6M5F | Pseudomonas aeruginosa | unknown | unpublished |
| HypT | 5YDW | Salmonella enterica serovar Typhimurium | HOCl | (176) |
| LysG | 6XTU | Corynebacterium glutamicum | l-arginine, l-lysine, l-histidine | (281) |
| NdhR/CcmR | 5Y2V | Synechocystis sp. | 2-phosphoglycolate, 2-oxoglutarate | (72) |
| OxyR | 4X6G | Pseudomonas aeruginosa | H2O2 | (228) |
| TsaR | 3FXQ | Comamonas testosterone | p-toluenesulfonate | (253) |
| VV2_1132 | 5Y9S | Vibrio vulnificus | unknown | (282) |
| PA01 | 3FZV | Pseudomonas aeruginosa | unknown | unpublished |
| PA0477 | 2ESN | Pseudomonas aeruginosa | unknown | unpublished |
Information gathered from the RCSB PDB database (https://www.rcsb.org/).
The first full LTTR that has been crystallized was the cis,cis-muconate-responsive regulator CbnR from Cupriavidus necator.133 Recently, further research on the structure of CbnR resulted in tetrameric protein complexes cocrystallized with target DNA, both in the presence and absence of the inducing ligand, for the first time in this family.132 This work has confirmed some of the hypotheses surrounding LTTR regulation such as the “sliding dimer” mechanism and the occurrence of conformational changes upon ligand binding. Due to the importance of the structure for functionality, it is further discussed below, starting from the monomeric protein up to the full tetramer protein complex. As the most extensively studied LTTR in terms of protein structure, the CbnR protein will be central in this discussion.
Monomer Domain Composition and Structure
Upon translation of the regulator gene, a monomeric TF unit is produced. None of the currently discovered LTTRs regulate gene expression as monomers, but its structure has important consequences for the higher order protein complexes. The encoded LTTR protein is roughly 300 amino acids long and is composed of several domains, namely the DNA-binding domain (DBD), the ligand-binding domain (LBD) and connecting the two is a linker helix (LH) and hinge (H)66,283 as shown in Figure 4. While many of the available structural studies annotate the proteins on their different domains, the exact boundaries are not well-defined. Establishing such boundaries facilitates the development of chimeric transcription factors via engineering techniques such as domain swapping.278,284−286 Within the LTTR family, this has led to novel transcriptional regulators with custom DNA- and ligand binding specificities.163
Figure 4.

(A) Annotation of the LysR-type transcriptional regulator (LTTR) protein. Using sequence and structural alignments, an amino acid consensus sequence was created, and specific boundaries of the protein domains were set. The sequence alignment was based on 134 unique and curated sequences from the Uniprot database (https://www.uniprot.org/),287 which can be found in Supplementary Table 1. The protein models used for the structural alignment were extracted from either the RCSB database (https://www.rcsb.org/)279 or the AlphaFold database (https://alphafold.com/),288 and are given in Supplementary Figure 1. (B) The structure of regulator CbnR is given, as well as a schematic illustration of the secondary structure of LTTRs, both annotated according to the domain definitions. The α-helices are given as squares and numbered from N-terminus to C-terminus, while the β-sheets are similarly given as arrows and letters A–F. Figure adapted from ref (265). DBD = DNA-binding domain, LH = linker helix, H = hinge, LBD = ligand-binding domain, RD = subdomain of the LBD, h = hydrophobic amino acids, X = any amino acid.
To further understand the domain organization of LTTRs, a sequence alignment was performed, of which the results are shown in Figure 4A. From the Uniprot database,287 134 different LTTR sequences were extracted, which are given in Supplementary Table 1. The sequences were selected among the “Reviewed” protein entries within the ‘LysR transcriptional regulatory family’ section of the database. This results in 433 proteins, of which many are homologues with highly similar sequences. Hence, to prevent a bias toward such clusters during sequence alignment, only one sequence was chosen per regulator, resulting in a total of 134 proteins. The boundaries of the domains were subsequently found by combining the alignment information with available protein models and crystal structures, the latter given in Supplementary Figure 1. An overview of the LTTR protein structure with boundaries for the DBD, LH, H and LBD is presented in Figure 4A.
The DNA-Binding Domain
The DBD structure is visualized in Figure 4B. It is located at the N-terminus, spanning approximately 65 amino acids that fold into three α-helices and two β-strands. Within this domain, α2, α3 and both β-strands make up the LTTR-specific wHTH structure,116,253,289 which is responsible for protein–DNA interactions. While there are important residues on α2 for stabilizing the DNA–protein complex via interactions with the phosphate backbone, most of the selective residues are located in α3.133,149 The 13 base pair long TFBS spans two DNA major grooves that each hold conserved nucleotides, including the thymine and adenine of the T-N11-A sequence logo. By entering deeply into the major grooves of the DNA, the α3-helices interact with these two nucleotides, which are important for sequence recognition. Subsequently, the selective interactions enabling LTTRs to target their specific TFBSs, are made with the surrounding nucleotides.134 The winged β-strands interact with adjacent DNA minor grooves, where they contact the sugar–phosphate moiety of the DNA backbone to provide stability for protein–DNA interaction.116,133,134,149 In addition, the wing assists in positioning the DBD’s α-helices to allow proper DNA recognition and binding.116 In total, a dimer of LTTRs interacts with roughly 25 nucleotides spread over both DNA strands.134 Outside of the wHTH, α1 is made up of a conserved hydrophobic core and less conserved polar residues at the surface facing the DNA sequences. The latter are believed to aid in the interaction with target DNA, although their exact function has yet to be determined.56,133
The Ligand-Binding Domain
The 200 C-terminal residues of the regulator gene encode for the ligand recognition domain. The LBD is split into two subdomains, RD1 and RD2, connected by two short antiparallel β- strands, as shown in Figure 4B. RD1 is composed of two different parts of the polypeptide chain, as the protein loops back onto itself.151 The subdomains are structured in α/β Rossmann like folds, consisting of nine α-helices and ten β-strands.121,290 At the interface of the subdomains, between the crossover strands, a cavity is formed that holds the ligand-binding cleft.61,133,151 Due to the nature of inducing molecules varying greatly between members of the LTTR family, the architecture of the ligand-binding cleft can be vastly different.265 Strongly polar ligands form multiple hydrogen bonds with amino acids lining the cleft interior,111,137,159,239,253,261 while other ligands are engulfed in a hydrophobic cavity.200,201 LTTRs can bind their target DNA in the absence of the inducing ligand, but this molecule is required for transcriptional activation of a target promoter. Upon recognition and association with the corresponding ligand, LTTRs alter the transcription of their target gene between 6- and 200-fold. Conformational changes occur when the inducing molecule enters the RD hinge region, changing the overall protein complex structure.132
In contrast to the DBD, there is a much lower conservation noted in the LBD, both in sequence composition and in length. Therefore, determining those residues essential for structure or the key amino acids for ligand recognition has been proven difficult.111 Different mutational studies highlighted the importance of both proximal and distal residues from the ligand-binding cleft for ligand recognition. This is likely due to the subsequent signal transduction, which is equally as important as ligand recognition for transcriptional activation of LTTRs.149,210,217,253,264,291 Furthermore, the C-terminus plays an important role in oligomerization and nonspecific DNA interactions,66,71,253 and additional protein sites are important for transcriptional activation because they allow protein–protein interactions with RNAP. These sites are hidden in the uninduced protein complex and are exposed by the conformational changes following ligand interactions.292 Next to the location of the mutation, the nature of the replacing residue in the LBD shows equal importance. As an example, comparing mutations of CysB, the cysteine biosynthesis regulator from Salmonella typhimurium, at the same site revealed that T149 M results in an activation upon cysteine binding comparable to the wild type, whereas T149P resulted in only 10% of this activity. The introduction of a proline reduces overall flexibility which potentially interferes with conformational changes required for the ligand inducing.153
The Linker Helix and Hinge Region
Coinciding with the transition from DBD to LH domains, a highly conserved TxxG sequence is noted in the amino acid consensus sequence, shown in Figure 4A. The LH is a single α-helix domain that is important for correct homodimerization,65 as well as positioning the DBDs for optimal contact with the TFBS. In contrast, the boundary between the hinge and LH is more difficult to establish. The hinge is defined here as the loop between the α-helix structure of the LH, and the first β-sheet of the LBD. Upon ligand interaction in the LBD, conformational changes in the LBD cause the DBDs of one dimer in the protein complex to move to a new TFBS. This occurs around the hinge, which transduces the signal of ligand interaction toward the DBDs via the LH. In addition, the hinge is believed to hold important residues for determining the degree of oligomerization. When analyzing the octameric regulator CrgA from Neisseria meningitidis, responsible for oxidative stress response, the hinge showed a more rigid structure due to the presence of a proline residue in comparison to the more mobile structure of the tetramer CbnR.149 This difference in structural flexibility could influence the higher order oligomerization pattern.149 The full extent of the influence of LH and hinge in the overall functionality of these regulators is not yet established. Small changes in the amino acid sequence of these domains have shown to cause changes in specificity163 and altered strength of transcriptional regulation upon ligand induction,275 emphasizing the importance of gaining deeper insights into this underexplored part of the LTTR protein structure.
Understanding the structure and function of the different domains constituting LTTR monomers is essential for accurate application and further optimization of these regulators. The DBD, located at the N-terminus, is crucial for specific DNA interactions, while the LBD, at the C-terminus, recognizes and binds ligands, inducing conformational changes necessary for transcriptional activation. The LH and hinge regions facilitate dimerization and signal transduction from the LBD to the DBD. These structural elements set the stage for understanding how these monomers oligomerize into multimeric structures, which form the functional units for transcriptional regulation.
LTTR Dimers As the Functional Unit
From early on in the discovery of LTTRs, it was clear that these TFs oligomerize into higher order protein complexes due to the size, dyad symmetry, number and distance between DNA sites that are bound simultaneously.66,122 Later it was postulated that the LTTRs form dimers in the absence of DNA, and further dimerize when they bind their target DNA, resulting in what is now called the “dimer of dimers” conformation. This was confirmed with the first successful crystallization of a full LTTR protein complex, namely CbnR.132,133 Dimerization occurs around the linker helix, with antiparallel coiled coil interactions between the helices,57,71,133,221 as shown in Figure 5B. Their hydrophobic residues interact resulting in a hydrophobic patch. At the helix ends, several hydrogen bonds occur, resulting in a strong interaction that further influences the protein structure of both monomers in the dimer.133 Within this structure, further stabilization is achieved via interactions of the C-terminal regions of each monomer. Due to the antiparallel linker helix interaction, the DBDs are mirrored toward each other and show a pseudo-2-fold axis, which matches the dyad symmetry palindromic sequence of the TFBSs.253 In addition, the DBDs are separated roughly 30 Å, resembling the distance between major grooves in the DNA helix, where the binding sites are located.253
Figure 5.
Oligomerization of CbnR as a model transcription factor for the LysR-type transcriptional regulators. (A) The monomers form two distinct conformations upon dimerization, i.e., the extended (left) and compact (right) form. These structural differences are stabilized by interacting residues between the linker helix and ligand-binding domain. A close-up of this region is given above the structures. (B) Dimerization occurs primarily around the linker helix in an antiparallel manner. As a result, the DNA-binding domains (DBDs) are positioned to enter two adjacent major grooves of the DNA, as shown in the box containing a close-up of the DBDs of CbnR in association with DNA. The wings of the winged helix-turn-helix structures interact with the nascent minor grooves. (C) Two dimers interact, forming the “dimer of dimers” conformations. The tetramerization interface is mainly located in the ligand-binding domains.
While the LTTRs form homodimers, the monomers show two different conformations, namely compact and extended subunits,133,253,268 as shown in Figure 5A. The difference between these conformations is found at the hinge domain, where different angles between the linker helix and LBD are noted. For CbnR, it was found that the compact monomer showed an angle of 50°.133 A set of three direct hydrogen bonds and several water-mediated hydrogen bonds were found between the two domains. In contrast, no interactions were noted for the extended subunit, where an angle of 130° is noted. This organization of different conformations was also noted in the structures of TsaR and ArgP.253,268 The underlying mechanism for the formation of these two conformations between monomers with the same amino acid composition has not been elucidated. It is possible that in the extended form, the residues responsible for hydrogen bond formation are occupied. This would happen randomly to some monomers, while others form the hydrogen bonds and become compact subunits upon expression. Alternatively, it is possible that the determination of conformations is proximity based. During dimerization, the monomers would then be pushed into either extended or compact forms due to protein–protein interactions, causing the interacting residues to be brought into proximity for the formation of hydrogen bonds.253,268
Formation of the Tetrameric Protein Complex
As dimers, the LTTRs are able to recognize the DNA-binding sites, but simultaneous tetramerization and DNA binding is required for the creation of a stable interaction at the AS1 site. The tetramerization interface is mainly situated in the LBDs, where several interactions occur to form the eventual protein complex.132 The full tetramer structure of regulator CbnR is given in Figure 5C. The monomers collectively constitute a parallelogram configuration with the compact monomer from the dimers interacting not only with its extended monomer counterpart, but also with both units from the other dimer.71,133 In this context, the RD1 and RD2 domains of the extended and compact monomers interact through the formation of salt bridges65,71,133,253 and hydrophobic interactions.221 The C-terminal region has also shown importance for tetramerization, yet the exact mechanism has not been found.268
Tetramerization can only occur by bending the DNA to bring the dimers within proximity. However, DNA bending already starts at the dimer state. When comparing regulators, the cumulative bending angles within a dimer interacting with DNA got up to roughly 46° for CatM, and 45° for BenM. Most of this is achieved between the recognition half sites, i.e., between the thymine and adenine of the T-N11-A LTTR sequence.116 Both the bending by the individual dimers, as the full DNA bending by the tetramer are caused by the presence of extended and compact monomers, which results in an asymmetric overall structure and creates a V-shape of DBDs all situated in the same plane.116,127,133,293 Consequently, an uninduced tetramer binds the target promoter in key RNAP-binding regions and causes the DNA to bend up to 100°. The combined effect of steric hindrance and DNA bending impedes the polymerase’s ability to attach to the promoter and initiate transcription.57,265
The functional mechanism of the LTTRs is referred to as the “sliding dimer” mechanism and is visualized in Figure 2. The dimer interacting with binding site AS1 moves to a site closer to the RS-bound dimer when the LTTR complex interacts with its ligand. This became evident in DNase I footprinting pattern studies, which depict variations in DNA binding when comparing induced and uninduced states.124,142,143,268 These studies also highlight hypersensitivity of DNA to DNase I, which is linked to the DNA-bending mechanism, and they depict a shift in this hypersensitivity upon ligand addition.71,123,142 To move the DBDs of the AS1-bound dimer, conformational changes within the protein complex are required in order to form a more compact protein structure and to bring the DBD pairs closer. This has been demonstrated through multiple crystal structure studies, revealing varying quaternary structures depending upon the cocrystallization of the inducing ligand.50,132,253 When the structures of CbnR133 and TsaR253 were first compared, large differences were noted in their tetrameric structure. The former was crystallized in the absence of any inducing ligand and showed interactions between helices of nearby RD2 domains. No such interactions were found for the structure of TsaR, which was cocrystallized p-toluenesulfonate. This led to a mechanism describing the conformational changes following ligand binding including a break in such helix interactions, leading to the transition from a closed to an open form,159,253 as also illustrated in Figure 2B. This has been confirmed by several LTTRs structures, including ArgP,268 DntR,159 BenM111,118 and AphB.102 However, there are large differences in these conformational changes depending on the regulator, indicating that there is more research needed on the subsequent events following ligand interaction.
In addition to the reduced DNA bending and release of key promoter sequences, ligand-induced LTTR tetramers actively recruit RNAP to the bound promoter system. During analysis of this mechanism in CatR, it became clear that C-terminal mutations in the α-subunit of RNAP resulted in a lack of transcription activation upon CatR induction.123 These findings were complemented with research on NahR where direct interactions with the RNAP complex were found.208 The exact mechanism of RNAP recruitment by LTTRs is not yet fully understood, but it was hypothesized that because of conformational changes following ligand induction, RNAP recruitment patches migrate from internal protein positions to the tetramer surface of the LTTR complex. Here, they would be able to interact with the α-subunit of RNAP122,125,208,215,226 to guide the polymerase to the promoter’s UP region and increase the stability of RNAP-DNA interactions.71,294,295 Additionally, residues in the DBD could interact with the α-subunit of RNAP, as seen for BenM.120
While further research is necessary, current insights into the structural composition of LTTRs have led to the development of a standard regulatory model centered on the ’sliding dimer’ mechanism. The organization of monomers within one dimer, and their subsequent dimerization into the ’dimer of dimers’, supports this regulatory mechanism. Moreover, conformational changes within the tetrameric structure facilitate the sliding of the complex upon interaction with the target ligand, as well as recruitment of RNAP. However, the immense diversity of regulators within the LTTR family, along with challenges in obtaining structural information, remains a significant hurdle for further refining this ’classic’ model.
Exceptions Among the LysR-type Transcriptional Regulators
The “classic” model for LTTR regulation, based on the “sliding dimer” principle, applies to a large part of the currently characterized LTTRs. Nevertheless, since the founding of this protein family, many exceptions to these rules have emerged. What were initially considered outliers have now become increasingly common. The CysB-like regulator, Cbl, is a prime example of the extent to which LTTRs can deviate from what is considered the LTTR regulatory mechanism.61 Cbl is an important regulator in Escherichia coli where it regulates many genes related to sulfur acquisition and virulence.61,131,296,297 The promoters of these genes show only a single Cbl interaction site,297 which led to the hypothesis that it is functional as a dimer.130 The regulator activates target promoters independently, without requiring any ligand interactions. Moreover, ligand interactions have a suppressive effect on the transcriptional activation. For instance, adenosine phosphosulfate leads to the repression of Cbl-bound promoters,298 while thiosulfate impedes DNA interactions.131 There was also no autoregulation noted for this LTTR.129 As a result, converting this regulator into an LTTR-based biosensor requires significantly different consideration in comparison to the ’classic’ model.
As it is the largest bacterial transcriptional regulator family with a wide range of important biological functions, it is unsurprising that a single model is inadequate to describe all LTTRs. The introduction of a subfamily system based on functional mechanisms, oligomerization patterns, or other distinguishing features, could greatly enhance our comprehension of this family. Such a system would also accelerate the characterization of newly discovered LTTRs by more accurately linking them to known functional mechanisms. It would also allow the development of biosensor construction strategies and designs for each subfamily. Figure 6A summarizes the currently known exceptions for each of the key features of transcriptional regulation by LTTRs, together with the impact that these variations can have on the regulatory mechanism (visualized in Figure 6B). In what follows, a selection of different characteristics, showing a significant number of exceptions to the “classic” LTTR mechanism, is given as an argument for the need for a subdivision of the LTTR family.
Figure 6.
Overview of the currently identified alterations in LysR-type transcriptional regulators. (A) The components of the “classic” LysR-type transcriptional regulator model, given with “*”; their exceptions are summarized in five categories. Red arrows indicate repression, green arrows indicate activation, and gray arrows indicate no effect on transcription. (B) The possible effects of the exceptions described in A on the response curves of these regulators are visualized per parameter (as described in Figure 1B). Genetic circuit parts are given according to SBOL conventions.41,42T = Temperature, TF = transcription factor, TFBS = transcription factor binding site.
Transcription Factor Binding Site Organization and Regulation of Target Genes
The “sliding dimer” mechanism requires three TFBSs, yet an increasing number of LTTRs deviate from this, as exemplified in Figure 6A. It is expected that, due to the diversity of LTTRs, a multitude of regulatory mechanisms are embedded in the family, each with specific TFBS architectures.265 Similar to the previously mentioned Cbl, aspartate transport and metabolism regulator AalR from Acinetobacter baylyi ADP1 shows single binding sites in their target promoters.81 The occurrence of two binding sites, which were also longer in sequence than the classic T-N11-A LTTR box, was noted for regulators AlsR and CrgA. Given that these regulators form octameric structures, their regulation mechanism is likely to vary, corresponding to their mode of oligomerization.149,299
As global regulators, LTTRs are known to regulate their target genes in different ways, which is mostly achieved by the variation in TFBS sequences that result in altered affinity of the TF.151 For CysB, it was noted that next to variation in the sequences, the amount of TFBSs also differs over its target promoters.130,131,300 This is taken even further for ArgP from E. coli, which interacts with seemingly random DNA sequences. While LTTR binding sites have been found for ArgP, it was noted that it is able to regulate promoters by nonspecific binding of AT-rich regions of DNA.106 However, one of the most challenging parts of characterizing LTTRs is the discovery of their binding sites due to this natural variation in TFBS sequences. Therefore, whether these variations in the number and nature of TFBS sequences are a result of undiscovered sites or altered regulatory mechanisms, remains unclear.
Due to colocalization of LTTRs with a target gene, binding to these TFBSs also enables the TF to regulate its own expression. This is an important mechanism, as it determines the TF concentration, which greatly impacts the overall response of the regulator. However, it was established early on that a variety of autoregulatory mechanisms are present within the family,66 as shown in Figure 6A. First, negative autoregulation has been reported for LTTRs that are not found around bidirectional promoters. This has been shown for CysB, which is expressed seemingly isolated from any of its regulatory targets but is negatively autoregulated. Another example is VirR, which is the first protein expressed in the virulence operon of Rhodococcus equi. VirR represses transcription of the operon, resulting in negative autoregulation, but is able to regulate the other genes within the operon at an additional internal promoter, past the VirR sequence.190 Second, different LTTRs have been proven to lack autoregulation, examples including the previously mentioned ArgP,105 as well as PA2206 from Pseudomonas aeruginosa.232 Third, despite negative autoregulation being regarded as the standard for LTTRs, there is a seemingly large set of regulators that show positive autoregulation, including SpvR from Salmonella typhimurium(245) and YtxR from Yersinia enterocolitica.262 Several other regulators, e.g., KaeR from Lactobacillus brevis,180 LttR from Lactobacillus plantarum,187 ThnR from Sphingomonas macrogolitabida strain TFA301 or DbdR from Thauera aromatica AR-1,158 show positive autoregulation upon ligand induction. Positive autoregulation could give an advantage to systems in environments where the inducing ligand shows limited availability, due to the increased sensitivity for signal changes.302 Lastly, there is a group of transcriptional repressors which show positive autoregulation, including LrhA from E. coli,186 homologues PecT from Dickeya dadantii(234) and HexA from Photorhabdus luminescens,171 and RovM from Yersinia pseudotuberculosis.243 For these repressors, positive autoregulation can aid in repression mechanisms by ensuring the availability of adequate concentrations of the regulator.301
In accordance with the variation in autoregulation, an increasing number of regulators are found to repress their target promoter upon ligand induction. Two regulators, GltC and NdhR, have the ability to perform both activation and repression depending on the interacting ligand. In the presence of α-ketoglutarate, GltC of Bacillus subtilis performs transcriptional activation over its regulon, but in the presence of glutamate, the regulator represses target promoters.168 Similarly, NdhR is important for coordinating carbon and nitrogen metabolisms in cyanobacteria and activates appropriate genes in the presence of 2-phosphoglycolate. In times of nitrogen starvation, 2-oxoglutarate accumulates and is able to interact with NdhR, causing the regulator to repress its targets.72 However, it is not clear if this repression is different from the uninduced repression noted for all regulators. While the presence of corepressors makes it likely that these are actual repressors, it is possible that the molecules are blocking the activator from interacting and thus prevent activation, rather than cause repression. Competitive interaction has been noted for ClcR from Pseudomonas putida, where fumarate competes with 2-chloro-cis,cis-muconate for interaction at the ligand-binding cleft.142 As the difference between competitive binding and corepressor is small, further investigation into the regulatory mechanism of these regulators and potential conformational shifts upon repressing molecule binding, could provide a clear definition.
Ligand Recognition
Several regulators seem to lack an inducing ligand, and their regulation is altered based on different signals, as illustrated in Figure 6A. The LTTR NAC, or nitrogen assimilation control protein, has been shown to possess a rare functional mechanism that does not require any inducing ligands.303 This TF of Klebsiella pneumoniae is a key regulator in nitrogen metabolism, and regulates many pathways within this context, both as a repressor and activator, depending on the target promoter. It does so in often nonspecific ways, similar to ArgP, where it seemingly binds all available AT-rich DNA regions.303 Instead of ligand induction, the activity of this regulator is mainly dependent on its intracellular concentration. A two-component system that monitors nitrogen availability in the environment, regulates the transcription of NAC. This transcriptional activation is balanced out by a unique form of negative autoregulation where NAC prevents DNA bending of its own promoter region, counteracting the functional mechanism of the activator.304,305
A different set of regulators are influenced in their activity by external factors or environmental conditions. The virulence genes of Dickeya dadantii are controlled by PecT, which can only bind and repress the target genes at higher temperatures (roughly 35–38°). In these conditions, the regulator undergoes conformational changes resulting in the formation of active complexes. In addition, the relaxation of DNA supercoiling at these temperatures helps the active PecT with binding its target DNA.234 OxyR from Pseudomonas aeruginosa is a well characterized LTTR that aids in the prevention of oxidation damage by sensing hydrogen peroxide (H2O2) and regulating antioxidant genes.306 While technically H2O2 functions as a ligand for this regulator, it is not bound by the TF, but causes conformational changes by oxidizing the protein resulting in the formation of disulfide bonds.228,229 In doing so, the regulator reduces the H2O2 molecule, which counters oxidative stress in addition to the subsequent activation of antioxidant genes.230 Similarly, HypT from Salmonella enterica serovar typhimurium helps the microorganism withstand hypochlorous acid (HOCl) by regulating methionine and cysteine biosynthesis pathways as well as reducing the toxic molecule upon interaction.176 Structural changes were noted upon reduction of HOCl, which also triggered the activation of target genes, similar to the mechanism of OxyR.176 AphB from Vibrio cholerae is responsible for the activation of virulence genes under oxygen-limited situations.100,101 Due to the decrease in oxygen, a cysteine residue is reduced, causing the regulator to undergo structural changes and form an active complex, as well as increase dimerization stability.101
Other exceptions to the classic LTTR ligand recognition include binding of multiple ligands. This was discussed above for regulators GltC and NdhR, which change transcriptional regulation patterns based on the available ligand. For these TFs, the ligands both bind at the ligand-binding cleft situated between RD subdomains of the LBD. An alternative mechanism of ligand interactions based on the availability of multiple ligand-binding sites has been seen for BenM and DntR.118 This enables synergistic actions of both inducing molecules upon the regulator, resulting in unique regulatory control over the target gene by integrating multiple inputs.159,293,307 For BenM, a benzoate degradation regulator in Acinetobacter sp. Strain ADP1, both cis,cis-muconate and benzoate are bound by the regulator on different sites.118,293,307 In Burkholderia sp. strain DNT, DntR is known for its role in the oxidative degradation of 2,4-dinitrotoluene (2,4-DNT) by binding this molecule and activating the appropriate genes in response.308,309 It was further shown that the regulator recognizes salicylate at a different ligand-binding site.159 For both BenM and DntR, it is known that conformational changes within the protein structures are only fully developed with the binding of both molecules. Similarly, sulfate can interact with the ligand-binding site of CysB, causing the protein to act as a repressor. By binding in a different location, O-acetylserine excludes sulfate from the regulator due to conformational changes, transitioning CysB to an activator. Exclusion of sulfate enables N-acetylserine to bind the regulator, causing further stabilization of the activated form.154
Oligomerization
Determining the degree of oligomerization of LTTRs has been challenging because of the issues related to crystallizing higher order protein structures in the absence of DNA, as discussed above. Hence, many of the initially characterized LTTRs were classified as dimers. However, none of the current oligomerization interfaces facilitate a functional dimeric LTTR.265 It was later established that most regulators from this family form tetramers. Even within the tetrameric structures, differences in organization of monomers have been noted, which has recently been reviewed by Baugh et al.265 Higher order oligomerization creates intriguing structures of octameric protein complexes, which can further dimerize to hexadecamers in association with DNA.149 CrgA of Neisseria meningitidis is the most well-known example of octameric regulators.149 When compared to the process of tetramerization, the main differences are found in the organization of monomeric structures. While the tetrameric regulator CbnR shows extended and compact subunits, as shown in Figure 5A, all CrgA monomers show exactly the same conformation. There are no hydrogen bonds formed between the hinge and LBD, and a much more rigid hinge sequence is found due to the presence of a PxG motive instead of the more mobile GxxG motif of CbnR.149 Several other regulators have been hypothesized to form octamers, e.g., AlsR from Bacillus subtilis(90) or HsdR from Comamonas testosteroni,172 but CrgA remains the only LTTR for which this was experimentally determined.
Subclassification of the LysR-type Transcriptional Regulators
The diversity embedded within the LTTRs has become extensive, and it is likely that more exceptions will arise during continuous expansion of this TF family. As illustrated in Figure 6, these variations significantly impact the regulator’s effect on the target promoter, resulting in distinct response curves when used as biosensors. While this variability can be advantageous for developing biosensors for a wide range of genetic circuit applications, it also complicates their design due to the difficulty of predicting their effects in vivo. To accelerate and simplify LTTR-based biosensor development, it would be beneficial to create a subclassification system that groups LTTRs with similar variations on the ’classic’ regulatory mechanism.
In the past, several attempts have been made to organize this diversity in subfamilies, yet this has been hampered by the large number of regulator features that vary. These attempts were mainly focused on the classification of LTTRs based on sequence similarities. However, homology seemingly lacks many of the intricate differences that are embedded in the LTTR amino acid sequence. In the hinge region, the presence of a single proline residue is potentially enough to prevent the formation of extended and compact monomer structures, resulting in octameric regulators. As an alternative to homology-based clustering, some suggestions can be made from the discussion presented here. Oligomerization is seemingly linked to many of the other features which are varied in the exceptions presented above. It is directly related to the overall structure of the protein complex and the DNA-interaction, making it important for the precise method of transcriptional regulation. The main problem is determining the degree of oligomerization, especially as higher order complexes are usually only formed in the presence of target DNA sequences. Several techniques have been used in the past, including SDS-PAGE, size-exclusion chromatograph, microscopy-based techniques and X-ray crystallography or NMR spectroscopy, yet the requirement of DNA–protein complexes, the difficulties with extracting LTTRs, and the low throughput of these methods, inhibit the use of oligomerization as an efficient grouping parameter. Alternatively, the architecture of TFBSs has important implications on the mechanism behind transcriptional regulation. Similar to oligomerization, accurate and reliable TFBS prediction is difficult due to the extent of variation in these sequences.
To effectively group LTTRs into subfamilies, comprehensive information on a wide range of parameters is essential. By integrating this data with amino acid sequence analysis, it is possible to identify patterns that can serve as the basis for establishing a more precise classification system. This process can be significantly expedited through the application of machine learning techniques, which excel at intricate pattern analysis when provided with sufficient data. Therefore, by establishing high-throughput experimental methods for characterizing LTTRs and integrating computational biology, significant progress can be achieved in understanding and organizing the LTTR family.
Conclusions
Despite advances in modifying natural transcriptional regulators for biosensor use, there is a need for a broader repertoire of sensors as genetic circuit inputs in order to fully exploit the technology. To expand on the currently limited set of used transcriptional regulators, the untapped pool of TFs from the LTTR family can be exploited. LTTRs are the largest group of bacterial transcriptional regulators and play pivotal roles in various biological processes. They have evolved into highly efficient dual-regulatory molecules that provide precise control over their regulon via the “sliding-dimers” mechanism. The tetrameric protein complexes bind their target promoters, repressing transcription by interacting with key RNAP interaction sites, and making them further inaccessible by bending the DNA. After interaction with their respective ligand, the LTTRs undergo conformational changes, causing the protein complex to bind an upstream binding site resulting in RNAP recruitment via direct and indirect mechanisms.
However, with the size of the family comes a highly diverse set of regulators that show many exceptions to this “classic” LTTR model. Due to the increasing quantity of exceptions, it is becoming clear that one model is not sufficient to describe a TF family of this size. Reorganization of the family, potentially into subfamilies of regulators with similar mechanisms of regulation, will aid in the understanding both known and newly discovered LTTRs. Despite the added complexity, the array of different mechanisms opens options to new applications and novel biosensors. Expanding our understanding of their structure and regulatory mechanisms is essential for harnessing their capabilities and expanding the toolbox of sensors in synthetic biology.
Acknowledgments
W.D. was supported by a Ph.D. grant (1SC6820N) from the Research Foundation Flanders (FWO). B.D.P. holds an FWO postdoctoral fellowship (1246323N). The authors thank the UGent Core Facility “HTS for SynBio” for training, support, and access to the instrument park.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acssynbio.4c00219.
Protein structures used to validate the LTTR domain definition and list of proteins used to establish the domain definition (PDF)
Author Contributions
All authors were involved in the conceptualization and design of the manuscript. W.D. drafted the manuscript, which was followed by critical revision by all authors.
The authors declare no competing financial interest.
Supplementary Material
References
- Gardner T. S.; Cantor C. R.; Collins J. J. Construction of a genetic toggle switch in Escherichia coli. Nature 2000, 403, 339–342. 10.1038/35002131. [DOI] [PubMed] [Google Scholar]
- Elowitz M. B.; Leibler S. A synthetic oscillatory networkof transcriptional regulators. Nature 2000, 403, 335–338. 10.1038/35002125. [DOI] [PubMed] [Google Scholar]
- Cameron D. E.; Bashor C. J.; Collins J. J. A brief history of synthetic biology. Nature Reviews Microbiology 2014, 12, 381–390. 10.1038/nrmicro3239. [DOI] [PubMed] [Google Scholar]
- Zhang H.; Jiang T. Synthetic circuits, devices and modules. Protein and Cell 2010, 1, 974–978. 10.1007/s13238-010-0133-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Engstrom M. D.; Pfleger B. F. Transcription control engineering and applications in synthetic biology. Synthetic Syst. Biotechnol. 2017, 2, 176–191. 10.1016/j.synbio.2017.09.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Smanski M. J.; Zhou H.; Claesen J.; Shen B.; Fischbach M. A.; Voigt C. A. Synthetic biology to access and expand nature’s chemical diversity. Nature Reviews Microbiology 2016, 14, 135–149. 10.1038/nrmicro.2015.24. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim S. G.; Noh M. H.; Lim H. G.; Jang S.; Jang S.; Koffas M. A.; Jung G. Y. Molecular parts and genetic circuits for metabolic engineering of microorganisms. FEMS Microbiol. Lett. 2018, 365, fny187. 10.1093/femsle/fny187. [DOI] [PubMed] [Google Scholar]
- Nielsen A. A.; Segall-Shapiro T. H.; Voigt C. A. Advances in genetic circuit design: Novel biochemistries, deep part mining, and precision gene expression. Curr. Opin. Chem. Biol. 2013, 17, 878–892. 10.1016/j.cbpa.2013.10.003. [DOI] [PubMed] [Google Scholar]
- Ong N. T.; Tabor J. J. A Miniaturized Escherichia coli Green Light Sensor with High Dynamic Range. ChemBioChem. 2018, 19, 1255–1258. 10.1002/cbic.201800007. [DOI] [PubMed] [Google Scholar]
- Daeffler K. N.; Galley J. D.; Sheth R. U.; Ortiz-Velez L. C.; Bibb C. O.; Shroyer N. F.; Britton R. A.; Tabor J. J. Engineering bacterial thiosulfate and tetrathionate sensors for detecting gut inflammation. Mol. Syst. Biol. 2017, 13, 923. 10.15252/msb.20167416. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhao J.; Sun H.; Wang G.; Wang Q.; Wang Y.; Li Q.; Bi S.; Qi Q.; Wang Q. Engineering Chimeric Chemoreceptors and Two-Component Systems for Orthogonal and Leakless Biosensing of Extracellular γ-Aminobutyric Acid. J. Agric. Food Chem. 2024, 72, 14216–14228. 10.1021/acs.jafc.4c00041. [DOI] [PubMed] [Google Scholar]
- Juárez J. F.; Lecube-Azpeitia B.; Brown S. L.; Johnston C. D.; Church G. M. Biosensor libraries harness large classes of binding domains for construction of allosteric transcriptional regulators. Nat. Commun. 2018, 9, 3101. 10.1038/s41467-018-05525-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stanton B. C.; Nielsen A. A.; Tamsir A.; Clancy K.; Peterson T.; Voigt C. A. Genomic mining of prokaryotic repressors for orthogonal logic gates. Nat. Chem. Biol. 2014, 10, 99–105. 10.1038/nchembio.1411. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang F.; Keasling J. Biosensors and their applications in microbial metabolic engineering. Trends in Microbiology 2011, 19, 323–329. 10.1016/j.tim.2011.05.003. [DOI] [PubMed] [Google Scholar]
- He W.; Yuan S.; Zhong W. H.; Siddikee M. A.; Dai C. C. Application of genetically engineered microbial whole-cell biosensors for combined chemosensing. Appl. Microbiol. Biotechnol. 2016, 100, 1109–1119. 10.1007/s00253-015-7160-6. [DOI] [PubMed] [Google Scholar]
- Ulrich L. E.; Koonin E. V.; Zhulin I. B. One-component systems dominate signal transduction in prokaryotes. Trends in Microbiology 2005, 13, 52–56. 10.1016/j.tim.2004.12.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tellechea-Luzardo J.; Stiebritz M. T.; Carbonell P. Transcription factor-based biosensors for screening and dynamic regulation. Front. Bioeng. Biotechnol. 2023, 11, 1118702. 10.3389/fbioe.2023.1118702. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kaczmarek J. A.; Prather K. L. Effective use of biosensors for high-Throughput library screening for metabolite production. J. Ind. Microbiol. Biotechnol. 2021, 48, kuab049. 10.1093/jimb/kuab049. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang Y.; Cortez J. D.; Hammer S. K.; Carrasco-López C.; García Echauri S.; Wiggins J. B.; Wang W.; Avalos J. L. Biosensor for branched-chain amino acid metabolism in yeast and applications in isobutanol and isopentanol production. Nat. Commun. 2022, 13, 270. 10.1038/s41467-021-27852-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- De Paepe B.; Peters G.; Coussement P.; Maertens J.; De Mey M. Tailor-made transcriptional biosensors for optimizing microbial cell factories. Journal of Industrial Microbiology and Biotechnology 2017, 44, 623–645. 10.1007/s10295-016-1862-3. [DOI] [PubMed] [Google Scholar]
- Dixon T. A.; Williams T. C.; Pretorius I. S. Sensing the future of bio-informational engineering. Nat. Commun. 2021, 12, 388. 10.1038/s41467-020-20764-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gui Q.; Lawson T.; Shan S.; Yan L.; Liu Y. The application of whole cell-based biosensors for use in environmental analysis and in medical diagnostics. Sensors 2017, 17, 1623. 10.3390/s17071623. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moraskie M.; Roshid M. H. O.; O’Connor G.; Dikici E.; Zingg J. M.; Deo S.; Daunert S. Microbial whole-cell biosensors: Current applications, challenges, and future perspectives. Biosens. Bioelectron. 2021, 191, 113359. 10.1016/j.bios.2021.113359. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Özyurt C.; Üstükarcı H.; Evran S.; Telefoncu A. MerR-fluorescent protein chimera biosensor for fast and sensitive detection of Hg2+ in drinking water. Biotechnology and Applied Biochemistry 2019, 66, 731–737. 10.1002/bab.1805. [DOI] [PubMed] [Google Scholar]
- Liu X.; Silverman A. D.; Alam K. K.; Iverson E.; Lucks J. B.; Jewett M. C.; Raman S. Design of a transcriptional biosensor for the portable, on-demand detection of cyanuric acid. ACS Synth. Biol. 2020, 9, 84–94. 10.1021/acssynbio.9b00348. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chou H. H.; Keasling J. D. Programming adaptive control to evolve increased metabolite production. Nat. Commun. 2013, 4, 2595. 10.1038/ncomms3595. [DOI] [PubMed] [Google Scholar]
- Hirasawa T.; Maeda T. Adaptive laboratory evolution of microorganisms: methodology and application for bioproduction. Microorganisms 2023, 11, 92. 10.3390/microorganisms11010092. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang X.; Sun Z.; Bian J.; Gao Y.; Zhang D.; Xu G.; Zhang X.; Li H.; Shi J.; Xu Z. Rational metabolic engineering combined with biosensor-mediated adaptive laboratory evolution for L-cysteine overproduction from glycerol in Escherichia coli. Fermentation 2022, 8, 299. 10.3390/fermentation8070299. [DOI] [Google Scholar]
- Shen Y. P.; Pan Y.; Niu F. X.; Liao Y. L.; Huang M.; Liu J. Z. Biosensor-assisted evolution for high-level production of 4-hydroxyphenylacetic acid in Escherichia coli. Metabolic Engineering 2022, 70, 1–11. 10.1016/j.ymben.2021.12.008. [DOI] [PubMed] [Google Scholar]
- Mahr R.; Gätgens C.; Gätgens J.; Polen T.; Kalinowski J.; Frunzke J. Biosensor-driven adaptive laboratory evolution of l-valine production in Corynebacterium glutamicum. Metabolic Engineering 2015, 32, 184–194. 10.1016/j.ymben.2015.09.017. [DOI] [PubMed] [Google Scholar]
- Wu Y.; Jameel A.; Xing X. H.; Zhang C. Advanced strategies and tools to facilitate and streamline microbial adaptive laboratory evolution. Trends Biotechnol. 2022, 40, 38–59. 10.1016/j.tibtech.2021.04.002. [DOI] [PubMed] [Google Scholar]
- Orsi E.; Claassens N. J.; Nikel P. I.; Lindner S. N. Growth-coupled selection of synthetic modules to accelerate cell factory development. Nat. Commun. 2021, 12, 5295. 10.1038/s41467-021-25665-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang Y.; Lin Y.; Wang J.; Wu Y.; Zhang R.; Cheng M.; Shen X.; Wang J.; Chen Z.; Li C.; Yuan Q.; Yan Y. Sensor-regulator and RNAi based bifunctional dynamic control network for engineered microbial synthesis. Nat. Commun. 2018, 9, 3043. 10.1038/s41467-018-05466-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu D.; Zhang F. Metabolic feedback circuits provide rapid control of metabolite dynamics. ACS Synth. Biol. 2018, 7, 347–356. 10.1021/acssynbio.7b00342. [DOI] [PubMed] [Google Scholar]
- Hartline C. J.; Schmitz A. C.; Han Y.; Zhang F. Dynamic control in metabolic engineering: Theories, tools, and applications. Metabolic Engineering 2021, 63, 126–140. 10.1016/j.ymben.2020.08.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cress B. F.; Trantas E. A.; Ververidis F.; Linhardt R. J.; Koffas M. A. Sensitive cells: Enabling tools for static and dynamic control of microbial metabolic pathways. Curr. Opin. Biotechnol. 2015, 36, 205–214. 10.1016/j.copbio.2015.09.007. [DOI] [PubMed] [Google Scholar]
- Tan S.; Shi F.; Liu H.; Yu X.; Wei S.; Fan Z.; Li Y. Dynamic Control of 4-Hydroxyisoleucine Biosynthesis by Modified l-Isoleucine Biosensor in Recombinant Corynebacterium glutamicum. ACS Synth. Biol. 2020, 9, 2378–2389. 10.1021/acssynbio.0c00127. [DOI] [PubMed] [Google Scholar]
- Xu X.; Li X.; Liu Y.; Zhu Y.; Li J.; Du G.; Chen J.; Ledesma-Amaro R.; Liu L. Pyruvate-responsive genetic circuits for dynamic control of central metabolism. Nat. Chem. Biol. 2020, 16, 1261–1268. 10.1038/s41589-020-0637-3. [DOI] [PubMed] [Google Scholar]
- Zhang F.; Carothers J. M.; Keasling J. D. Design of a dynamic sensor-regulator system for production of chemicals and fuels derived from fatty acids. Nat. Biotechnol. 2012, 30, 354–359. 10.1038/nbt.2149. [DOI] [PubMed] [Google Scholar]
- Zhou S.; Yuan S. F.; Nair P. H.; Alper H. S.; Deng Y.; Zhou J. Development of a growth coupled and multi-layered dynamic regulation network balancing malonyl-CoA node to enhance (2S)-naringenin biosynthesis in Escherichia coli. Metabolic Engineering 2021, 67, 41–52. 10.1016/j.ymben.2021.05.007. [DOI] [PubMed] [Google Scholar]
- Beal J.; et al. Communicating structure and function in Synthetic Biology diagrams. ACS Synth. Biol. 2019, 8, 1818–1825. 10.1021/acssynbio.9b00139. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Quinn J. Y.; Cox R. S.; Adler A.; Beal J.; Bhatia S.; Cai Y.; Chen J.; Clancy K.; Galdzicki M.; Hillson N. J.; Le Novere N.; Maheshwari A. J.; McLaughlin J. A.; Myers C. J.; P U.; Pocock M.; Rodriguez C.; Soldatova L.; Stan G.-B. V.; Swainston N.; Wipat A.; Sauro H. M.; et al. SBOL visual: A graphical language for genetic designs. PLoS Biol. 2015, 13, e1002310. 10.1371/journal.pbio.1002310. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brophy J. A.; Voigt C. A. Principles of genetic circuit design. Nat. Methods 2014, 11, 508–520. 10.1038/nmeth.2926. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rogers J. K.; Taylor N. D.; Church G. M. Biosensor-based engineering of biosynthetic pathways. Curr. Opin. Biotechnol. 2016, 42, 84–91. 10.1016/j.copbio.2016.03.005. [DOI] [PubMed] [Google Scholar]
- Bintu L.; Buchler N. E.; Garcia H. G.; Gerland U.; Hwa T.; Kondev J.; Kuhlman T.; Phillips R. Transcriptional regulation by the numbers: Applications. Current Opinion in Genetics and Development 2005, 15, 125–135. 10.1016/j.gde.2005.02.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tabor J. J.; Groban E. S.; Voigt C. A. Systems Biology and Biotechnology of Escherichia coli 2009, 401–439. 10.1007/978-1-4020-9394-4_19. [DOI] [Google Scholar]
- Dietrich J. a.; McKee A. E.; Keasling J. D. High-throughput metabolic engineering: advances in small-molecule screening and selection. Annual review of biochemistry 2010, 79, 563–90. 10.1146/annurev-biochem-062608-095938. [DOI] [PubMed] [Google Scholar]
- Kiel C.; Yus E.; Serrano L. Engineering signal transduction pathways. Cell 2010, 140, 33–47. 10.1016/j.cell.2009.12.028. [DOI] [PubMed] [Google Scholar]
- Koch M.; Pandi A.; Delépine B.; Faulon J. L. A dataset of small molecules triggering transcriptional and translational cellular responses. Data in Brief 2018, 17, 1374–1378. 10.1016/j.dib.2018.02.061. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lerche M.; Dian C.; Round A.; Lönneborg R.; Brzezinski P.; Leonard G. A. The solution configurations of inactive and activated DntR have implications for the sliding dimer mechanism of LysR transcription factors. Sci. Rep. 2016, 6, 19988. 10.1038/srep19988. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Perez-Rueda E.; Hernandez-Guerrero R.; Martinez-Nuñez M. A.; Armenta-Medina D.; Sanchez I.; Ibarra J. A. Abundance, diversity and domain architecture variability in prokaryotic DNA-binding transcription factors. PLoS One 2018, 13, e0195332. 10.1371/journal.pone.0195332. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pareja E.; Pareja-Tobes P.; Manrique M.; Pareja-Tobes E.; Bonal J.; Tobes R. ExtraTrain: A database of extragenic regions and transcriptional information in prokaryotic organisms. BMC Microbiol. 2006, 6, 29. 10.1186/1471-2180-6-29. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Henikoff S.; Haughn G. W.; Calvo J. M.; Wallace J. C. A large family of bacterial activator proteins. Proc. Natl. Acad. Sci. U.S.A. 1988, 85, 6602–6606. 10.1073/pnas.85.18.6602. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Henikoff S.; Wallace J. C.; Brown J. P. Finding protein similarities with nucleotide sequence databases. Methods Enzymol. 1990, 183, 111–132. 10.1016/0076-6879(90)83009-X. [DOI] [PubMed] [Google Scholar]
- Stragier P.; Patte J. C. Regulation of diaminopimelate decarboxylase synthesis in Escherichia coli. III. nucleotide sequence and regulation of the lysR gene. J. Mol. Biol. 1983, 168, 333–350. 10.1016/S0022-2836(83)80022-9. [DOI] [PubMed] [Google Scholar]
- Issa K. H. B.; Phan G.; Broutin I. Functional mechanism of the efflux pumps transcription regulators from Pseudomonas aeruginosa based on 3D structures. Front. Mol. Biosci. 2018, 5, 57. 10.3389/fmolb.2018.00057. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maddocks S. E.; Oyston P. C. F. Structure and function of the LysR-type transcriptional regulator (LTTR) family proteins. Microbiology 2008, 154, 3609–23. 10.1099/mic.0.2008/022772-0. [DOI] [PubMed] [Google Scholar]
- Dik D. A.; Fisher J. F.; Mobashery S. Cell-Wall recycling of the Gram-negative bacteria and the nexus to antibiotic resistance. Chem. Rev. 2018, 118, 5952–5984. 10.1021/acs.chemrev.8b00277. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lahiri A.; Das P.; Chakravortty D. Salmonella Typhimurium: Insight into the multi-faceted role of the LysR-type transcriptional regulators in Salmonella. Int. J. Biochem. Cell Biol. 2009, 41, 2129–2133. 10.1016/j.biocel.2009.05.006. [DOI] [PubMed] [Google Scholar]
- Momany C.; Neidle E. L. Defying stereotypes: The elusive search for a universal model of LysR-type regulation. Mol. Microbiol. 2012, 83, 453–456. 10.1111/j.1365-2958.2011.07960.x. [DOI] [PubMed] [Google Scholar]
- Stec E.; Witkowska-Zimny M.; Hryniewicz M. M.; Neumann P.; Wilkinson A. J.; Brzozowski A. M.; Verma C. S.; Zaim J.; Wysocki S.; Bujacz G. D. Structural basis of the sulphate starvation response in E. coli: crystal structure and mutational analysis of the cofactor-binding domain of the Cbl transcriptional regulator. J. Mol. Biol. 2006, 364, 309–322. 10.1016/j.jmb.2006.06.033. [DOI] [PubMed] [Google Scholar]
- Pérez-Rueda E.; Collado-Vides J. Common history at the origin of the position - Function correlation in transcriptional regulators in archaea and bacteria. Journal of Molecular Evolution 2001, 53, 172–179. 10.1007/s002390010207. [DOI] [PubMed] [Google Scholar]
- Sun J.; Klein A. A LysR-type regulator is involved in the negative regulation of genes encoding selenium-free hydrogenases in the archaeon Methanococcus voltae. Mol. Microbiol. 2004, 52, 563–571. 10.1111/j.1365-2958.2004.03998.x. [DOI] [PubMed] [Google Scholar]
- Oliveira Monteiro L. M.; Saraiva J. P.; Brizola Toscan R.; Stadler P. F.; Silva-Rocha R.; Nunes da Rocha U. PredicTF: prediction of bacterial transcription factors in complex microbial communities using deep learning. Environ. Microbiomes 2022, 17, 7. 10.1186/s40793-021-00394-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Knapp G. S.; Hu J. C. Specificity of the E. coli LysR-type transcriptional regulators. PLoS One 2010, 5, e15189. 10.1371/journal.pone.0015189. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schell M. A. Molecular biology of the LysR family of transcriptional regulators. Annual Reviews Microbiolgy 1993, 47, 597–626. 10.1146/annurev.mi.47.100193.003121. [DOI] [PubMed] [Google Scholar]
- Reen F. J.; Barret M.; Fargier E.; O’Muinneacháin M.; O’Gara F. Molecular evolution of LysR-type transcriptional regulation in Pseudomonas aeruginosa. Mol. Phylogenet. Evol. 2013, 66, 1041–1049. 10.1016/j.ympev.2012.12.014. [DOI] [PubMed] [Google Scholar]
- Mao D.; Bushin L. B.; Moon K.; Wu Y.; Seyedsayamdost M. R. Discovery of scmR as a global regulator of secondary metabolism and virulence in Burkholderia thailandensis E264. Proc. Natl. Acad. Sci. U.S.A. 2017, 114, 2920–2928. 10.1073/pnas.1619529114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Le Guillouzer S.; Groleau M. C.; Mauffrey F.; Déziel E. ScmR, a global regulator of gene expression, quorum sensing, pH homeostasis, and virulence in burkholderia thailandensis. J. Bacteriol. 2020, 10.1128/JB.00776-19. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fragel S. M.; Montada A.; Heermann R.; Baumann U.; Schacherl M.; Schnetz K. Characterization of the pleiotropic LysR-type transcription regulator LeuO of Escherichia coli. Nucleic Acids Res. 2019, 47, 7363–7379. 10.1093/nar/gkz506. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tropel D.; van der Meer J. R. Bacterial transcriptional regulators for degradation pathways of aromatic compounds. Microbiology and Molecular Biology Reviews 2004, 68, 474–500. 10.1128/MMBR.68.3.474-500.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jiang Y. L.; Wang X. P.; Sun H.; Han S. J.; Li W. F.; Cui N.; Lin G. M.; Zhang J. Y.; Cheng W.; Cao D. D.; Zhang Z. Y.; Zhang C. C.; Chen Y.; Zhou C. Z. Coordinating carbon and nitrogen metabolic signaling through the cyanobacterial global repressor NdhR. Proc. Natl. Acad. Sci. U.S.A. 2018, 115, 403–408. 10.1073/pnas.1716062115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lindberg F.; Westman L.; Normark S. Regulatory components in Citrobacter freundii ampC β-lactamase induction. Proc. Natl. Acad. Sci. U.S.A. 1985, 82, 4620–4624. 10.1073/pnas.82.14.4620. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Van den Bergh E. R.; Dijkhuizen L.; Meijer W. G. CbbR, a LysR-type transcriptional activator, is required for expression of the autotrophic CO2 fixation enzymes of Xanthobacter flavus. J. Bacteriol. 1993, 175, 6097–6104. 10.1128/jb.175.19.6097-6104.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Saint-Girons I.; Parsot C.; Zakin M. M.; Barzu O.; Cohen G. N.; Weissbach H. Methionine biosynthesis in enterobacteriaceae: Biochemical, regulatory, and evolutionary aspect. Crit. Rev. Biochem. Mol. Biol. 1988, 23, S1–S42. 10.3109/10409238809083374. [DOI] [PubMed] [Google Scholar]
- Steiert J. G.; Rolfes R. J.; Zalkin H.; Stauffer G. V. Regulation of the Escherichia coli glyA gene by the purR gene product. J. Bacteriol. 1990, 172, 3799–3803. 10.1128/jb.172.7.3799-3803.1990. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Goldberg M. B.; Boyko S. A.; Calderwood S. B. Transcriptional regulation by iron of a Vibrio cholerae virulence gene and homology of the gene to the Escherichia coli Fur system. J. Bacteriol. 1990, 172, 6863–6870. 10.1128/jb.172.12.6863-6870.1990. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Appelbaum E. R.; Thompson D. V.; Idler K.; Chartrain N. Rhizobium japonicum USDA 191 has two nodD genes that differ in primary structure and function. Journal of bacteriology 1988, 170, 12–20. 10.1128/jb.170.1.12-20.1988. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lu Z.; Takeuchi M.; Sato T. The LysR-type transcriptional regulator YofA controls cell division through the regulation of expression of ftsW in Bacillus subtilis. J. Bacteriol. 2007, 189, 5642–5651. 10.1128/JB.00467-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Reen F. J.; Romano S.; Dobson A. D.; O’Gara F. The sound of silence: Activating silent biosynthetic gene clusters in marine microorganisms. Marine Drugs 2015, 13, 4754–4783. 10.3390/md13084754. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bedore S. R.; Schmidt A. L.; Slarks L. E.; Duscent-Maitland C. V.; Elliott K. T.; Andresen S.; Costa F. G.; Weerth R. S.; Tumen-Velasquez M. P.; Nilsen L. N.; Dean C. E.; Karls A. C.; Hoover T. R.; Neidle E. L. Regulation of L- and D-aspartate transport and metabolism in Acinetobacter baylyi ADP1. Appl. Environ. Microbiol. 2022, 88, e00883-22. 10.1128/aem.00883-22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bolla J. R.; Howes A. C.; Fiorentino F.; Robinson C. V. Assembly and regulation of the chlorhexidine-specific efflux pump AceI. Proc. Natl. Acad. Sci. U.S.A. 2020, 117, 17011–17018. 10.1073/pnas.2003271117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu Q.; Hassan K. A.; Ashwood H. E.; Gamage H. K.; Li L.; Mabbutt B. C.; Paulsen I. T. Regulation of the aceI multidrug efflux pump gene in Acinetobacter baumannii. J. Antimicrob. Chemother. 2018, 73, 1492–1500. 10.1093/jac/dky034. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gavira J. A.; Rico-Jiménez M.; Ortega A.; Petukhova N. V.; Bug D. S.; Castellví A.; Porozov Y. B.; Zhulin I. B.; Krell T.; Matilla M. A. Emergence of an auxin sensing domain in plant-associated bacteria. mBio 2023, 14, e03363-22. 10.1128/mbio.03363-22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Matilla M. A.; Nogellova V.; Morel B.; Krell T.; Salmond G. P. Biosynthesis of the acetyl-coA carboxylase-inhibiting antibiotic, andrimid in Serratia is regulated by hfq and the LysR-type transcriptional regulator, admX. Environmental Microbiology 2016, 18, 3635–3650. 10.1111/1462-2920.13241. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Matilla M. A.; Daddaoua A.; Chini A.; Morel B.; Krell T. An auxin controls bacterial antibiotics production. Nucleic Acids Res. 2018, 46, 11229–11238. 10.1093/nar/gky766. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rintoul M. R.; Cusa E.; Baldomà L.; Badia J.; Reitzer L.; Aguilar J. Regulation of the Escherichia coli allantoin regulon: Coordinated function of the repressor AllR and the activator AllS. J. Mol. Biol. 2002, 324, 599–610. 10.1016/S0022-2836(02)01134-8. [DOI] [PubMed] [Google Scholar]
- Xue J.; Tan B.; Yang S.; Luo M.; Xia H.; Zhang X.; Zhou X.; Yang X.; Yang R.; Li Y.; Qiu J. Influence of cAMP receptor protein (CRP) on bacterial virulence and transcriptional regulation of allS by CRP in Klebsiella pneumoniae. Gene 2016, 593, 28–33. 10.1016/j.gene.2016.08.006. [DOI] [PubMed] [Google Scholar]
- Hasegawa A.; Ogasawara H.; Kori A.; Teramoto J.; Ishihama A. The transcription regulator AIIR senses both allantoin and glyoxylate and controls a set of genes for degradation and reuilization of purines. Microbiology 2008, 154, 3366–3378. 10.1099/mic.0.2008/020016-0. [DOI] [PubMed] [Google Scholar]
- Frädrich C.; March A.; Fiege K.; Hartmann A.; Jahn D.; Härtig E. The transcription factor AlsR binds and regulates the promoter of the alssd operon responsible for acetoin formation in bacillus subtilis. J. Bacteriol. 2012, 194, 1100–1112. 10.1128/JB.06425-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Renna M. C.; Najimudin N.; Winik L. R.; Zahler S. A. Regulation of the Bacillus subtilis alsS, alsD, and alsR genes involved in post-exponential-phase production of acetoin. J. Bacteriol. 1993, 175, 3863–3875. 10.1128/jb.175.12.3863-3875.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- De Oliveira R. R.; Nicholson W. L. The LysR-type transcriptional regulator (LTTR) AlsR indirectly regulates expression of the Bacillus subtilis bdhA gene encoding 2,3-butanediol dehydrogenase. Appl. Microbiol. Biotechnol. 2013, 97, 7307–7316. 10.1007/s00253-013-4871-4. [DOI] [PubMed] [Google Scholar]
- Bishop R. E.; Weiner J. H. Overproduction, solubilization, purification and DNA-binding properties of AmpR from Citrobacter freundii. Eur. J. Biochem. 1993, 213, 405–412. 10.1111/j.1432-1033.1993.tb17775.x. [DOI] [PubMed] [Google Scholar]
- Lindquist S.; Lindberg F.; Normark S. Binding of the Citrobacter freundii ampR regulator to a single DNA site provides both autoregulation and activation of the inducible ampC β-lactamase gene. J. Bacteriol. 1989, 171, 3746–3753. 10.1128/jb.171.7.3746-3753.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu C.; Li C.; Chen Y.; Hao H.; Liang J.; Duan R.; Guo Z.; Zhang J.; Zhao Z.; Jing H.; Wang X.; Shao S. Role of low-molecular-mass penicillin-binding proteins, NagZ and AmpR in AmpC β-lactamase regulation of Yersinia enterocolitica. Front. Cell. Infect. Microbiol. 2017, 7, 425. 10.3389/fcimb.2017.00425. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bartowsky E.; Normark S. Purification and mutant analysis of Citrobacter freundii AmpR, the regulator for chromosomal AmpC β-lactamase. Mol. Microbiol. 1991, 5, 1715–1725. 10.1111/j.1365-2958.1991.tb01920.x. [DOI] [PubMed] [Google Scholar]
- Balcewich M. D.; Reeve T. M.; Orlikow E. A.; Donald L. J.; Vocadlo D. J.; Mark B. L. Crystal structure of the AmpR effector binding domain provides insight into the molecular regulation of inducible AmpC β-lactamase. J. Mol. Biol. 2010, 400, 998–1010. 10.1016/j.jmb.2010.05.040. [DOI] [PubMed] [Google Scholar]
- Vadlamani G.; Thomas M. D.; Patel T. R.; Donald L. J.; Reeve T. M.; Stetefeld J.; Standing K. G.; Vocadlo D. J.; Mark B. L. The β-lactamase gene regulator AmpR is a tetramer that recognizes and binds the D-Ala-D-Ala motif of its repressor UDP-N-acetylmuramic acid (MurNAc)-pentapeptide. J. Biol. Chem. 2015, 290, 2630–2643. 10.1074/jbc.M114.618199. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kovacikova G.; Lin W.; Skorupski K. The LysR-type virulence activator AphB regulates the expression of genes in Vibrio cholerae in response to low pH and anaerobiosis. J. Bacteriol. 2010, 192, 4181–4191. 10.1128/JB.00193-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen J.; Byun H.; She Q.; Liu Z.; Ruggeberg K. G.; Pu Q.; Jung I. J.; Zhu D.; Brockett M. R.; Hsiao A.; Zhu J. S-Nitrosylation of the virulence regulator AphB promotes Vibrio cholerae pathogenesis. PLoS Pathogens 2022, 18, e1010581. 10.1371/journal.ppat.1010581. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu Z.; Yang M.; Peterfreund G. L.; Tsou A. M.; Selamoglu N.; Daldal F.; Zhong Z.; Kan B.; Zhu J. Vibrio cholerae anaerobic induction of virulence gene expression is controlled by thiol-based switches of virulence regulator AphB. Proc. Natl. Acad. Sci. U.S.A. 2011, 108, 810–815. 10.1073/pnas.1014640108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Taylor J. L.; de Silva R. S.; Kovacikova G.; Lin W.; Taylor R. K.; Skorupski K.; Kull F. J. The crystal structure of AphB, a virulence gene activator from Vibrio cholerae, reveals residues that influence its response to oxygen and pH. Mol. Microbiol. 2012, 83, 457–470. 10.1111/j.1365-2958.2011.07919.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kovacikova G.; Skorupski K. A Vibrio cholerae LysR homolog, AphB, cooperates with AphA at the tcpPH promoter to activate expression of the ToxR virulence cascade. J. Bacteriol. 1999, 181, 4250–4256. 10.1128/JB.181.14.4250-4256.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arai H.; Ohishi T.; Chang M. Y.; Kudo T. Arrangement and regulation of the genes for meta-pathway enzymes required for degradation of phenol in Comamonas testosteroni TA441. Microbiology 2000, 146, 1707–1715. 10.1099/00221287-146-7-1707. [DOI] [PubMed] [Google Scholar]
- Nandineni M. R.; Gowrishankar J. Evidence for an arginine exporter encoded by yggA (argO) that is regulated by the LysR-type transcriptional regulator argP in Escherichia coli. J. Bacteriol. 2004, 186, 3539–3546. 10.1128/JB.186.11.3539-3546.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nguyen Le Minh P.; Velázquez Ruiz C.; Vandermeeren S.; Abwoyo P.; Bervoets I.; Charlier D. Differential protein-DNA contacts for activation and repression by ArgP, a LysR-type (LTTR) transcriptional regulator in Escherichia coli. Microbiological Research 2018, 206, 141–158. 10.1016/j.micres.2017.10.009. [DOI] [PubMed] [Google Scholar]
- Celis R. T. Repression and activation of arginine transport genes in Escherichia coli K 12 by the ArgP protein. J. Mol. Biol. 1999, 294, 1087–1095. 10.1006/jmbi.1999.3308. [DOI] [PubMed] [Google Scholar]
- Porrúa O.; García-Jaramillo M.; Santero E.; Govantes F. The LysR-type regulator AtzR binding site: DNA sequences involved in activation, repression and cyanuric acid-dependent repositioning. Mol. Microbiol. 2007, 66, 410–427. 10.1111/j.1365-2958.2007.05927.x. [DOI] [PubMed] [Google Scholar]
- Platero A. I.; López-Sánchez A.; Tomás-Gallardo L.; Santero E.; Govantes F. Mechanism of antiactivation at the pseudomonas sp. strain ADP σN-dependent PatzT promoter. Appl. Environ. Microbiol. 2016, 82, 4350–4362. 10.1128/AEM.00906-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yao X.; He W.; Lu C. D. Functional characterization of seven γ-glutamylpolyamine synthetase genes and the bauRABCD locus for polyamine and β-alanine utilization in Pseudomonas aeruginosa PAO1. J. Bacteriol. 2011, 193, 3923–3930. 10.1128/JB.05105-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ezezika O. C.; Haddad S.; Clark T. J.; Neidle E. L.; Momany C. Distinct effector-binding sites enable synergistic transcriptional activation by BenM, a LysR-type regulator. J. Mol. Biol. 2007, 367, 616–629. 10.1016/j.jmb.2006.09.090. [DOI] [PubMed] [Google Scholar]
- Clark T.; Haddad S.; Neidle E.; Momany C. Crystallization of the effector-binding domains of BenM and CatM, LysR-type transcriptional regulators from Acinetobacter sp. ADP1. Acta Crystallographica Section D: Biological Crystallography 2004, 60, 105–108. 10.1107/S0907444903021589. [DOI] [PubMed] [Google Scholar]
- Clark T. J.; Phillips R. S.; Bundy B. M.; Momany C.; Neidle E. L. Benzoate decreases the binding of cis,cis-muconate to the BenM regulator despite the synergistic effect of both compounds on transcriptional activation. J. Bacteriol. 2004, 186, 1200–1204. 10.1128/JB.186.4.1200-1204.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Collier L. S.; Gaines G. L.; Neidle E. L. Regulation of benzoate degradation in Acinetobacter sp. strain ADP1 by BenM, a LysR-type transcriptional activator. J. Bacteriol. 1998, 180, 2493–2501. 10.1128/JB.180.9.2493-2501.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ruangprasert A.; Craven S. H.; Neidle E. L.; Momany C. Full-length structures of BenM and two variants reveal different oligomerization schemes for LysR-type transcriptional regulators. J. Mol. Biol. 2010, 404, 568–586. 10.1016/j.jmb.2010.09.053. [DOI] [PubMed] [Google Scholar]
- Alanazi A. M.; Neidle E. L.; Momany C. The DNA-binding domain of BenM reveals the structural basis for the recognition of a T-N11-A sequence motif by LysR-type transcriptional regulators. Acta Crystallographica Section D: Biological Crystallography 2013, 69, 1995–2007. 10.1107/S0907444913017320. [DOI] [PubMed] [Google Scholar]
- Ezezika O. C.; Haddad S.; Neidle E. L.; Momany C. Oligomerization of BenM, a LysR-type transcriptional regulator: Structural basis for the aggregation of proteins in this family. Acta Crystallographica Section F: Structural Biology and Crystallization Communications 2007, 63, 361–368. 10.1107/S1744309107019185. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Craven S. H.; Ezezika O. C.; Haddad S.; Hall R. A.; Momany C.; Neidle E. L. Inducer responses of BenM, a LysR-type transcriptional regulator from Acinetobacter baylyi ADP1. Mol. Microbiol. 2009, 72, 881–894. 10.1111/j.1365-2958.2009.06686.x. [DOI] [PubMed] [Google Scholar]
- Magdalena J.; Gérard C.; Joris B.; Forsman M.; Dusart J. The two β-lactamase genes of Streptomyces cacaoi, blaL and blaU, are under the control of the same regulatory system. Molecular and General Genetics 1997, 255, 187–193. 10.1007/s004380050488. [DOI] [PubMed] [Google Scholar]
- Tumen-Velasquez M. P.; Laniohan N. S.; Momany C.; Neidle E. L. Engineering CatM, a LysR-type transcriptional regulator, to respond synergistically to two effectors. Genes 2019, 10, 421. 10.3390/genes10060421. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Neidle E. L.; Hartnett C.; Ornston L. N. Characterization of Acinetobacter calcoaceticus catM, a repressor gene homologous in sequence to transcriptional activator genes. J. Bacteriol. 1989, 171, 5410–5421. 10.1128/jb.171.10.5410-5421.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chugani S. A.; Parsek M. R.; Hershberger C. D.; Murakami K.; Ishihama A.; Chakrabarty A. M. Activation of the catBCA Promoter: Probing the Interaction of CatR and RNA Polymerase through In Vitro Transcription. J. Bacteriol. 1997, 179, 2221–2227. 10.1128/jb.179.7.2221-2227.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McFall S. M.; Klem T. J.; Fujita N.; Ishihama A.; Chakrabarty A. M. DNase I footprinting, DNA bending and in vitro transcription analyses of ClcR and CatR interactions with the clcABD promoter: Evidence of a conserved transcriptional activation mechanism. Mol. Microbiol. 1997, 24, 965–976. 10.1046/j.1365-2958.1997.4041763.x. [DOI] [PubMed] [Google Scholar]
- Parsek M. R.; Ye R. W.; Pun P.; Chakrabarty A. M. Critical nucleotides in the interaction of a LysR-type regulator with its target promoter region. catBC promoter activation by CatR. J. Biol. Chem. 1994, 269, 11279–11284. 10.1016/S0021-9258(19)78122-8. [DOI] [PubMed] [Google Scholar]
- Parsek M. R.; Shinabarger D. L.; Rothmel R. K.; Chakrabarty A. M. Roles of CatR and cis,cis-muconate in activation of the catBC operon, which is involved in benzoate degradation in Pseudomonas putida. J. Bacteriol. 1992, 174, 7798–7806. 10.1128/jb.174.23.7798-7806.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rothmel R. K.; Shinabarger D. L.; Parsek M. R.; Aldrich T. L.; Chakrabarty A. M. Functional analysis of the Pseudomonas putida regulatory protein CatR: Transcriptional studies and determination of the CatR DNA-binding site by hydroxyl-radical footprinting. J. Bacteriol. 1991, 173, 4717–4724. 10.1128/jb.173.15.4717-4724.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Van Keulen G.; Ridder A. N.; Dijkhuizen L.; Meijer W. G. Analysis of DNA binding and transcriptional activation by the LysR-type transcriptional regulator CbbR of Xanthobacter flavus. J. Bacteriol. 2003, 185, 1245–1252. 10.1128/JB.185.4.1245-1252.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Van Keulen G.; Girbal L.; Van Den Bergh E. R.; Dijkhuizen L.; Meijer W. G. The LysR-type transcriptional regulator CbbR controlling autotrophic CO2 fixation by Xanthobacter flavus is an NADPH sensor. J. Bacteriol. 1998, 180, 1411–1417. 10.1128/JB.180.6.1411-1417.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Iwanicka-Nowicka R.; Hryniewicz M. M. A new gene, cbl, encoding a member of the LysR family of transcriptional regulators belongs to Escherichia coli cys regulon. Gene 1995, 166, 11–17. 10.1016/0378-1119(95)00606-8. [DOI] [PubMed] [Google Scholar]
- Van Der Ploeg J. R.; Iwanicka-Nowicka R.; Kertesz M. A.; Leisinger T.; Hryniewicz M. M. Involvement of CysB and Cbl regulatory proteins in expression of the tauABCD operon and other sulfate starvation-inducible genes in Escherichia coli. J. Bacteriol. 1997, 179, 7671–7678. 10.1128/jb.179.24.7671-7678.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Van Der Ploeg J. R.; Iwanicka-Nowicka R.; Bykowski T.; Hryniewicz M. M.; Leisinger T. The Escherichia coli ssuEADCB gene cluster is required for the utilization of sulfur from aliphatic sulfonates and is regulated by the transcriptional activator Cbl. J. Biol. Chem. 1999, 274, 29358–29365. 10.1074/jbc.274.41.29358. [DOI] [PubMed] [Google Scholar]
- Giannopoulou E. A.; Senda M.; Koentjoro M. P.; Adachi N.; Ogawa N.; Senda T. Crystal structure of the full-length LysR-type transcription regulator CbnR in complex with promoter DNA. FEBS Journal 2021, 288, 4560–4575. 10.1111/febs.15764. [DOI] [PubMed] [Google Scholar]
- Muraoka S.; Okumura R.; Ogawa N.; Nonaka T.; Miyashita K.; Senda T. Crystal Structure of a full-length LysR-type transcriptional regulator, CbnR: Unusual combination of two subunit forms and molecular bases for causing and changing DNA bend. J. Mol. Biol. 2003, 328, 555–566. 10.1016/S0022-2836(03)00312-7. [DOI] [PubMed] [Google Scholar]
- Koentjoro M. P.; Adachi N.; Senda M.; Ogawa N.; Senda T. Crystal structure of the DNA-binding domain of the LysR-type transcriptional regulator CbnR in complex with a DNA fragment of the recognition-binding site in the promoter region. FEBS Journal 2018, 285, 977–989. 10.1111/febs.14380. [DOI] [PubMed] [Google Scholar]
- Jourlin-Castelli C.; Mani N.; Nakano M. M.; Sonenshein A. L. CcpC, a novel regulator of the LysR family required for glucose repression of the citB gene in Bacillus subtilis. J. Mol. Biol. 2000, 295, 865–878. 10.1006/jmbi.1999.3420. [DOI] [PubMed] [Google Scholar]
- Mittal M.; Pechter K. B.; Picossi S.; Kim H. J.; Kerstein K. O.; Sonenshein A. L. Dual role of CcpC protein in regulation of aconitase gene expression in Listeria monocytogenes and Bacillus subtilis. Microbiology 2013, 159, 68–76. 10.1099/mic.0.063388-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen J.; Shang F.; Wang L.; Zou L.; Bu T.; Jin L.; Dong Y.; Ha N. C.; Quan C.; Nam K. H.; Xu Y. Structural and biochemical analysis of the citrate-responsive mechanism of the regulatory domain of catabolite control protein E from Staphylococcus aureus. Biochemistry 2018, 57, 6054–6060. 10.1021/acs.biochem.8b00671. [DOI] [PubMed] [Google Scholar]
- Windhöver U.; Bowien B. Identification of cfxR, an activator gene of autotrophic CO2 fixation in Alcaligenes eutrophus. Mol. Microbiol. 1991, 5, 2695–2705. 10.1111/j.1365-2958.1991.tb01978.x. [DOI] [PubMed] [Google Scholar]
- Suzuki K.; Uchiyama T.; Suzuki M.; Nikaidou N.; Regue M.; Watanabe T. LysR-type transcriptional regulator ChiR is essential for production of all chitinases and a chitin-binding protein, CBP21, in Serratia marcescens 2170. Biosci., Biotechnol., Biochem. 2001, 65, 338–347. 10.1271/bbb.65.338. [DOI] [PubMed] [Google Scholar]
- Yang S. J.; Rice K. C.; Brown R. J.; Patton T. G.; Liou L. E.; Park Y. H.; Bayles K. W. A LysR-type regulator, CidR, is required for induction of the Staphylococcus aureus cidABC operon. J. Bacteriol. 2005, 187, 5893–5900. 10.1128/JB.187.17.5893-5900.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jin S.; Sonenshein A. L. Transcriptional regulation of Bacillus subtilis citrate synthase genes. J. Bacteriol. 1994, 176, 4680–4690. 10.1128/jb.176.15.4680-4690.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mcfall S. M.; Parsek M. R.; Chakrabarty A. M. 2-Chloromuconate and Clcr-mediated activation of the clcABD operon: In vitro transcriptional and DNase I footprint analyses. J. Bacteriol. 1997, 179, 3655–3663. 10.1128/jb.179.11.3655-3663.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Coco W. M.; Parsek M. R.; Chakrabarty A. M. Purification of the LysR family regulator, ClcR, and its interaction with the Pseudomonas putida clcABD chlorocatechol operon promoter. J. Bacteriol. 1994, 176, 5530–5533. 10.1128/jb.176.17.5530-5533.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mahounga D. M.; Sun H.; Jiang Y. L. Crystal structure of the effector-binding domain of Synechococcus elongatus CmpR in complex with ribulose 1,5-bisphosphate. Acta Crystallographica Section F: Structural Biology Communications 2018, 74, 506–511. 10.1107/S2053230X18008841. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Takahashi Y.; Yamaguchi O.; Omata T. Roles of CmpR, a LysR family transcriptional regulator, in acclimation of the cyanobacterium Synechococcus sp. strain PCC 7942 to low-CO2 and high-light conditions. Mol. Microbiol. 2004, 52, 837–845. 10.1111/j.1365-2958.2004.04021.x. [DOI] [PubMed] [Google Scholar]
- Nishimura T.; Takahashi Y.; Yamaguchi O.; Suzuki H.; Maeda S. I.; Omata T. Mechanism of low CO2-induced activation of the cmp bicarbonate transporter operon by a LysR family protein in the cyanobacterium Synechococcus elongatus strain PCC 7942. Mol. Microbiol. 2008, 68, 98–109. 10.1111/j.1365-2958.2008.06137.x. [DOI] [PubMed] [Google Scholar]
- Deghmane A. E.; Giorgini D.; Maigre L.; Taha M. K. Analysis in vitro and in vivo of the transcriptional regulator CrgA of Neisseria meningitidis upon contact with target cells. Mol. Microbiol. 2004, 53, 917–927. 10.1111/j.1365-2958.2004.04167.x. [DOI] [PubMed] [Google Scholar]
- Fan X.; Zhao Z.; Sun T.; Rou W.; Gui C.; Saleem T.; Zhao X.; Xu X.; Zhuo T.; Hu X.; Zou H. The LysR-type transcriptional regulator CrgA negatively regulates the flagellar master regulator flhDC in ralstonia solanacearum GMI1000. J. Bacteriol. 2020, 10.1128/JB.00419-20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sainsbury S.; Lane L. A.; Ren J.; Gilbert R. J.; Saunders N. J.; Robinson C. V.; Stuart D. I.; Owens R. J. The structure of CrgA from Neisseria meningitidis reveals a new octameric assembly state for LysR transcriptional regulators. Nucleic Acids Res. 2009, 37, 4545–4558. 10.1093/nar/gkp445. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sung Y. C.; Fuchs J. A. The Escherichia coli K-12 cyn operon is positively regulated by a member of the lysR family. J. Bacteriol. 1992, 174, 3645–3650. 10.1128/jb.174.11.3645-3650.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tyrrell R.; Verschueren K. H.; Dodson E. J.; Murshudov G. N.; Addy C.; Wilkinson A. J. The structure of the cofactor-binding fragment of the LysR family member, CysB: A familiar fold with a surprising subunit arrangement. Structure 1997, 5, 1017–1032. 10.1016/S0969-2126(97)00254-2. [DOI] [PubMed] [Google Scholar]
- Verschueren K. H.; Addy C.; Dodson E. J.; Wilkinson A. J. Crystallization of full-length CysB of Klebsiella aerogenes, a LysR-type transcriptional regulator. Acta Crystallographica Section D: Biological Crystallography 2001, 57, 260–262. 10.1107/S0907444900016851. [DOI] [PubMed] [Google Scholar]
- Colyer T. E.; Kredich N. M. In vitro characterization of constitutive CysB proteins from Salmonella typhimurium. Mol. Microbiol. 1996, 21, 247–256. 10.1046/j.1365-2958.1996.6301347.x. [DOI] [PubMed] [Google Scholar]
- Mittal M.; Singh A. K.; Kumaran S. Structural and biochemical characterization of ligand recognition by CysB, the master regulator of sulfate metabolism. Biochimie 2017, 142, 112–124. 10.1016/j.biochi.2017.08.011. [DOI] [PubMed] [Google Scholar]
- Kredich N. M. Regulation Salmonella of L-Cysteine typhimurium Biosynthesis in. Biological chemistry 1971, 246, 3474–3484. 10.1016/S0021-9258(18)62154-4. [DOI] [PubMed] [Google Scholar]
- Guillouard I.; Auger S.; Hullo M. F.; Chetouani F.; Danchin A.; Martin-Verstraete I. Identification of Bacillus subtilis CysL, a regulator of the cysJI operon, which encodes sulfite reductase. J. Bacteriol. 2002, 184, 4681–4689. 10.1128/JB.184.17.4681-4689.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jones R. M.; Popham D. L.; Schmidt A. L.; Neidle E. L.; Stabb E. V. Vibrio fischeri DarR directs responses to D-aspartate and represents a group of similar LysR-type transcriptional regulators. J. Bacteriol. 2018, 10.1128/JB.00773-17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pacheco-Sanchez D.; Molina-Fuentes A.; Marín P.; Díaz-Romero A.; Marqués S. DbdR, a new member of the LysR family of transcriptional regulators, coordinately controls four promoters in the Thauera aromatica AR-1 3,5-Dihydroxybenzoate anaerobic degradation pathway. Appl. Environ. Microbiol. 2019, 85, e02295-18. 10.1128/AEM.02295-18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Devesse L.; Smirnova I.; Lönneborg R.; Kapp U.; Brzezinski P.; Leonard G. A.; Dian C. Crystal structures of DntR inducer binding domains in complex with salicylate offer insights into the activation of LysR-type transcriptional regulators. Mol. Microbiol. 2011, 81, 354–367. 10.1111/j.1365-2958.2011.07673.x. [DOI] [PubMed] [Google Scholar]
- Smirnova I. A.; Dian C.; Leonard G. A.; McSweeney S.; Birse D.; Brzezinski P. Development of a bacterial biosensor for nitrotoluenes: The crystal structure of the transcriptional regulator DntR. J. Mol. Biol. 2004, 340, 405–418. 10.1016/j.jmb.2004.04.071. [DOI] [PubMed] [Google Scholar]
- de las Heras A.; Chavarría M.; de Lorenzo V. Association of dnt genes of Burkholderia sp. DNT with the substrate-blind regulator DntR draws the evolutionary itinerary of 2,4-dinitrotoluene biodegradation. Mol. Microbiol. 2011, 82, 287–299. 10.1111/j.1365-2958.2011.07825.x. [DOI] [PubMed] [Google Scholar]
- De Paepe B.; Maertens J.; Vanholme B.; De Mey M. Modularization and response curve engineering of a naringenin-responsive transcriptional biosensor. ACS Synth. Biol. 2018, 7, 1303–1314. 10.1021/acssynbio.7b00419. [DOI] [PubMed] [Google Scholar]
- De Paepe B.; Maertens J.; Vanholme B.; De Mey M. Chimeric LysR-Type transcriptional biosensors for customizing ligand specificity profiles toward flavonoids. ACS Synth. Biol. 2019, 8, 318–331. 10.1021/acssynbio.8b00326. [DOI] [PubMed] [Google Scholar]
- Lee Y.; Peña-Llopis S.; Kang Y. S.; Shin H. D.; Demple B.; Madsen E. L.; Jeon C. O.; Park W. Expression analysis of the fpr (ferredoxin-NADP+ reductase) gene in Pseudomonas putida KT2440. Biochem. Biophys. Res. Commun. 2006, 339, 1246–1254. 10.1016/j.bbrc.2005.11.135. [DOI] [PubMed] [Google Scholar]
- Stauffer L. T.; Stauffer G. V. GcvA interacts with both the α and σ subunits of RNA polymerase to activate the Escherichia coli gcvB gene and the gcvTHP operon. FEMS Microbiology Letters 2005, 242, 333–338. 10.1016/j.femsle.2004.11.027. [DOI] [PubMed] [Google Scholar]
- Shi J.; Feng Z.; Song Q.; Wang F.; Zhang Z.; Liu J.; Li F.; Wen A.; Liu T.; Ye Z.; Zhang C.; Das K.; Wang S.; Feng Y.; Lin W. Structural and functional insights into transcription activation of the essential LysR-type transcriptional regulators. Protein Sci. 2024, 33, e5012. 10.1002/pro.5012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gebhardt M. J.; Czyz D. M.; Singh S.; Zurawski D. V.; Becker L.; Shuman H. A. GigC, a LysR family transcription regulator, is required for cysteine metabolism and virulence in acinetobacter baumannii. Infect. Immun. 2020, 10.1128/IAI.00180-20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Picossi S.; Belitsky B. R.; Sonenshein A. L. Molecular mechanism of the regulation of Bacillus subtilis gltAB expression by GltC. J. Mol. Biol. 2007, 365, 1298–1313. 10.1016/j.jmb.2006.10.100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Belitsky B. R.; Janssen P. J.; Sonenshein A. L. Sites required for GltC-dependent regulation of Bacillus subtilis glutamate synthase expression. J. Bacteriol. 1995, 177, 5686–5695. 10.1128/jb.177.19.5686-5695.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bohannon D. E.; Sonenshein A. L. Positive regulation of glutamate biosynthesis in Bacillus subtilis. Journal of bacteriology 1989, 171, 4718–4727. 10.1128/jb.171.9.4718-4727.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Langer A.; Moldovan A.; Harmath C.; Joyce S. A.; Clarke D. J.; Heermann R. HexA is a versatile regulator involved in the control of phenotypic heterogeneity of Photorhabdus luminescens. PLoS One 2017, 12, e0176535. 10.1371/journal.pone.0176535. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gong W.; Xiong G.; Maser E. Oligomerization and negative autoregulation of the LysR-type transcriptional regulator HsdR from Comamonas testosteroni. Journal of Steroid Biochemistry and Molecular Biology 2012, 132, 203–211. 10.1016/j.jsbmb.2012.05.012. [DOI] [PubMed] [Google Scholar]
- Gong W.; Xiong G.; Maser E. Identification and characterization of the LysR-type transcriptional regulator HsdR for steroid-inducible expression of the 3α-Hydroxysteroid dehydrogenase/carbonyl reductase gene in Comamonas testosteroni. Appl. Environ. Microbiol. 2012, 78, 941–950. 10.1128/AEM.06872-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Litwin C. M.; Quackenbush J. Characterization of a Vibrio vulnificus LysR homologue, HupR, which regulates expression of the haem uptake outer membrane protein, HupA. Microbial Pathogenesis 2001, 31, 295–307. 10.1006/mpat.2001.0472. [DOI] [PubMed] [Google Scholar]
- Buggy J. J.; Sganga M. W.; Bauer C. E. Nucleotide sequence and characterization of the Rhodobacter capsulatus hvrB gene: HvrB is an activator of S-adenosyl-L-homocysteine hydrolase expression and is a member of the LysR family. J. Bacteriol. 1994, 176, 61–69. 10.1128/jb.176.1.61-69.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jo I.; Kim D.; No T.; Hong S.; Ahn J.; Ryu S.; Ha N. C. Structural basis for HOCl recognition and regulation mechanisms of HypT, a hypochlorite-specific transcriptional regulator. Proc. Natl. Acad. Sci. U.S.A. 2019, 116, 3740–3745. 10.1073/pnas.1811509116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Malakooti J.; Ely B. Identification and characterization of the ilvR gene encoding a LysR-type regulator of Caulobacter crescentus. J. Bacteriol. 1994, 176, 1275–1281. 10.1128/jb.176.5.1275-1281.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wek R. C.; Hatfield G. W. Transcriptional activation at adjacent operators in the divergent-overlapping ilvY and ilvC promoters of Escherichia coli. J. Mol. Biol. 1988, 203, 643–663. 10.1016/0022-2836(88)90199-4. [DOI] [PubMed] [Google Scholar]
- Goldberg M. B.; Boyko S. A.; Calderwood S. B. Positive transcriptional regulation of an iron-regulated virulence gene in Vibrio cholerae. Proc. Natl. Acad. Sci. U.S.A. 1991, 88, 1125–1129. 10.1073/pnas.88.4.1125. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pande S. G.; Pagliai F. A.; Gardner C. L.; Wrench A.; Narvel R.; Gonzalez C. F.; Lorca G. L. Lactobacillus brevis responds to flavonoids through KaeR, a LysR-type of transcriptional regulator. Mol. Microbiol. 2011, 81, 1623–1639. 10.1111/j.1365-2958.2011.07796.x. [DOI] [PubMed] [Google Scholar]
- Sánchez-Popoca D.; Serrano-Fujarte I.; Fernández-Mora M.; Calva E. The LeuO regulator and quiescence: About transcriptional roadblocks, multiple promoters, and crispr-cas. Mol. Microbiol. 2022, 118, 503–509. 10.1111/mmi.14990. [DOI] [PubMed] [Google Scholar]
- Hernández-Lucas I.; Calva E. The coming of age of the LeuO regulator. Mol. Microbiol. 2012, 85, 1026–1028. 10.1111/j.1365-2958.2012.08175.x. [DOI] [PubMed] [Google Scholar]
- Hernández-Lucas I.; Gallego-Hernández A. L.; Encarnación S.; Fernández-Mora M.; Martínez-Batallar A. G.; Salgado H.; Oropeza R.; Calva E. The LysR-type transcriptional regulator LeuO controls expression of several genes in Salmonella enterica serovar typhi. J. Bacteriol. 2008, 190, 1658–1670. 10.1128/JB.01649-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tabata M.; Endo R.; Ito M.; Ohtsubo Y.; Kumar A.; Tsuda M.; Nagata Y. The lin genes for γ-hexachlorocyclohexane degradation in Sphingomonas sp. MM-1 proved to be dispersed across multiple plasmids. Biosci., Biotechnol., Biochem. 2011, 75, 466–472. 10.1271/bbb.100652. [DOI] [PubMed] [Google Scholar]
- Endo R.; Kamakura M.; Miyauchi K.; Fukuda M.; Ohtsubo Y.; Tsuda M.; Nagata Y. Identification and characterization of genes involved in the downstream degradation pathway of γ-hexachlorocyclohexane in Sphingomonas paucimobilis UT26. J. Bacteriol. 2005, 187, 847–853. 10.1128/JB.187.3.847-853.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lehnen D.; Blumer C.; Polen T.; Wackwitz B.; Wendisch V. F.; Unden G. LrhA as a new transcriptional key regulator of flagella, motility and chemotaxis genes in Escherichia coli. Mol. Microbiol. 2002, 45, 521–532. 10.1046/j.1365-2958.2002.03032.x. [DOI] [PubMed] [Google Scholar]
- Liu X. X.; Xiong Z. Q.; Wang G. Q.; Wang L. F.; Xia Y. J.; Song X.; Ai L. Z. LysR Family Regulator LttR Controls Production of Conjugated Linoleic Acid in Lactobacillus plantarum by Directly Activating the cla Operon. Appl. Environ. Microbiol. 2021, 87, e02798-20. 10.1128/AEM.02798-20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Binder S.; Schendzielorz G.; Stäbler N.; Krumbach K.; Hoffmann K.; Bott M.; Eggeling L. A high-throughput approach to identify genomic variants of bacterial metabolite producers at the single-cell level. Genome Biol. 2012, 13, R40. 10.1186/gb-2012-13-5-r40. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bellmann A.; Vrljić M.; Pátek M.; Sahm H.; Krämer R.; Eggeling L. Expression control and specificity of the basic amino acid exporter lysE of Corynebacterium glutamicum. Microbiology 2001, 147, 1765–1774. 10.1099/00221287-147-7-1765. [DOI] [PubMed] [Google Scholar]
- Byrne G. A.; Russell D. A.; Chen X.; Meijer W. G. Transcriptional regulation of the virR operon of the intracellular pathogen Rhodococcus equi. J. Bacteriol. 2007, 189, 5082–5089. 10.1128/JB.00431-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Peng H. L.; Shiou S. R.; Chang H. Y. Characterization of mdcR, a regulatory gene of the malonate catabolic system in Klebsiella pneumoniae. J. Bacteriol. 1999, 181, 2302–2306. 10.1128/JB.181.7.2302-2306.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cai X. Y.; Redfield B.; Maxon M.; Weissbach H.; Brot N. The effect of homocysteine on metR regulation of metE, metR and metH expression in vitro. Biochem. Biophys. Res. Commun. 1989, 163, 79–83. 10.1016/0006-291X(89)92101-3. [DOI] [PubMed] [Google Scholar]
- Fritsch P. S.; Urbanowski M. L.; Stauffer G. V. Role of the RNA polymerase α subunits in MetR-dependent activation of metE and metH: Important residues in the C-terminal domain and orientation requirements within RNA polymerase. J. Bacteriol. 2000, 182, 5539–5550. 10.1128/JB.182.19.5539-5550.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maxon M. E.; Redfield B.; Cai X. Y.; Shoeman R.; Fujita K.; Fisher W.; Stauffer G.; Weissbach H.; Brot N. Regulation of methionine synthesis in Escherichia coli: Effect of the MetR protein on the expression of the metE and metR genes. Proc. Natl. Acad. Sci. U.S.A. 1989, 86, 85–89. 10.1073/pnas.86.1.85. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu W. F.; Urbanowski M. L.; Stauffer G. V. Characterization of a second MetR-binding site in the metE metR regulatory region of Salmonella typhimurium. J. Bacteriol. 1995, 177, 1834–1839. 10.1128/jb.177.7.1834-1839.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Punekar A. S.; Porter J.; Carr S. B.; Phillips S. E. Structural basis for DNA recognition by the transcription regulator MetR. Acta Crystallographica Section:F Structural Biology Communications 2016, 72, 417–426. 10.1107/S2053230X16006828. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maxon M. E.; Wigboldus J.; Brot N.; Weissbach H. Structure-function studies on Escherichia coli MetR protein, a putative prokaryotic leucine zipper protein. Proc. Natl. Acad. Sci. U.S.A. 1990, 87, 7076–7079. 10.1073/pnas.87.18.7076. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kovaleva G. Y.; Gelfand M. S. Transcriptional regulation of the methionine and cysteine transport and metabolism in streptococci. FEMS Microbiology Letters 2007, 276, 207–215. 10.1111/j.1574-6968.2007.00934.x. [DOI] [PubMed] [Google Scholar]
- Lemme A.; Sztajer H.; Wagner-Döbler I. Characterization of mleR, a positive regulator of malolactic fermentation and part of the acid tolerance response in Streptococcus mutans. BMC Microbiol. 2010, 10, 58. 10.1186/1471-2180-10-58. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ilangovan A.; Fletcher M.; Rampioni G.; Pustelny C.; Rumbaugh K.; Heeb S.; Cámara M.; Truman A.; Chhabra S. R.; Emsley J.; Williams P. Structural basis for native agonist and synthetic inhibitor recognition by the Pseudomonas aeruginosa quorum sensing regulator PqsR (MvfR). PLoS Pathogens 2013, 9, e1003508. 10.1371/journal.ppat.1003508. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kitao T.; Lepine F.; Babloudi S.; Walte F.; Steinbacher S.; Maskos K.; Blaesse M.; Negri M.; Pucci M.; Zahler B.; Felici A.; Rahme L. G. Molecular insights into function and competitive inhibition of pseudomonas aeruginosa multiple virulence factor regulator. mBio 2018, 10.1128/mBio.02158-17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xu N.; Yu S.; Moniot S.; Weyand M.; Blankenfeldt W. Crystallization and preliminary crystal structure analysis of the ligand-binding domain of PqsR (MvfR), the Pseudomonas quinolone signal (PQS) responsive quorum-sensing transcription factor of Pseudomonas aeruginosa. Acta Crystallographica Section F: Structural Biology and Crystallization Communications 2012, 68, 1034–1039. 10.1107/S1744309112032538. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xiao G.; Déziel E.; He J.; Lépine F.; Lesic B.; Castonguay M. H.; Milot S.; Tampakaki A. P.; Stachel S. E.; Rahme L. G. MvfR, a key Pseudomonas aeruginosa pathogenicity LTTR-class regulatory protein, has dual ligands. Mol. Microbiol. 2006, 62, 1689–1699. 10.1111/j.1365-2958.2006.05462.x. [DOI] [PubMed] [Google Scholar]
- McGrath S.; Wade D. S.; Pesci E. C. Dueling quorum sensing systems in Pseudomonas aeruginosa control the production of the Pseudomonas quinolone signal (PQS). FEMS Microbiology Letters 2004, 230, 27–34. 10.1016/S0378-1097(03)00849-8. [DOI] [PubMed] [Google Scholar]
- Cao H.; Krishnan G.; Goumnerov B.; Tsongalis J.; Tompkins R.; Rahme L. G. A quorum sensing-associated virulence gene of Pseudomonas aeruginosa encodes a LysR-like transcription regulator with a unique self-regulatory mechanism. Proc. Natl. Acad. Sci. U.S.A. 2001, 98, 14613–14618. 10.1073/pnas.251465298. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jones R. M.; Britt-Compton B.; Williams P. A. The naphthalene catabolic (nag) genes of Ralstonia sp. strain U2 are an operon that is regulated by NagR, a LysR-type transcriptional regulator. J. Bacteriol. 2003, 185, 5847–5853. 10.1128/JB.185.19.5847-5853.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhou N. Y.; Fuenmayor S. L.; Williams P. A. nag genes of Ralstonia (formerly pseudomonas) sp. Strain U2 encoding enzymes for gentisate catabolism. J. Bacteriol. 2001, 183, 700–708. 10.1128/JB.183.2.700-708.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Park W.; Jeon C. O.; Madsen E. L. Interaction of NahR, a LysR-type transcriptional regulator, with the α subunit of RNA polymerase in the naphthalene degrading bacterium, Pseudomonas putida NCIB 9816–4. FEMS Microbiol. Lett. 2002, 213, 159–165. 10.1111/j.1574-6968.2002.tb11300.x. [DOI] [PubMed] [Google Scholar]
- Shin H. J. Development of highly-sensitive microbial biosensors by mutation of the nahR regulatory gene. J. Biotechnol. 2010, 150, 246–250. 10.1016/j.jbiotec.2010.09.936. [DOI] [PubMed] [Google Scholar]
- Cebolla A.; Sousa C.; De Lorenzo V. Effector specificity mutants of the transcriptional activator NahR of naphthalene degrading Pseudomonas define protein sites involved in binding of aromatic inducers. J. Biol. Chem. 1997, 272, 3986–3992. 10.1074/jbc.272.7.3986. [DOI] [PubMed] [Google Scholar]
- Figge R. M.; Cassier-Chauvat C.; Chauvat F.; Cerff R. Characterization and analysis of an NAD(P)H dehydrogenase transcriptional regulator critical for the survival of cyanobacteria facing inorganic carbon starvation and osmotic stress. Mol. Microbiol. 2001, 39, 455–469. 10.1046/j.1365-2958.2001.02239.x. [DOI] [PubMed] [Google Scholar]
- Dover N.; Padan E. Transcription of nhaA, the main Na+/H+ antiporter of Escherichia coli, is regulated by Na+ and growth phase. J. Bacteriol. 2001, 183, 644–653. 10.1128/JB.183.2.644-653.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Von Lintig J.; Kreusch D.; Schroder J. Opine-regulated promoters and LysR-type regulators in the nopaline (noc) and octopine (occ) catabolic regions of Ti plasmids of Agrobacterium tumefaciens. J. Bacteriol. 1994, 176, 495–503. 10.1128/jb.176.2.495-503.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schlaman H. R.; Okker R. J.; Lugtenberg B. J. Regulation of nodulation gene expression by NodD in rhizobia. J. Bacteriol. 1992, 174, 5177–5182. 10.1128/jb.174.16.5177-5182.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hou B.; Li F.; Yang X.; Hong G. The properties of NodD were affected by mere variation in length within its hinge region. Acta Biochimica et Biophysica Sinica 2009, 41, 963–971. 10.1093/abbs/gmp090. [DOI] [PubMed] [Google Scholar]
- Chen X. C.; Feng J.; Hou B. H.; Li F. Q.; Li Q.; Hong G. F. Modulating DNA bending affects NodD-mediated transcriptional control in Rhizobium leguminosarum. Nucleic Acids Res. 2005, 33, 2540–2548. 10.1093/nar/gki537. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Burn J. E.; Hamilton W. D.; Wootton J. C.; Johnston A. W. Single and multiple mutations affecting properties of the regulatory gene nodD of Rhizobium. Mol. Microbiol. 1989, 3, 1567–1577. 10.1111/j.1365-2958.1989.tb00142.x. [DOI] [PubMed] [Google Scholar]
- Akakura R.; Winans S. C. Mutations in the occQ operator that decrease OccR-induced DNA bending do not cause constitutive promoter activity. J. Biol. Chem. 2002, 277, 15773–15780. 10.1074/jbc.M200109200. [DOI] [PubMed] [Google Scholar]
- Wang L.; Winans S. C. The sixty nucleotide OccR operator contains a subsite essential and sufficient for OccR binding and a second subsite required for ligand-responsive DNA bending. J. Mol. Biol. 1995, 253, 691–702. 10.1006/jmbi.1995.0583. [DOI] [PubMed] [Google Scholar]
- Wang L.; Helmann J. D.; Winans S. C. The A. tumefaciens transcriptional activator OccR causes a bend at a target promoter, which is partially relaxed by a plant tumor metabolite. Cell 1992, 69, 659–667. 10.1016/0092-8674(92)90229-6. [DOI] [PubMed] [Google Scholar]
- Kim Y.; Chhor G.; Tsai C. S.; Winans J. B.; Jedrzejczak R.; Joachimiak A.; Winans S. C. Crystal structure of the ligand-binding domain of a LysR-type transcriptional regulator: transcriptional activation via a rotary switch. Mol. Microbiol. 2018, 110, 550–561. 10.1111/mmi.14115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Habeeb L. F.; Wang L.; Winans S. C. Transcription of the octopine catabolism operon of the Agrobacterium tumor-inducing plasmid pTiA6 is activated by a LysR-type regulatory protein. Molecular plant-microbe interactions 1991, 4, 379–385. 10.1094/MPMI-4-379. [DOI] [PubMed] [Google Scholar]
- Chen K.; Ke Z.; Wang S.; Wang S.; Yang K.; Jia W.; Zhu J.; Jiang J. Precise regulation of differential transcriptions of various catabolic genes by OdcR via a single nucleotide mutation in the promoter ensures the safety of metabolic flux. Appl. Environ. Microbiol. 2022, 88, e01182-22. 10.1128/aem.01182-22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Choi H. J.; Kim S. J.; Mukhopadhyay P.; Cho S.; Woo J. R.; Storz G.; Ryu S. E. Structural basis of the redox switch in the OxyR transcription factor. Cell 2001, 105, 103–113. 10.1016/S0092-8674(01)00300-2. [DOI] [PubMed] [Google Scholar]
- Toledano M. B.; Kullik I.; Trinh F.; Baird P. T.; Schneider T. D.; Storz G. Redox-dependent shift of OxyR-DNA contacts along an extended DNA-binding site: A mechanism for differential promoter selection. Cell 1994, 78, 897–909. 10.1016/S0092-8674(94)90702-1. [DOI] [PubMed] [Google Scholar]
- Tao K.; Zou C.; Fujita N.; Ishihama A. Mapping of the OxyR protein contact site in the C-terminal region of RNA polymerase α subunit. J. Bacteriol. 1995, 177, 6740–6744. 10.1128/jb.177.23.6740-6744.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zaim J.; Kierzek A. M. The structure of full-length LysR-type transcriptional regulators. Modeling of the full-length OxyR transcription factor dimer. Nucleic Acids Res. 2003, 31, 1444–1454. 10.1093/nar/gkg234. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jo I.; Chung I. Y.; Bae H. W.; Kim J. S.; Song S.; Cho Y. H.; Ha N. C. Structural details of the OxyR peroxide-sensing mechanism. Proc. Natl. Acad. Sci. U.S.A. 2015, 112, 6443–6448. 10.1073/pnas.1424495112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee C.; Lee S. M.; Mukhopadhyay P.; Kim S. J.; Lee S. C.; Ahn W. S.; Yu M. H.; Storz G.; Ryu S. E. Redox regulation of OxyR requires specific disulfide bond formation involving a rapid kinetic reaction path. Nature Structural and Molecular Biology 2004, 11, 1179–1185. 10.1038/nsmb856. [DOI] [PubMed] [Google Scholar]
- Pedre B.; Young D.; Charlier D.; Mourenza A.; Rosado L. A.; Marcos-Pascual L.; Wahni K.; Martens E.; de la Rubia A. G.; Belousov V. V.; Mateos L. M.; Messens J. Structural snapshots of OxyR reveal the peroxidatic mechanism of H2O2 sensing. Proc. Natl. Acad. Sci. U.S.A. 2018, 115, 11623–11632. 10.1073/pnas.1807954115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zheng M.; Storz G. Redox sensing by prokaryotic transcription factors. Biochem. Pharmacol. 2000, 59, 1–6. 10.1016/S0006-2952(99)00289-0. [DOI] [PubMed] [Google Scholar]
- Reen F. J.; Haynes J. M.; Mooij M. J.; O’Gara F. A non-classical LysR-type transcriptional regulator PA2206 is required for an effective oxidative stress response in Pseudomonas aeruginosa. PLoS One 2013, 8, e54479. 10.1371/journal.pone.0054479. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yuan X.; Zeng Q.; Khokhani D.; Tian F.; Severin G. B.; Waters C. M.; Xu J.; Zhou X.; Sundin G. W.; Ibekwe A. M.; Liu F.; Yang C. H. A feed-forward signalling circuit controls bacterial virulence through linking cyclic di-GMP and two mechanistically distinct sRNAs, ArcZ and RsmB. Environmental Microbiology 2019, 21, 2755–2771. 10.1111/1462-2920.14603. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hérault E.; Reverchon S.; Nasser W. Role of the LysR-type transcriptional regulator PecT and DNA supercoiling in the thermoregulation of pel genes, the major virulence factors in Dickeya dadantii. Environmental Microbiology 2014, 16, 734–745. 10.1111/1462-2920.12198. [DOI] [PubMed] [Google Scholar]
- Perrier A.; Barlet X.; Peyraud R.; Rengel D.; Guidot A.; Genin S. Comparative transcriptomic studies identify specific expression patterns of virulence factors under the control of the master regulator PhcA in the Ralstonia solanacearum species complex. Microbial Pathogenesis 2018, 116, 273–278. 10.1016/j.micpath.2018.01.028. [DOI] [PubMed] [Google Scholar]
- Brumbley S. M.; Carney B. F.; Denny T. P. Phenotype conversion in Pseudomonas solanacearum due to spontaneous inactivation of PhcA, a putative LysR transcriptional regulator. J. Bacteriol. 1993, 175, 5477–5487. 10.1128/jb.175.17.5477-5487.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Z.; Huang X.; Nie C.; Xiang T.; Zhang X. The Lon protease negatively regulates pyoluteorin biosynthesis through the Gac/Rsm-RsmE cascade and directly degrades the transcriptional activator PltR in Pseudomonas protegens H78. Environmental Microbiology Reports 2022, 14, 506–519. 10.1111/1758-2229.13057. [DOI] [PubMed] [Google Scholar]
- Li S.; Huang X.; Wang G.; Xu Y. Transcriptional activation of pyoluteorin operon mediated by the LysR-type regulator PltR bound at a 22 bp lys box in pseudomonas aeruginosa M18. PLoS One 2012, 7, e39538. 10.1371/journal.pone.0039538. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Prezioso S. M.; Xue K.; Leung N.; Gray-Owen S. D.; Christendat D. Shikimate induced transcriptional activation of protocatechuate biosynthesis genes by QuiR, a LysR-type transcriptional regulator, in Listeria monocytogenes. J. Mol. Biol. 2018, 430, 1265–1283. 10.1016/j.jmb.2018.03.003. [DOI] [PubMed] [Google Scholar]
- Viale A. M.; Kobayashi H.; Akazawa T.; Henikoff S. rcbR, a gene coding for a member of the LysR family of transcriptional regulators, is located upstream of the expressed set of ribulose 1,5-bisphosphate carboxylase/oxygenase genes in the photosynthetic bacterium Chromatium vinosum. J. Bacteriol. 1991, 173, 5224–5229. 10.1128/jb.173.16.5224-5229.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ki N.; Kim J.; Jo I.; Hyun Y.; Ryu S.; Ha N. C. Isocitrate binds to the itaconic acid-responsive LysR-type transcriptional regulator RipR in Salmonella pathogenesis. J. Biol. Chem. 2022, 298, 102562. 10.1016/j.jbc.2022.102562. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hersch S. J.; Navarre W. W. The salmonella lysr family regulator ripr activates the SPI-13-Encoded itaconate degradation cluster. Infect. Immun. 2020, 10.1128/IAI.00303-20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Heroven A. K.; Dersch P. RovM, a novel LysR-type regulator of the virulence activator gene rovA, controls cell invasion, virulence and motility of Yersinia pseudotuberculosis. Mol. Microbiol. 2006, 62, 1469–1483. 10.1111/j.1365-2958.2006.05458.x. [DOI] [PubMed] [Google Scholar]
- Quade N.; Dieckmann M.; Haffke M.; Heroven A. K.; Dersch P.; Heinz D. W. Structure of the effector-binding domain of the LysR-type transcription factor RovM from Yersinia pseudotuberculosis. Acta Crystallographica Section D: Biological Crystallography 2011, 67, 81–90. 10.1107/S0907444910049681. [DOI] [PubMed] [Google Scholar]
- Kowarz L.; Robbe-Saule V.; Norel F. Identification of cis-acting DNA sequences involved in the transcription of the virulence regulatory gene spvR in Salmonella typhimurium. Mol. Gen. Genet. 1996, 251, 225–235. 10.1007/BF02172922. [DOI] [PubMed] [Google Scholar]
- Sheehan B. J.; Dorman C. J. In vivo analysis of the interactions of the Lysr-like regulator SpvR with the operator sequences of the spvA and spvR virulence genes of Salmonella typhimurium. Mol. Microbiol. 1998, 30, 91–105. 10.1046/j.1365-2958.1998.01041.x. [DOI] [PubMed] [Google Scholar]
- Barnett M. J.; Swanson J. A.; Long S. R. Multiple genetic controls on Rhizobium meliloti syrA, a regulator of exopolysaccharide abundance. Genetics 1998, 148, 19–32. 10.1093/genetics/148.1.19. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Van der Meer J. R.; Frijters A. C.; Leveau J. H.; Eggen R. I.; Zehnder A. J.; De Vos W. M. Characterization of the Pseudomonas sp. strain P51 gene tcbR, a LysR-type transcriptional activator of the tcbCDEF chlorocatechol oxidative operon, and analysis of the regulatory region. J. Bacteriol. 1991, 173, 3700–3708. 10.1128/jb.173.12.3700-3708.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rivas-Marín E.; Floriano B.; Santero E. Genetic dissection of independent and cooperative transcriptional activation by the LysR-type activator ThnR at close divergent promoters. Sci. Rep. 2016, 6, 24538. 10.1038/srep24538. [DOI] [PMC free article] [PubMed] [Google Scholar]
- López-Sánchez A.; Floriano B.; Andújar E.; Hernáez M. J.; Santero E. Tetralin-induced and ThnR-regulated aldehyde dehydrogenase and β-oxidation genes in sphingomonas macrogolitabida strain TFA. Appl. Environ. Microbiol. 2010, 76, 110–118. 10.1128/AEM.01846-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim J.; Kim J. G.; Kang Y.; Jang J. Y.; Jog G. J.; Lim J. Y.; Kim S.; Suga H.; Nagamatsu T.; Hwang I. Quorum sensing and the LysR-type transcriptional activator ToxR regulate toxoflavin biosynthesis and transport in Burkholderia glumae. Mol. Microbiol. 2004, 54, 921–934. 10.1111/j.1365-2958.2004.04338.x. [DOI] [PubMed] [Google Scholar]
- Junker F.; Kiewitz R.; Cook A. M. Characterization of the p-toluenesulfonate operon tsaMBCD and tsaR in Comamonas testosteroni T-2. J. Bacteriol. 1997, 179, 919–927. 10.1128/jb.179.3.919-927.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Monferrer D.; Tralau T.; Kertesz M. A.; Dix I.; Solà M.; Usón I. Structural studies on the full-length LysR-type regulator TsaR from Comamonas testosteroni T-2 reveal a novel open conformation of the tetrameric LTTR fold. Mol. Microbiol. 2010, 75, 1199–1214. 10.1111/j.1365-2958.2010.07043.x. [DOI] [PubMed] [Google Scholar]
- Tralau T.; Mampel J.; Cook A. M.; Ruff J. Characterization of TsaR, an oxygen-sensitive LysR-type regulator for the degradation of p-toluenesulfonate in Comamonas testosteroni T-2. Appl. Environ. Microbiol. 2003, 69, 2298–2305. 10.1128/AEM.69.4.2298-2305.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Crouzet P.; Otten L. Sequence and mutational analysis of a tartrate utilization operon from Agrobacterium vitis. J. Bacteriol. 1995, 177, 6518–6526. 10.1128/jb.177.22.6518-6526.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gao X.; Wang X.; Mao Q.; Xu R.; Zhou X.; Ma Y.; Liu Q.; Zhang Y.; Wang Q. VqsA, a novel LysR-type transcriptional regulator, coordinates quorum sensing (QS) and is controlled by QS to regulate virulence in the pathogen Vibrio alginolyticus. Appl. Environ. Microbiol. 2018, 84, e00444-18. 10.1128/AEM.00444-18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Budnick J. A.; Sheehan L. M.; Ginder M. J.; Failor K. C.; Perkowski J. M.; Pinto J. F.; Kohl K. A.; Kang L.; Michalak P.; Luo L.; Heindl J. E.; Caswell C. C. A central role for the transcriptional regulator VtlR in small RNA-mediated gene regulation in Agrobacterium tumefaciens. Sci. Rep. 2020, 10, 14968. 10.1038/s41598-020-72117-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Eisfeld J.; Kraus A.; Ronge C.; Jagst M.; Brandenburg V. B.; Narberhaus F. A LysR-type transcriptional regulator controls the expression of numerous small RNAs in Agrobacterium tumefaciens. Mol. Microbiol. 2021, 116, 126–139. 10.1111/mmi.14695. [DOI] [PubMed] [Google Scholar]
- García-Tomsig N. I.; Robledo M.; DiCenzo G. C.; Mengoni A.; Millán V.; Peregrina A.; Uceta A.; Jiménez-Zurdo J. I. Pervasive RNA regulation of metabolism enhances the root colonization ability of nitrogen-fixing symbiotic a-Rhizobia. mBio 2022, 13, e03576-21. 10.1128/mbio.03576-21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rodionova I. A.; Gao Y.; Monk J.; Hefner Y.; Wong N.; Szubin R.; Lim H. G.; Rodionov D. A.; Zhang Z.; Saier M. H.; Palsson B. O. A systems approach discovers the role and characteristics of seven LysR type transcription factors in Escherichia coli. Sci. Rep. 2022, 12, 7274. 10.1038/s41598-022-11134-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hong S.; Kim J.; Cho E.; Na S.; Yoo Y. J.; Cho Y. H.; Ryu S.; Ha N. C. Crystal structures of YeiE from Cronobacter sakazakii and the role of sulfite tolerance in gram-negative bacteria. Proc. Natl. Acad. Sci. U.S.A. 2022, 119, e2118002119. 10.1073/pnas.2118002119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Axler-DiPerte G. L.; Miller V. L.; Darwin A. J. YtxR, a conserved LysR-like regulator that induces expression of genes encoding a putative ADP-ribosyltransferase toxin homologue in Yersinia enterocolitica. J. Bacteriol. 2006, 188, 8033–8043. 10.1128/JB.01159-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Snoek T.; Chaberski E. K.; Ambri F.; Kol S.; Bjørn S. P.; Pang B.; Barajas J. F.; Welner D. H.; Jensen M. K.; Keasling J. D. Evolution-guided engineering of small-molecule biosensors. Nucleic Acids Res. 2020, 48, e3. 10.1093/nar/gkz954. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pérez-Rueda E.; Collado-Vides J. The repertoire of DNA-binding transcriptional regulators in Escherichia coli K-12. Nucleic Acids Res. 2000, 28, 1838–1847. 10.1093/nar/28.8.1838. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Baugh A. C.; Momany C.; Neidle E. L. Versatility and complexity: Common and uncommon facets of LysR-type transcriptional regulators. Annu. Rev. Microbiol. 2023, 77, 317–339. 10.1146/annurev-micro-050323-040543. [DOI] [PubMed] [Google Scholar]
- Rosenfeld N.; Elowitz M. B.; Alon U. Negative autoregulation speeds the response times of transcription networks. J. Mol. Biol. 2002, 323, 785–793. 10.1016/S0022-2836(02)00994-4. [DOI] [PubMed] [Google Scholar]
- Xu J. Y.; Xu Y.; Chu X.; Tan M.; Ye B. C. Protein Acylation Affects the Artificial Biosynthetic Pathway for Pinosylvin Production in Engineered E. coli. ACS Chem. Biol. 2018, 13, 1200–1208. 10.1021/acschembio.7b01068. [DOI] [PubMed] [Google Scholar]
- Zhou X.; Lou Z.; Fu S.; Yang A.; Shen H.; Li Z.; Feng Y.; Bartlam M.; Wang H.; Rao Z. Crystal structure of ArgP from mycobacterium tuberculosis confirms two distinct conformations of full-length LysR transcriptional regulators and reveals its function in dna binding and transcriptional regulation. J. Mol. Biol. 2010, 396, 1012–1024. 10.1016/j.jmb.2009.12.033. [DOI] [PubMed] [Google Scholar]
- Porrúa O.; García-González V.; Santero E.; Shingler V.; Govantes F. Activation and repression of a σn-dependent promoter naturally lacking upstream activation sequences. Mol. Microbiol. 2009, 73, 419–433. 10.1111/j.1365-2958.2009.06779.x. [DOI] [PubMed] [Google Scholar]
- Goethals K.; Van Montagu M.; Holsters M. Conserved motifs in a divergent nod box of Azorhizobium caulinodans ORS571 reveal a common structure in promoters regulated by LysR-type proteins. Proc. Natl. Acad. Sci. U.S.A. 1992, 89, 1646–1650. 10.1073/pnas.89.5.1646. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Oliver P.; Peralta-Gil M.; Tabche M. L.; Merino E. Molecular and structural considerations of TF-DNA binding for the generation of biologically meaningful and accurate phylogenetic footprinting analysis: The LysR-type transcriptional regulator family as a study model. BMC Genomics 2016, 17, 686. 10.1186/s12864-016-3025-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pomposiello P. J.; Janes B. K.; Bender R. A. Two roles for the DNA recognition site of the Klebsiella aerogenes nitrogen assimilation control protein. J. Bacteriol. 1998, 180, 578–585. 10.1128/JB.180.3.578-585.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Demeester W.; De Baets J.; Duchi D.; De Mey M.; De Paepe B. MoBioS: modular platform technology for high-throughput construction and characterization of tunable transcriptional. Biosensors 2023, 13, 590. 10.3390/bios13060590. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Beck C. F.; Warren R. A. Divergent promoters, a common form of gene organization. Microbiological reviews 1988, 52, 318–326. 10.1128/mr.52.3.318-326.1988. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pu W.; et al. Directed evolution of linker helix as an efficient strategy for engineering LysR-type transcriptional regulators as whole-cell biosensors. Biosens. Bioelectron. 2023, 222, 115004. 10.1016/j.bios.2022.115004. [DOI] [PubMed] [Google Scholar]
- Browning D. F.; Busby S. J. Local and global regulation of transcription initiation in bacteria. Nature Reviews Microbiology 2016, 14, 638–650. 10.1038/nrmicro.2016.103. [DOI] [PubMed] [Google Scholar]
- Galvão T. C.; De Lorenzo V. Transcriptional regulators á la carte: Engineering new effector specificities in bacterial regulatory proteins. Curr. Opin. Biotechnol. 2006, 17, 34–42. 10.1016/j.copbio.2005.12.002. [DOI] [PubMed] [Google Scholar]
- Garmendia J.; Devos D.; Valencia A.; De Lorenzo V. Á la carte transcriptional regulators: Unlocking responses of the prokaryotic enhancer-binding protein XylR to non-natural effectors. Mol. Microbiol. 2001, 42, 47–59. 10.1046/j.1365-2958.2001.02633.x. [DOI] [PubMed] [Google Scholar]
- Berman H. M.; Westbrook J.; Feng Z.; Gilliland G.; Bhat T. N.; Weissig H.; Shindyalov I. N.; Bourne P. E. The Protein Data Bank. Nucleic Acids Res. 2000, 28, 235–242. 10.1093/nar/28.1.235. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang W.; Wu H.; Xiao Q.; Zhou H.; Li M.; Xu Q.; Wang Q.; Yu F.; He J. Crystal structure details of Vibrio fischeri DarR and mutant DarR-M202I from LTTR family reveals their activation mechanism. Int. J. Biol. Macromol. 2021, 183, 2354–2363. 10.1016/j.ijbiomac.2021.05.186. [DOI] [PubMed] [Google Scholar]
- Della Corte D.; van Beek H. L.; Syberg F.; Schallmey M.; Tobola F.; Cormann K. U.; Schlicker C.; Baumann P. T.; Krumbach K.; Sokolowsky S.; Morris C. J.; Grünberger A.; Hofmann E.; Schröder G. F.; Marienhagen J. Engineering and application of a biosensor with focused ligand specificity. Nat. Commun. 2020, 11, 4851. 10.1038/s41467-020-18400-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jang Y.; Choi G.; Hong S.; Jo I.; Ahn J.; Choi S. H.; Ha N. C. A novel tetrameric assembly configuration in VV2_1132, a LysR-type transcriptional regulator in vibrio vulnificus. Mol. Cells 2018, 41, 301–310. 10.14348/molcells.2018.2190. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee N.; Francklyn C.; Hamilton E. P. Arabinose-induced binding of AraC protein to araI2 activates the araBAD operon promoter. Proc. Natl. Acad. Sci. U.S.A. 1987, 84, 8814–8818. 10.1073/pnas.84.24.8814. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Meinhardt S.; Swint-Kruse L. Experimental identification of specificity determinants in the domain linker of a LacI/GalR protein: Bioinformatics-based predictions generate true positives and false negatives. Proteins: Struct., Funct., Genet. 2008, 73, 941–957. 10.1002/prot.22121. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Meinhardt S.; Manley M. W.; Becker N. A.; Hessman J. A.; Maher L. J.; Swint-Kruse L. Novel insights from hybrid LacI/GalR proteins: Family-wide functional attributes and biologically significant variation in transcription repression. Nucleic Acids Res. 2012, 40, 11139–11154. 10.1093/nar/gks806. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tungtur S.; Egan S. M.; Swint-Kruse L. Functional consequences of exchanging domains between LacI and PurR are mediated by the intervening linker sequence. Proteins: Struct., Funct., Genet. 2007, 68, 375–388. 10.1002/prot.21412. [DOI] [PMC free article] [PubMed] [Google Scholar]
- UniProt: the universal protein knowledgebase in 2023. Nucleic Acids Res. 2023, 51, D523–D531. 10.1093/nar/gkac1052. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Varadi M.; et al. AlphaFold Protein Structure Database: Massively expanding the structural coverage of protein-sequence space with high-accuracy models. Nucleic Acids Res. 2022, 50, D439–D444. 10.1093/nar/gkab1061. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jordan S. R.; Pabo C. O. Structure of the lambda complex at 2.5 Å resolution: Details of the repressor-operator interactions. Science 1988, 242, 893–899. 10.1126/science.3187530. [DOI] [PubMed] [Google Scholar]
- Medvedev K. E.; Kinch L. N.; Dustin Schaeffer R.; Pei J.; Grishin N. V. A fifth of the protein world: Rossmann-like proteins as an evolutionarily successful structural unit. J. Mol. Biol. 2021, 433, 166788. 10.1016/j.jmb.2020.166788. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jørgensen C.; Dandanell G. Isolation and characterization of mutations in the Escherichia coli regulatory protein XapR. J. Bacteriol. 1999, 181, 4397–4403. 10.1128/JB.181.14.4397-4403.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McFall S. M.; Chugani S. A.; Chakrabarty A. M. Transcriptional activation of the catechol and chlorocatechol operons: Variations on a theme. Gene 1998, 223, 257–267. 10.1016/S0378-1119(98)00366-7. [DOI] [PubMed] [Google Scholar]
- Bundy B. M.; Collier L. S.; Hoover T. R.; Neidle E. L. Synergistic transcriptional activation by one regulatory protein in response to two metabolites. Proc. Natl. Acad. Sci. U.S.A. 2002, 99, 7693–7698. 10.1073/pnas.102605799. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fernandez-López R.; Ruiz R.; de la Cruz F.; Moncalián G. Transcription factor-based biosensors enlightened by the analyte. Front. Microbiol. 2015, 6, 648. 10.3389/fmicb.2015.00648. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Murakami K.; Fujita N.; Ishihama A. Transcription factor recognition surface on the RNA polymerase α subunit is involved in contact with the DNA enhancer element. EMBO J. 1996, 15, 4358–4367. 10.1002/j.1460-2075.1996.tb00809.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kredich N. M. Biosynthesis of cysteine. EcoSal Plus 2008, 3, 111–134. 10.1128/ecosalplus.3.6.1.11. [DOI] [PubMed] [Google Scholar]
- Pokhrel A.; Dinh H.; Li L.; Hassan K. A.; Cain A. K.; Paulsen I. T. Identification of a Novel LysR Family Transcriptional Regulator Controlling Acquisition of Sulfur Sources in Acinetobacter baumannii. Microbial Physiology 2023, 33, 27–35. 10.1159/000529038. [DOI] [PubMed] [Google Scholar]
- Bykowski T.; van der Ploeg J. R.; Iwanicka-Nowicka R.; Hryniewicz M. M. The switch from inorganic to organic sulphur assimilation in Escherichia coli: Adenosine 5′-phosphosulphate (APS) as a signalling molecule for sulphate excess. Mol. Microbiol. 2002, 43, 1347–1358. 10.1046/j.1365-2958.2002.02846.x. [DOI] [PubMed] [Google Scholar]
- Härtig E.; Frädrich C.; Behringer M.; Hartmann A.; Neumann-Schaal M.; Jahn D. Functional definition of the two effector binding sites, the oligomerization and DNA binding domains of the Bacillus subtilis LysR-type transcriptional regulator AlsR. Mol. Microbiol. 2018, 109, 845–864. 10.1111/mmi.14089. [DOI] [PubMed] [Google Scholar]
- Ostrowski J.; Kredich N. M. Molecular characterization of the cysJIH promoters of Salmonella typhimurium and Escherichia coli: Regulation of cysB protein and N-acetyl-L-serine. J. Bacteriol. 1989, 171, 130–140. 10.1128/jb.171.1.130-140.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
- López-Sánchez A.; Rivas-Marín E.; Martínez-Pérez O.; Floriano B.; Santero E. Co-ordmated regulation of two divergent promoters through higher-order complex formation by the LysR-type regulator ThnR. Mol. Microbiol. 2009, 73, 1086–1100. 10.1111/j.1365-2958.2009.06834.x. [DOI] [PubMed] [Google Scholar]
- Wall M. E.; Hlavacek W. S.; Savageau M. A. Design of gene circuits: Lessons from bacteria. Nat. Rev. Genet. 2004, 5, 34–42. 10.1038/nrg1244. [DOI] [PubMed] [Google Scholar]
- Bender R. A. A NAC for regulating metabolism: The nitrogen assimilation control protein (NAC) from Klebsiella pneumoniae. J. Bacteriol. 2010, 192, 4801–4811. 10.1128/JB.00266-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Feng J.; Goss T. J.; Bender R. A.; Ninfa A. J. Repression of the Klebsiella aerogenes nac promoter. J. Bacteriol. 1995, 177, 5535–5538. 10.1128/jb.177.19.5535-5538.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rosario C. J.; Bender R. A. Importance of tetramer formation by the nitrogen assimilation control protein for strong repression of glutamate dehydrogenase formation in Klebsiella pneumoniae. J. Bacteriol. 2005, 187, 8291–8299. 10.1128/JB.187.24.8291-8299.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Imlay J. A. Cellular defenses against superoxide and hydrogen peroxide. Annu. Rev. Biochem. 2008, 77, 755–776. 10.1146/annurev.biochem.77.061606.161055. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Clark T. J.; Momany C.; Neidle E. L. The benPK, operon, proposed to play a role in transport, is part of a regulon for benzoate catabolism in Acinetobacter sp. strain ADP1. Microbiology 2002, 148, 1213–1223. 10.1099/00221287-148-4-1213. [DOI] [PubMed] [Google Scholar]
- Suen W. C.; Spain J. C. Cloning and characterization of Pseudomonas sp. strain DNT genes for 2,4- dinitrotoluene degradation. J. Bacteriol. 1993, 175, 1831–1837. 10.1128/jb.175.6.1831-1837.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Spanggord R. J.; Spain J. C.; Nishino S. F.; Mortelmans K. E. Biodegradation of 2,4-dinitrotoluene by a Pseudomonas sp. Appl. Environ. Microbiol. 1991, 57, 3200–3205. 10.1128/aem.57.11.3200-3205.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.



