Evolution of pH-sensitive transcription termination in Escherichia coli during adaptation to repeated long-term starvation

Edited by Richard Lenski, Michigan State University, East Lansing, MI; received March 17, 2024; accepted August 19, 2024
September 19, 2024
121 (39) e2405546121

Significance

We identified a pH-sensitive amino acid substitution that repeatedly arose in bacterial populations experimentally evolved under extreme cycles of starvation. This substitution affects the RNA binding of a transcription termination factor and bypasses traditional cell signaling to facilitate a rapid transcriptional response during alkaline/neutral pH shifts. As fluctuating conditions represent a universal challenge faced by macro- and microorganisms, pH-sensing mutations within RNA- or DNA-binding domains of global regulators represent an adaptation that can coordinate immediate cellular responses based on seasonal and recurring environmental cues.

Abstract

Fluctuating environments that consist of regular cycles of co-occurring stress are a common challenge faced by cellular populations. For a population to thrive in constantly changing conditions, an ability to coordinate a rapid cellular response is essential. Here, we identify a mutation conferring an arginine-to-histidine (Arg to His) substitution in the transcription terminator Rho. The rho R109H mutation frequently arose in Escherichia coli populations experimentally evolved under repeated long-term starvation conditions, during which the accumulation of metabolic waste followed by transfer into fresh media results in drastic environmental pH fluctuations associated with feast and famine. Metagenomic sequencing revealed that populations containing the rho mutation also possess putative loss-of-function mutations in ydcI, which encodes a recently characterized transcription factor associated with pH homeostasis. Genetic reconstructions of these mutations show that the rho allele confers plasticity via an alkaline-induced reduction of Rho function that, when found in tandem with a ΔydcI allele, leads to intracellular alkalization and genetic assimilation of Rho mutant function. We further identify Arg to His substitutions at analogous sites in rho alleles from species that regularly experience neutral to alkaline pH fluctuations in their environments. Our results suggest that Arg to His substitutions in Rho may serve to rapidly coordinate complex physiological responses through pH sensing and shed light on how cellular populations use environmental cues to coordinate rapid responses to complex, fluctuating environments.
Cellular populations regularly encounter fluctuating environmental stressors in their natural contexts, such as changes in resource availability, temperature, pH, and oxidative stress (13). This holds particularly true for microorganisms, which have evolved to thrive in some of Earth’s most extreme and fluctuation-prone environments. Even in the most forgiving of habitats, microbial growth is primarily limited due, in part, to the scarcity of carbon and nitrogen sources and competition between microbes (46). Given that resource availability fluctuates temporally and spatially in natural environments (3), microbial growth in nature occurs in bursts as resources become accessible, resulting in long, stagnant periods of starvation punctuated by transient periods of growth, commonly referred to as feast/famine cycles (4, 7). Previous work has revealed several common microbial responses to persistent famine conditions, particularly those experienced by batch cultures during long-term stationary phase. These responses include an increase in genetic diversity, elevated mutation rates, morphological changes, and the emergence of the growth advantage in stationary phase (GASP) phenotype which confers an increased ability to catabolize alternative energy sources, such as amino acids, present in necromass as well as modulation of the stress response (814). Building upon this, recent work suggests that bacteria exhibit distinct responses depending on the frequency and magnitude of nutrient resource fluctuations, indicating that these fluctuations are an important determinant of bacterial behavior (3, 1517). Notably, microbial populations subjected to cycles of prolonged starvation followed by a rapid influx of resources (i.e., 100-d feast/famine cycles) still exhibit GASP-like behavior but evolve mutations that are distinct from the mutations traditionally associated with the GASP phenotype (18). Thus, further investigation of how bacteria adapt to repeated nutrient resource fluctuations is critical to contextualizing microbial behavior in the wild.
Among the complications that microbes face in the wild, the reintroduction of nutrient resources into environments presents an inherent complex stress. Here, microbes must contend with additional stressors that coincide with nutrient resource replenishment, such as shifts in pH. Microbes are further challenged as they metabolize the newly introduced resources since excreted metabolic by-products can result in further environmental changes. These ecological feedbacks are also common in laboratory environments, particularly the significant pH fluctuations that can accompany growth (19). For example, when Escherichia coli populations are cultivated in the amino acid-based LB broth, their metabolic activities induce remarkable shifts in pH (20, 21). As these bacteria metabolize the amino acids for carbon, they release ammonia, causing the medium pH to undergo a dramatic transformation, shifting from a neutral pH of 7.0 to an alkaline pH of 9.0 in approximately 24 h (22, 23).
Consequently, many bacterial species have evolved mechanisms to sense and rapidly respond to changes in proton concentrations (24). For pathogens and marine bacteria which are frequently exposed to drastic shifts in external pH, the ability to sense these changes and respond through phenotypically plastic traits is critical to their survival (2528). As such, microbes encode a number of systems dedicated to pH homeostasis, including amino acid–dependent acid resistance systems such as the GAD regulon in E. coli (29), two-component regulatory systems such as the MtrAB system in Actinobacteria (30) or the BatRS system in Bartonella, and one-component systems such as CadC in E. coli (31). Additionally, more global systems like the RpoS regulon control the expression of many components involved in the pH response (32). Experimental evolution under high-pH conditions has recapitulated the role of RpoS in the pH response through mutations, as well as revealed unexpected targets such as the two-component response regulator PhoB (33). Yet, another more universal plastic mechanism for pH sensing is the modulation of protein function via the protonation state of histidine residues, which is present across the Tree of Life. Histidine residues are often situated within functionally important sites of proteins, where their protonation status dictates the protein’s functionality (3436). The protonation state of histidine residues also regulates the activity of particular cell-surface receptors and transporters, such as G-coupled receptors (37) and ion transporters (38, 39). Moreover, tumorigenic arginine to histidine mutations are reported to contribute to cancers where they often confer pH sensing abilities (40, 41). Thus, histidine residues, acting as pH sensors, can serve as a plastic mechanism to enable rapid responses to changing environments.
One powerful approach for investigating how organisms navigate complex stresses and identifying the resulting mutations contributing to adaptation is experimental evolution (42, 43). Phenotypically plastic traits can evolve rapidly in evolution experiments and through these investigations, it has become evident that genes encoding global regulators of gene expression are frequent targets of selection when adapting to diverse environments, primarily due to their widespread impacts on various biological processes (44, 45). One such essential global regulator is Rho, which terminates transcription at hundreds of sites throughout the E. coli genome (4649). Rho-dependent termination plays critical cellular roles, including the delineation of transcription unit boundaries and surveillance of the gene expression machinery (50), as well as involvement in many conditional gene regulatory mechanisms (51, 52). Additionally, Rho prevents unwanted transcription from loci antisense to genes or horizontally acquired genes, such as toxic prophage genes (46, 47). Functioning as a ring-shaped homohexamer, Rho utilizes its crown-like primary binding site to bind nascent transcripts at poorly conserved C-rich regions called rut (Rho utilization) sites. Once bound to a rut site, the Rho hexamer guides the transcript into its central ring channel, where its secondary binding site mediates ATP-dependent [5′ to 3′] movement along the RNA chain until Rho catches up with and dissociates RNA polymerase, effectively terminating transcription (53, 54).
Given the widespread prevalence of Rho-dependent termination in E. coli, any alteration in Rho function has the propensity to trigger extensive physiological changes. As such, numerous previous studies employing adaptive laboratory evolution under various conditions have reported nonsynonymous mutations in the rho gene. These conditions range from adaptation to heat stress (5558) and exposure to copper (59) and ethanol (60, 61) to the utilization of various carbon sources (62, 63). Notably, previously reported adaptive mutations in rho exhibit variability, affecting residues distributed throughout the gene, particularly in regions associated with its core functions, including the primary and secondary binding site domains. Hence, the pleiotropic nature of Rho positions it as a frequently positively selected genetic target in E. coli when confronted with novel and intricate environmental challenges.
Here, we investigate a mutation in rho that results in an R109H substitution that fixed in 5 out of 16 E. coli populations during experimental evolution in repeated long-term starvation (RLTS) conditions consisting of 100 d between transfers into fresh LB-Miller broth (64). Each of these populations also contained putative loss of function mutations in an additional gene, ydcI, which encodes a poorly characterized transcriptional regulator related to pH homeostasis (65). Since arginine-to-histidine substitutions can alter a protein’s affinity to bind nucleic acids and are shown to grant pH-dependent activity to transcription factors in cancer (41), we hypothesize that the activity of RhoR109H exhibits a similar pH dependence that serves as a plastic adaptation to the neutral/alkaline fluctuations experienced during repeated starvation in LB broth. Using a combination of in vitro biochemical assays, fluorescence microscopy, and competitive coculture, we examined how rho R109H and ΔydcI mutations contribute to the physiological adaptation of E. coli during repeated starvation. Additionally, through bioinformatic approaches, we further identify rho R109H alleles in wild bacterial strains that regularly experience neutral/alkaline fluctuations and infer amino acid residues of structural and functional importance within YdcI.

Results

Extensive Parallel Mutation in rho and ydcI During RLTS.

To gain insight into how microbes adapt to recurrent nutrient resource fluctuations following prolonged periods of famine, we evolved 16 E. coli populations in parallel to a RLTS regime of 100-d feast/famine cycles for a total of 900 d (64). Every 100 d, we transferred 1 mL of the aged culture into 9 mL of fresh LB-Miller broth, replenishing each culture’s resources. At this time, pretransfer aliquots of the populations were frozen to maintain a historical record of the evolution experiment. Growth in LB broth is supported through the catabolism of amino acids and leads to alkalinization of the medium within 24 h which remains between pH 8 and pH 9 until additional resources are provided (21, 22, 66). As such, E. coli cultures are subjected to resource-depleted, alkaline conditions for the vast majority of the RLTS regime, essentially 99 d (SI Appendix, Fig. S1A). Every 100 d, when cultures are replenished with resources, the medium pH decreases dramatically from approximately pH 9 to pH 7, before slowly returning to an alkaline pH over the next 24 h. Thus, populations adapted to these conditions must not only survive extended starvation under alkaline conditions but also cope with the sudden downward shift in pH and repletion of resources concurrently.
Metagenomic sequencing throughout the 900-d experiment revealed that approximately half (7/16) of the populations evolved fixed mutations in rho, which encodes the Rho-dependent transcription termination factor (Fig. 1A and SI Appendix, Fig. S2A). The majority of rho mutations (5/7) occurred in the first half of the RLTS experiment and swept to fixation within the first 400 d (Fig. 1B). By 700 d, rho mutations were fixed in a total of seven populations after which time no additional rho mutations were detected in the remaining nine populations. All identified rho mutations occur within the primary RNA-binding site of Rho. Thus, we suspect these mutations are likely to perturb interactions between Rho and its RNA target (Fig. 1 C and E and SI Appendix, Fig. S3). Furthermore, all five rho mutations that are fixed within the first 400 d result in an R109H nonsynonymous substitution, suggesting it may contribute to adaptation in RLTS conditions. After reconstructing the rho R109H allele into experimental ancestor background, we found that the R109H allele did not confer any considerable growth effects in conditions that previously produced other experimentally evolved rho mutations, including heat stress (56), exposure to ethanol (61), and osmotic stress (67) (SI Appendix, Fig. S4). This lack of phenotype is consistent with Rho’s role as a pleiotropic “master-regulator” (68) and indicates that any phenotypes associated with the rho R109H allele may be contingent on specific conditions that occur during RTLS or dependent on an earlier arising mutation.
Fig. 1.
RLTS selects for mutations in rho and ydcI. (A) Number of populations containing polymorphic (dashed line) or fixed (solid line) mutations in the rho (red) or ydcI (blue) loci over 900 d of evolution to RLTS conditions. (B) Number of fixed mutations in rho (top half) and ydcI (bottom half) genes indicating the timing at which each particular mutation arose over 900 d. Purple lines indicate alleles co-occurring within populations. The genic location of affected (C) rho (orange: R109H, purple: G63S, green: K105E) and (D) ydcI residues (red: R10H, blue: R30C, purple: F74Y, magenta: D83G, cyan: S91P, green: V181A, orange: L198Q) and (E) structural locations of affected rho and (F) ydcI residues identified in populations over 900 d of evolution to RLTS conditions. Only the A and B chains of the YdcI homotetramer are visualized for simplicity. Structural locations colored in green indicate H-T-H domain residues. NTD: N-terminal domain, CTD: C-terminal domain, NHB: N-terminal helix bundle, CSD: cold shock domain, RBD: RNA-binding domain, DBD: DNA-binding domain, RD: regulatory domain, HTH: helix–turn–helix.
Since prior investigations implicate the evolution of arginine to histidine substitutions as a response to pH (40, 41, 69, 70), and histidine residues can serve as pH sensors (39, 71, 72), we refocused our investigation to include genes shaped by positive selection during RTLS and associated with pH homeostasis. We found that all populations with mutant rho alleles also contain mutations in the recently annotated ydcI gene, which encodes a LysR-type transcriptional regulator reported to affect the expression of around 60 genes in E. coli, some of which are associated with the pH response (65, 73) (Fig. 1 A and B and SI Appendix, Fig. S2B). Mutations in ydcI are pervasive and evolved in the vast majority (13/16) of RTLS populations (SI Appendix, Table S1). In each of these cases, the ydcI mutation arose simultaneously or before the rho mutation, suggesting that mutations in ydcI may be necessary for the fixation of mutations in rho.
In contrast to the localized mutations identified in rho, the mutations in ydcI are diverse, ranging from putative loss of function mutations—caused by IS-element insertions and small indels—to nonsynonymous SNPs whose functional consequences on YdcI activity are currently unknown (Fig. 1D). Nonsynonymous mutations in ydcI tend to cluster into three general genic regions. Using previous structural data for LysR-type proteins and the AlphaFold predicted structure for YdcI, we mapped the locations of these mutations to specific domains to gain insight into their potential functional impact (Fig. 1F). The affected residues R10 and R30 are found in the predicted helix–turn–helix domain and could lead to an abolishment of DNA-binding activity. Residues F74, D83, and S91 are all found in the α4 linker helix, and residues V181 and L198 are present in the junction between the two regulatory domains, regions that are critical to dimerization and effector binding, respectively, in well-studied LysR-type proteins (74). Together, this suggests that mutations resulting in loss or attenuation of YdcI function are under positive selection in RLTS conditions, and disruption of YdcI function may be necessary for the rho R109H allele to be beneficial.

RhoR109H Mutants Display a pH-Sensitive In Vitro Phenotype.

Previous investigations have identified arginine to histidine substitutions that alter protein functionality in alkaline conditions (41, 70). Thus, we hypothesized that the activity of RhoR109H may differ from that of WT Rho (Rho+) under alkaline conditions. To this end, we performed in vitro assays evaluating the capability of Rho to function as a helicase, terminate transcription, and bind RNA in neutral (pH 7) and alkaline (pH 9) conditions (Fig. 2 and SI Appendix, Fig. S5). At pH 7, we found no significant differences in the helicase activity (t test; P = 0.829) or transcription termination patterns of RhoR109H and only modest differences in RNA affinity (t test, P = 0.033) compared to Rho+. By contrast, at pH 9, RhoR109H exhibits significant reductions in its helicase activity (t test; P = 5.54 × 10−5), a complete inability to terminate transcription, and a significant decline in RNA affinity (t test, P = 0.003), while Rho+ remained functional. Moreover, Rho+ binding to RNA is largely unaffected by the change in pH conditions (t test, P = 0.068), while there is a significant reduction in RhoR109H RNA binding as pH increases (t test, P = 0.004). These data collectively demonstrate that the activity of the RhoR109H mutant is similar to the WT protein under neutral conditions, but is strongly reduced under alkaline conditions due, in part, to compromised RNA binding.
Fig. 2.
The in vitro Rho assay reveals diminished activity of RhoR109H at pH 9. The in vitro activities of Rho+ (WT) and RhoR109H (R109H) were assessed at pH 7 and pH 9 in three ways: (A) efficiency of ATP-dependent unwinding of a RNA:DNA duplex containing a 5′-single-stranded rut overhang (graph: helicase activity after 20 min), (B) ability to terminate transcription at the λtR1 terminator, and (C) ability to bind to RNA. A schematic of the DNA template used in the transcription termination assay is shown above gel images in (B). Chart above plots in (C) contains the calculated average Kd for WT and R109H proteins. Additional information regarding binding parameters can be found in Table 1.
Table 1.
Equilibrium binding parameters for Rho–RNA complexes
 WTR109H
Parameters*pH 7pH 9pH 7pH 9
n1.2 ± 0.21.3 ± 0.21.6 ± 0.21.7 ± 0.2
Kd (nM)1.5 ± 0.20.8 ± 0.22.4 ± 0.220 ± 3
*
Binding parameters were determined by nonlinear least-squares fit of the slot-blot data to the Hill equation: Fs = Fmax[Rho]n/([Rho]n +Kdn), where Fmax is the maximal fraction of Rho–RNA complexes, Kd is the dissociation constant, and n is the Hill parameter reflecting binding cooperativity.
Of the transcripts where termination is Rho dependent, a fraction of termination events require an additional transcription-termination factor, NusG (47, 49), such as transcription termination that represses toxic prophage genes (46). These NusG-stimulated terminators have been shown to contain distinct rut features (47, 49) and are less sensitive to two disruptive primary binding site mutations, RhoF62S and RhoY80C, than canonical terminators (75). To test the possibility that the R109H mutation may differentially impact the two classes of terminators, we conducted in vitro transcription experiments with two DNA templates encoding the NusG-stimulated terminators located within aspS and yaiC (49). With Rho+, termination at the aspS and yaiC sites is strongly stimulated by NusG at both pH 7 and pH 9 (SI Appendix, Fig. S6). Termination with the RhoR109H mutant is also stimulated by NusG at pH 7, albeit to a lesser extent than Rho+. At pH 9, however, RhoR109H can no longer trigger termination at the aspS and yaiC sites, even in the presence of NusG. These data indicate that pH regulates the activity of the RhoR109H mutant at both NusG-stimulated and canonical Rho-dependent terminators (Fig. 2B).

Cells Containing rho R109H/ΔydcI Alleles Have an Altered Intracellular pH Response.

The finding that RhoR109H displays altered activity under alkaline conditions implies there must be changes in intracellular pH (pHi) to reveal the pH-dependent phenotype. To ascertain whether the introduction of the rho R109H and/or ΔydcI alleles results in an alteration of pHi homeostasis, we measured pHi across a range of pH-adjusted buffers (pH 6.92 to 9.59) by staining cells with BCECF-AM (7679) and compared their pHi response to that of WT cells. As the intensity of the fluorescent signal emitted from BCECF-stained cells depends on pH (78), we quantified the pHi of single cells with fluorescent microscopy.
Across the environmental pH conditions, WT and rho R109H single-mutant cells exhibited similar pHi values with a fitted range from 6.7 to 9.05 and 7.04 to 8.96, respectively (Fig. 3 and SI Appendix, Figs. S8 and S9). Alternatively, cells containing a single ΔydcI mutation showed a reduced pHi range of 7.48 to 8.92, as ΔydcI cells maintain a higher pHi than WT cells in acidic and neutral conditions (Fig. 3 and SI Appendix, Fig. S10). This effect is amplified in the double mutant rho R109H/ΔydcI cells which exhibit a pHi range of 7.53 to 9.67, maintaining a significantly elevated pHi across the entire range of environmental pH conditions compared to the WT (sign test; n = 8; P = 0.008) (Fig. 3C and SI Appendix, Fig. S11). Because pHi elevation in neutral conditions is only realized in mutant backgrounds containing ΔydcI, it is likely that double mutant rho R109H/ΔydcI cells constantly maintain an intracellular pH environment that continuously manifests the pH-sensitive Rho termination activity via reduced RNA binding.
Fig. 3.
rho R109H/ΔydcI cells have increased intracellular pH. The intracellular pH (pHi) of WT, rho R109H, ΔydcI, and rho R109H/ΔydcI single cells was measured in buffers with increasing pH values. These measurements were derived from a ladder control (Control) containing WT cells treated with nigericin, yielding the pHi and environmental pH equal. (A) Representative microscopic images of single cells from tested strains in each pH buffered solution with colors of cells corresponding to their calculated pHi. Each image is cropped to 5.5 × 5.5 μm; a 1 μm scale bar is included in the bottom right image. The remaining microscopic images can be found in supplementary files. (B) pHi measurements derived from these images across the environmental pH conditions. Lightly shaded data points represent single-cell replicates with the color denoting the pHi. Error bars display 95% CI. (C) Plot displaying the pHi range of each strain in tested conditions using data from panel (B) illustrates an upward shift in the pHi range in the rho R109H/ΔydcI cells.

Benefits of rho R109H and ΔydcI Alleles Vary in RLTS Environment.

Considering the frequent co-occurrence of the rho R109H and ΔydcI alleles in evolved populations and the necessity of both alleles for a significant alteration in pHi response, we hypothesized that the benefit of carrying the rho R109H allele could only be present in a ΔydcI background. To address this, we “replayed” one cycle of the RLTS regime by growing cultures from different starting genotypes (WT, rho R109H, ΔydcI) for 100 d and sequenced the rho and ydcI loci from the aged cultures (Fig. 4A and SI Appendix, Fig. S12). We found that when starting from a rho R109H genotype, 8/12 populations evolved putative ΔydcI mutations which swept to fixation within 100 d. Alternatively, when starting from a ΔydcI genotype, only 3/12 populations evolved a rho R109H mutation, and only one of these rho R109H mutations had fixed. Importantly, when starting from a WT genotype, we did not observe the emergence of the rho R109H allele in a population without also detecting a fixed mutation in ydcI. Thus, ydcI loss-of-function mutations are recurrent in RLTS conditions and the appearance of the rho R109H mutation is unlikely to occur without a prior mutation in the ydcI gene.
Fig. 4.
Benefits of rho R109H and ΔydcI alleles vary throughout the RLTS regime. (A) Frequency of final genotypes identified in populations after 12 replicate cultures of WT, ΔydcI, and rho R109H (initial genotypes) were grown in RLTS conditions for 100 d. A subscript of p denotes that the allele is polymorphic as indicated by the Sanger sequence. (B) Pairwise competitions between WT ancestor and mutant strains when cocultured for 1 or 14 d in unbuffered LB, LB buffered at pH 9, and LB buffered to pH 7. Colored bars indicate the median selection rate (s) for each mutant strain relative to the WT with single points representing each of three biological replicates. (C) Median CFU/mL of WT and mutant strains over 14 d of growth. Lightly shaded data points represent each of three biological replicates. Shaded vertical bars indicate the day of the first initial growth rebound postdeath phase. Error bars display 95% CI.
Identifying the precise conditions under which each allele confers an advantage poses a challenge due to the continual evolution of the strains and the rapid accumulation of mutations as cultures age. While this is evident after “replaying” the evolution experiment, it may also account for the lack of significant differences in population cell density between strains over 100 d (SI Appendix, Fig. S1B). Consequently, we opted to conduct relevant phenotyping experiments over shorter periods to minimize any confounding effects potentially associated with the strains’ evolution. To mimic conditions experienced by populations during the RLTS regime, we performed pairwise competitions between each mutant and the WT ancestor in unbuffered, buffered pH 9, and buffered pH 7 LB medium for 1 and 14 d (Fig. 4B). When competing the WT ancestor against the rho R109H strain, we observe an early benefit to carrying the mutant rho allele after the first day of coculture, particularly in the pH 9 medium. However, after 14 d of coculture the WT ancestor strongly outcompetes the rho R109H strain in the unbuffered, buffered pH 9, and buffered pH 7 media. In contrast to rho R109H, ΔydcI appears to exhibit considerably greater selection rates across the 14 d as the ΔydcI strain outcompetes the WT ancestor at both time points and in all of the tested conditions. Notably, when the rho R109H allele is present in a ΔydcI background, we no longer observe highly negative selection rates after 14 d. Instead, the double rho R109H/ΔydcI mutant strain outcompetes the WT strain under the unbuffered and buffered pH 9 conditions and maintains a high selection rate throughout the experiment, while in unbuffered pH 7 conditions it is neutral in coculture with the ancestor. These results indicate that the rho R109H allele is advantageous when resources are plentiful, especially in alkaline conditions, and disrupting the ydcI gene can mediate any tradeoffs associated with the rho R109H allele in resource-limited long-term stationary phase conditions.
In addition to assessing growth and survival in coculture, assessment of each strain in monoculture revealed no significant differences in the mutants’ survival over 14 d. The end of the stationary phase and subsequent onset of the death phase marks a considerable decline in population cell viability, which we consistently observe on day 3 of growth across all strains (80) (Fig. 4C and SI Appendix, Fig. S1B). Next, the transition into long-term stationary phase is characterized by a resurgence in population density following the death phase, attributed to the substantial availability of dead cells providing resources for the surviving population (80). This resurgence in population cell density is highly repeatable, occurring on the 6th day of growth for the WT ancestor, rho R109H, and ΔydcI strains (Fig. 4C). Notably, rho R109H/ΔydcI populations exhibit a resurgence a day earlier on day 5 of growth. This early population resurgence following the death phase highlights a unique advantage of rho R109H/ΔydcI as these cells can better utilize new resources as they become available, such as necromass and other metabolites released by dying cells.

Sequence Analysis Identifies rho R109H Alleles in Natural Isolates.

Given the widespread presence of Rho across the bacterial domain and its essential role in many prokaryotes (81), we investigated the prevalence and distribution of a histidine residue at analogous positions within Rho (Fig. 5). Our initial search of the RefSeq database did not return any bacterial sequences with histidine at this location. However, a deeper search using PHI-BLAST identified several sequences from species isolated from diverse environments that harbor a histidine residue at positions analogous to the E. coli Rho R109. Many of these isolates originated from alkaline environments or those which fluctuate between alkaline and neutral conditions. The soil bacterium Inquilinus limosus was isolated from a region that reports alkalization of the soil (82, 83), while the Poribacteria genome was identified as part of a microbial mat in hypersaline, alkaline water (84, 85). The pathogen Bartonella bacilliformis, the causative agent of Carrion’s disease, a neglected tropical disease, also encodes a histidine residue in Rho that may confer a pH-sensing phenotype as the pathogen alternates between neutral human blood and the alkaline midgut of sand flies (25, 86). The remaining species are isolated from aquatic environments with reported pH gradients. Planctomycetota was identified in sediment from the Auka marine hydrothermal vent (pH 6.5) in the Pacific Ocean (pH 8) (87), while Elusimicrobia was isolated from groundwater with a pH ranging from pH 6.5 to 8.5 (88, 89). Together, the absence of a histidine residue in the analogous position 109 of Rho in various lab-adapted bacterial reference strains, as well as the identification of Rho H109 residues in isolates that experience alkaline environments, is consistent with our results showcasing that this allele provides beneficial effects in alkaline environments.
Fig. 5.
Amino acid alignment of the Rho N-terminal RNA-binding domain in laboratory strains and natural isolates. Alignments of Rho sequences from selected (A). RefSeq genomes and the mutated R109H E. coli clone with natural (B) non–E. coli isolates that encode a Histidine residue at the analogous position. Natural non–E. coli isolates were identified with PSI-BLAST. For both alignments, NCBI accession numbers for the protein sequence are included on the y axis. Coordinates on the x-axis indicate analogous positions in E. coli Rho.

Discussion

Here, we report an adaptive pH-sensing mutation in the global transcriptional terminator Rho (rho R109H) that emerged independently in multiple E. coli populations subjected to 900 d of RLTS conditions. In vitro analyses revealed a pronounced pH-dependent loss of function, where RhoR109H activity exhibited a sharp decline in the transition from neutral pH 7 to alkaline pH 9 conditions. This mutation was exclusively observed in strains that also contained a mutation in the transcriptional regulator gene ydcI, implying that in E. coli the rho R109H allele is beneficial in the RLTS environment when YdcI activity is attenuated or nonfunctional. Measurements of the pHi provide insight into the coexistence of these two alleles as we found that double mutant cells that carry both the rho R109H and ΔydcI alleles exhibit a significantly more alkaline pHi, with respect to WT cells, across a spectrum of pH-adjusted solutions. These results imply that the intracellular environment of rho R109H/ΔydcI cells is conducive to conditions where RhoR109H termination would be impaired. Pairwise competitions between the mutants and WT ancestor pinpoint distinct advantages of each allele in the RTLS environment. Results of these competitions suggest that the rho R109H allele is beneficial during nutrient resource repletion as this mutation alone confers a fitness benefit only when resources are abundant, with an enhanced fitness when grown in LB broth adjusted to pH 9. However, this advantage is transient, as by 14 d of coculture, the rho R109H allele is no longer beneficial by itself. Instead, we observed that the deletion of ydcI masks the detrimental effects of the rho R109H allele during resource-depleted conditions, illustrating that the elimination of YdcI function generates an intracellular environment that better tolerates the plasticity of the rho R109H allele. Collectively, these data reveal a nuanced interplay between both alleles as they jointly provide unique contributions to navigating the complex stresses inherent to RLTS.
Prior research investigating the consequences of Rho impairment and the effects of various rho mutations on gene expression offers insight into the potential impacts of the rho R109H allele on resource utilization during pH fluctuations. In general, the impairment of Rho activity leads to an overall enrichment in transcripts due to transcriptional readthrough of Rho-dependent terminators (90). However, it appears that the identity and degree of differentially expressed genes vary based on the location and nature of the rho mutation, creating a challenge in pinpointing the exact benefit of the rho R109H allele. Although adaptive Rho mutations in residues that constitute the primary binding site are less common (91, 92), these mutants do appear to be less disruptive to gene expression and have less enrichment in prophage gene expression, compared to evaluated secondary binding site mutants, which could partially explain why the rho R109H mutation is not detrimental (46, 75). In some primary binding site mutants, a defect in RNA binding can be overcome at terminators with high NusG dependence (75). However, the RhoR109H in vitro activity at two such terminators, aspS and yaiC, is also impaired in pH 9 conditions, signifying that the effects of this mutation vary from previously described primary binding site mutants. Adaptive mutations in rho can confer the ability to use novel substrates due to the upregulation of many metabolic pathways (60, 93). Rho has been demonstrated to play a role in the pH response as changes in Rho activity can alter pH homeostasis (94). Rho also controls the expression of genes related to the pH response (46), including tryptophanase, one of the most abundant proteins present in LB broth buffered to pH 9 (95). Adaptation to complex stresses elicits a distinct response that differs from the response to individual stresses (96). Thus, the pleiotropic nature of Rho may enable it to effectively navigate complex stress (68).
In contrast to our current understanding of Rho’s function, very little is known about YdcI and its regulatory targets. What is understood is that ydcI is found across closely related enterobacterial genera including, Escherichia, Salmonella, Citrobacter, and Klebsiella, and its regulatory function has since diverged from its most characterized role in regulating acid-stress responses in S. Typhimurium to a more ambiguous role in mediating acid and alkaline stress in E. coli (73). While there are differing reports of the loci regulated by ydcI in E. coli, there is agreement on its role in activating expression of the rpoS stabilization protein iraP (65, 73). Thus, while we currently do not know exactly which genes are differentially expressed due to the rho R109H/ΔydcI double mutation, it is likely that changes in gene expression induced by these mutations in alkaline conditions impact pH and other stress responses. Since we do not observe fixation of the rho R109H mutation in cells without a deletion of ydcI, it is possible that the combination of these two alleles may relieve any potential detrimental effects on gene expression induced by alteration of Rho activity, especially in the long-term stationary phase portion of the RLTS regime, and help to explain why YdcI function has to be eliminated in order for the rho R109H allele to be beneficial.
The necessity of a permissive mutation, such as the disruption of ydcI, is not unprecedented as adaptive nonsynonymous mutations in rho often do not confer immediate fitness advantages and typically necessitate additional mutations for the development of specific phenotypes (57, 60, 97). This occurrence is exemplified in the rho R109H mutant as its mutant state is likely not actualized unless the cellular environment is altered through the elimination of YdcI. Within rho R109H/ΔydcI cells, the elevated pHi would likely result in a more uniform activity state for Rho since the majority of histidine residues would be deprotonated, effectively genetically assimilating the plastic rho R109H allele in its beneficial mutant state (98). Physiologically, the alkalinized pHi of rho R109H/ΔydcI cells would help alleviate stress associated with maintaining pH homeostasis in an alkalized environment. In these conditions, pH homeostasis is dependent on the activity of electrogenic antiporters, such as the NhaA Na+/H+ pump, to power proton entry, placing a large energetic burden on maintaining a proton motive force (24). Upstream of nhaA lies sokC, which contains a Rho-dependent terminator (47, 99). Thus, it is possible that when Rho activity is compromised, transcriptional readthrough from sokC leads to increased expression of nhaA. Moreover, by maintaining a pHi that is more alkaline than the environment, rho R109H/ΔydcI cells would be less dependent on the activity of cation/proton antiporters to maintain homeostasis. This physiological benefit extends to the pH fluctuation period during resource replenishment when there is an abrupt downshift to a neutral environmental pH. In this condition, our data indicate that rho R109H/ΔydcI cells would remain highly alkaline, leading to an even larger difference in protons across the membrane that would favor proton entry. Given that many metabolic substrates are often cotransported with H+ via symporters, the increased ΔpH of rho R109H/ΔydcI cells would help to drive the import of resources (100). Thus, the alkalinization of rho R109H/ΔydcI cells would not only help to reduce the energy required to maintain pH homeostasis but would also enable the ability to better capitalize on resources through more efficient transport of the available resources.
In summary, these data report a pH-sensitive mutation in an essential global regulator that confers significant physiological effects when paired with the inactivation of a mostly unknown transcription factor related to pH homeostasis. These mutations primarily evolve in the first 300 d of experimental evolution, illustrating how phenotypically plastic traits can facilitate rapid adaptation to complex, fluctuating environments. Despite the absence of this mutation from the current literature, we identify several natural isolates from fluctuating neutral/alkaline conditions carrying this allele, illustrating the power of experimental evolution for identifying functionally important mutations relevant to natural environments. Finally, the phenomenon described herein may signify genetic preassimilation via permissive mutations as a widespread adaptive mechanism for challenging complex environments. An emerging hallmark of cancer involves perturbations in pHi that manifest similarly to rho R109H/ΔydcI cells, where the pHi is maintained higher than that of the extracellular environment (101, 102). In cancers, these alterations can be induced through arginine to histidine mutations that confer pH-sensing capabilities, particularly in global regulators, subsequently altering gene expression and reshaping the cellular environment (40, 41). This suggests particular amino acid mutational signatures could yield effects applicable across all domains of life and that a greater understanding of how these patterns could contribute to changes in cellular physiology is of paramount importance.

Materials and Methods

Strain Construction and Growth Conditions.

All bacterial strains originate from PFM2, a prototrophic derivative of E. coli MG1655 K-12 that was provided by Dr. Patricia Foster (Indiana University) (103). The rho R109H substitution was cloned into the PFM2 ΔaraBAD background using the Church protocol (104). The ΔydcI and rho R109H/ΔydcI mutants were generated by moving the ydcI723(del)::kan cassette from the Keio collection strain JW5226 (105) to the PFM2 WT and rho R109H strains, respectively, using P1vir transduction (106). Mutations were confirmed via PCR and Sanger sequencing. The rho R109H mutation was confirmed using a forward primer sequence of 5′-GGTGATGGCGTACTGGAGAT-3′ and a reverse primer sequence of 5′-GTGCCACAATCAGACCACG-3′. For the ydcI locus, mutations were confirmed using the forward primer 5′-CGGCACGATTAACTGAGGCCGT-3′ and reverse primer 5′-CGCCGTGATGCTGTTCGCCAA-3′.

Identification of Mutations in Time-Series Whole-Population Metagenomic Sequencing.

Mutations in rho and ydcI were initially detected during the analysis of longitudinal metagenomic sequencing of 16 E. coli populations that were experimentally evolved in 100-d feast/famine cycle conditions (64). These populations originate from four different genetic backgrounds: PFM2; PFM2, ΔaraBAD; PFM2, ΔmutL; or PFM2, ΔaraBAD, ΔmutL. The deletion of the ΔaraBAD locus serves as a neutral marker that enables the detection of cross-contamination between lines during experimental evolution by plating on TA agar (10 g/L tryptone, 1 g/L yeast extract, 5 g/L NaCl, 16 g/L agar, 10 g/L L-arabinose, and 0.005% tetrazolium chloride) (107), while the deletion of ΔmutL disrupts methyl-directed mismatch repair and allows for the examination of the evolutionary process in a mutator phenotype. The relevant genotypes of populations adapted to 100-d feast/famine cycles can be found in SI Appendix, Table S1.

Structural Analysis.

The quaternary structure of YdcI was predicted with SWISS-MODEL from an alpha-fold predicted structure for YdcI (P77171) on UniprotKB. As YdcI is annotated as a LysR-family transcriptional regulator, which are typically homotetramers (108, 109), we focused our search on the homology options in the SWISS-MODEL repository that were homotetramers and based on OxyR—another member of the LysR family (templates: 4 x 6g.1.A and 6g1d.1.A). Residues that acquired nonsynonymous mutations throughout experimental evolution were highlighted in PyMol. For the structure of Rho, the most recent and most detailed cryoEM reconstruction of an E. coli Rho–RNA complex was used (#8E6W), while PyMol was used to highlight features of interest.

In Vitro Analysis of Rho Activity.

Plasmid pET28bR109H for overexpression of the R109H mutant was obtained by site-directed mutagenesis of plasmid pET28bRho (kindly provided by Pr. James Berger, Johns Hopkins University) using the NEBuilder HiFi kit (New England Biolabs). The R109H mutant was overexpressed in BL21(DE3)pLysS cells carrying the pET28bR109H plasmid and purified as described for WT Rho (110).
Duplex unwinding experiments were performed as described (110) with minor modifications. Briefly, 5 nM RNA:DNA duplexes bearing a 32P label and a rut site [substrate C (111)] was incubated for 5 min at 37 °C with 20 nM Rho hexamers in helicase buffer (150 mM potassium acetate and 20 mM of MES, HEPES, or EPPS at the indicated pH). Then, 1 mM MgCl2, 1 mM ATP, and 400 nM Trap oligonucleotide (complementary to the DNA strand of the duplexes) were added to the reaction mixture before incubation at 30 °C. Reaction aliquots were withdrawn at various times, quenched with two volumes of quench buffer (20 mM EDTA, 0.5% SDS, 150 mM sodium acetate, and 4% Ficoll 400), and analyzed by 9% polyacrylamide gel electrophoresis (PAGE) and phosphorimaging (Typhoon FLA-9500 instrument and ImageQuant TL v8.1 software, Cytiva).
Transcription termination experiments were performed as described (112) with minor modifications. Briefly, DNA template encoding the λtR1 terminator (0.1 pmol), E. coli RNA polymerase (0.4 pmol; New England Biolabs), Rho (1.4 pmol), and SUPERase-In (0.5 U/µL; Ambion) were mixed in 18 µL of transcription buffer (150 mM potassium acetate, 5 mM MgCl2, 1 mM DTT, 0.05 mg/mL BSA, and either 20 mM HEPES, pH 7.0, or EPPS, pH 9.0) and incubated for 10 min at 37 °C. Then, 2 µL of initiation mix (2 mM ATP, GTP, and CTP, 0.2 mM UTP, 2.5 µCi/µL of 32P-aUTP, and 250 µg/mL rifampicin) were added to the mixtures before further incubation for 20 min at 37 °C. Reactions were quenched with 70 mM EDTA, 0.1 mg/mL tRNA, and 0.3 M sodium acetate, phenol extracted, concentrated by ethanol precipitation, and then analyzed by denaturing 8% PAGE and phosphorimaging.
Rho–RNA dissociation constants (Kd) were determined using a filter-binding assay, as described (113). Briefly, ~10 fmoles of 32P-labeled RNA substrate (RNA strand from substrate C) were mixed with various amounts of Rho in 100 µL of binding buffer (0.1 mM EDTA, 1 mM DTT, 150 mM potassium acetate, 20 µg/mL BSA, and either 20 mM HEPES, pH 7.0, or 20 mM EPPS, pH 9.0). After incubation for 10 min at 37 °C, the samples were filtered through stacked [top] nitrocellulose (Amersham Protran) and [bottom] cationic nylon (Pall Biodyne B) membranes using a Bio-dot SF apparatus (Bio-Rad). The fractions of free and Rho-bound RNA (retained on the nylon and nitrocellulose membranes, respectively) as a function of Rho concentration (expressed in hexamers) were then determined by phosphorimaging of the membranes.

Quantification of Intracellular pH.

Intracellular pH was determined using the pH-sensitive fluorescent dye BCECF-AM. In brief, the fluorescence of this dye depends on the environmental pH, with high pH environments showing increased levels of fluorescence. It is a ratiometric dye, meaning that we utilize the ratio of fluorescence emission (525 nm) when excited at two different wavelengths (470 nm and 440 nm). Because the relative intensities in these two channels depend on both the imaging conditions (light intensity, emission filters, exposure time, etc.) and the pH, the conversion from the ratio of intensities to pH units needs to be calibrated utilizing the same approach as the experimental collection. Here, we used the ionophore nigericin as a way to collapse the pH gradient between the inside of control cells and their external buffered environment. This allowed us to measure intensities ratios and build a conversion function from intensities to pH while the BCECF dye was still in the presence of cellular components.
Cultures of each experimental strain, including an additional WT culture used for calibration of ratiometric fluorescence calibration (control), were grown overnight in LB broth. Due to the significant autofluorescence exhibited by LB broth, it could not be utilized for this microscopy-based approach. To address this issue, we thoroughly washed and resuspended cells from the overnight cultures in Ringer’s buffer, which is specifically formulated to minimize osmotic stress and commonly used to wash cells prior to staining and visualization on a microscope (114, 115). Specifically, a 1 mL aliquot from each culture was washed two times via centrifugation and resuspended in 500 µL of unbuffered Ringer’s buffer (1/4X) (Oxoid). We stained cells for 30 min at 37 °C with 2 µM BCECF-AM (Invitrogen). In addition to BCECF-AM staining, we added 5 µL of nigericin (Sigma Aldrich) to the tube containing the WT ladder standard to allow intracellular pH to equilibrate with the external environment. Following staining, we performed two additional washes via centrifugation using 500 µL of unbuffered 1/4X Ringers’ solution (Oxoid). 100 µL aliquots of each experimental strain and the ladder sample were transferred into five separate tubes. We centrifuged each tube for 1 min at 13,000×g and resuspended the pellets in one of seven 1/4X buffered Ringers’ solutions (pH 6.92, 7.25, 7.72, 8.33, 8.69, 9.06, and 9.59). We prepared buffered Ringers’ solutions using 50 mM of 1,3-Bis[tris(hydroxymethyl)amino]propane (Acros Organics) and adjusted the pH to the appropriate level using 1 M HCl. Once resuspended into pH-adjusted solutions, we imaged cells after approximately 30 min.
For each condition, 7 µL of cells was sandwiched between a 22 mm square #1.5 coverslip and a standard glass slide, providing sufficient fluid volume for cells to remain in solution but in approximately the same focal plane of the microscope. Cells were imaged on a Nikon Ti inverted microscope (Nikon Instruments, Melville, NY) equipped with an Andor Zyla sCMOS camera (Oxford Instruments, Abingdon, Oxfordshire), Spectra/Aura light engine (Lumencor, Beaverton, OR), and an Apo TIRF 60X Oil DIC N2 NA 1.49 objective (Nikon Instruments, Melville, NY). The BCECF fluorescence intensity was imaged in two channels (1: Ex 470 nm, Em 525/50 nm, and 2: Ex 440 nm, Em 525/50 nm). We used a semiautomated MATLAB script (MathWorks, Natwick, MA) to segment cells and quantify the fluorescence intensity. In brief, the process of segmentation identifies which pixels in the image correspond to individual objects or regions. Within each of these regions, the total intensity from each fluorescence channel was computed. Regions around these cells were used as proxies for the local background intensity and subtracted from the total intensity in each region. From these background subtracted intensities, the intensity ratio was calculated. To assist in the automation of this process, we utilized standard approaches in image processing which are described below. Images underwent a spatial bandpass filter (spatial frequencies between 220 nm and 1,100 nm) to remove salt and pepper noise and background intensity. Objects were then segmented as individual regions if above a certain intensity threshold. Because cells were plated at a sufficiently low density to ensure single objects, a relative threshold for each image was set as the intensity such that approximately the top 1,000 pixel values were above the threshold (quantile of 0.9998). To ensure that the entire region of each cell was included, and to deal with the fact that these two imaging channels have a small spatial shift from one channel to the other, these regions were then dilated by disks of radius 2 µm. Background regions were defined by dilating these regions by a further 2 µm, and removing the objects from these masks, yielding regions that are near, but not overlapping with, individual objects. The intensity for each object in each channel was determined as the background-subtracted sum of all the pixels in each region. Lists of relative intensities of these objects were then imported into RStudio. Samples treated with nigericin were used to determine the coefficients A and B in a fit of pHobs = A*log(intensityRatio)+B.
To statistically determine differences in intracellular pH across the WT and engineered mutant strains, we fit the relationship between intracellular pH and environmental pH to a logarithmic function [Intracellular pH ~ a*log(Environmental pH)+k] using a nonlinear least squares method in the R package “nlm.” The fit was calculated across the measured interval of pH 7 to 9.5, and CI were calculated via uncertainty propagation with the function predictNLS which calculates CI for the fitted values of nonlinear models by using first-/second-order Taylor expansion and Monte Carlo simulation.

Replaying of RLTS Regime.

To determine the prevalence of the rho R109H and ΔydcI alleles in the RLTS condition and to parse out any patterns in their co-occurrence, we replayed one cycle of the RLTS regime by incubating WT, rho R109H, and ΔydcI cultures for 100 d after which time we examined the rho and ydcI loci for acquired mutations. We inoculated 12 cultures of WT, rho R109H, and ΔydcI strains in 10 mL of LB broth each originating from a separate colony from an LB agar plate freshly prepared from frozen glycerol stocks. We cultivated each culture at 37 °C and shaken upright at 180 rpm for a total of 100 d. We left cultures undisturbed during this time except for a brief period every 10 d when the culture tubes were gently removed from the incubator and added sterile deionized water to achieve a volume of 10 mL to counteract evaporation. After 100 d, we removed the cultures and centrifuged them 2,500×g for 10 min. We froze the resulting pellet of bacterial biomass in liquid nitrogen and it was stored at −80 °C until DNA extraction was performed. We extracted DNA from the pellets using a DNeasy UltraClean Microbial Kit (Qiagen) and used the resulting DNA samples as template for PCR. We performed PCR using the same primers previously mentioned that were used for confirmation of the initial mutation. The PCR products were purified using the Monarch PCR & DNA Cleanup Kit (NEB) and sent for Sanger sequencing (GENEWIZ).

Pairwise Competition Assays.

To determine the fitness outcomes of the mutant strains relative to the WT ancestral strain, we performed pairwise competition assays (107, 116). Here, culture tubes containing a 50:50 starting proportion of two competing strains are grown for a given period, at which time they are sampled to determine the resulting proportions of each strain following growth. Each competing strain contains either an intact araBAD operon or ΔaraBAD genotype and can be differentiated from each other by plating on TA agar (10 g/L tryptone, 1 g/L yeast extract, 5 g/L NaCl, 16 g/L agar, 10 g/L L-arabinose, 0.005% tetrazolium chloride) where the WT araBAD strain produces light pink colonies and the ΔaraBAD strain produces bright red colonies. Overnight cultures were prepared for each strain using colonies picked from freshly streaked LB agar plates obtained from frozen stocks. For each competition, we inoculated 10 mL of LB broth with 50 µL of overnight culture from both competitors. We set up individual competitions for each sampling time point, as vortexing and resampling of the same competition tube has been shown to alter fitness outcomes (23). Before incubation, we sampled each competition to determine precise starting proportions by collecting a 100 µL aliquot, serially diluting in PBS, and plating on TA agar. All competition tubes were incubated at 37 °C until the sampling time point at which time a 100 µL sample was removed, serially diluted in PBS, and plated on TA agar. All TA agar plates were incubated at 37 °C for 24 h before colonies were enumerated. For competitions using buffered media, we added 50 mM 1,3-Bis[tris(hydroxymethyl)amino]propane (Acros Organics) and adjusted the pH to either 7.0 or 9.0 using 1 M HCl.

Survival Assays.

We assessed the survival of our strains by monitoring their population density over the course of 14 d. We inoculated three replicate overnight cultures of WT, rho R109H, ΔydcI, and rho R109H/ΔydcI strains in 10 mL of LB broth each originating from a separate colony from an LB agar plate freshly prepared from frozen glycerol stocks and incubated them for 18 h at 37 °C. The survival assay was initiated by aliquoting 100 µL of each overnight culture into 14 individual culture tubes filled with 10 mL of fresh LB broth. We incubated each culture at 37 °C, shaken upright at 180 rpm until the appropriate sampling time point. We set up individual cultures for each sampling time point, as vortexing and resampling of the same culture tube has been shown to alter fitness outcomes (23). Every 24 h, the culture tube corresponding to the sampling time point was removed from the incubator and gently vortexed. We then removed 100 µL from the culture and performed serial dilutions in PBS. We plated 100 µL aliquots of the dilutions on LB agar plates and enumerated the colonies from countable plates after 24 h of incubation at 37 °C.

Identification of Other Bacterial rho R109H Alleles.

Rho homologs containing a histidine at the analogous E. coli R109 residue were identified using blastp with the Pattern Hit Initiated (PHI-BLAST) algorithm (117). The following PHI pattern EH[YF][YFG][AGS][LM][LVT] targets R109H substitutions to the conserved ERFYALL motif from the E. coli Rho sequence (NP_418230.1) while allowing for any variation observed after alignment to Rho sequences with COBALT (118) across nine additional species of bacteria spanning the bacterial tree of life: Bacillus subtilis (AIY95021.1); Borreliella burgdorferi (WP_002657659.1); Deinococcus radiodurans (AAF10910.1); Helicobacter pylori (EIE30697.1); Pseudomonas aeruginosa (NP_253926.1); Salmonella enterica typhimurium (NP_462807.1); Staphylococcus aureus (YP_500838.1); Streptococcus pneumoniae (CVN06062.1); and Vibrio cholerae (WP_001054524.1) (Fig. 5).

Data, Materials, and Software Availability

All raw data and code needed to reproduce the figures in this study are available at https://github.com/BehringerLab/Rho_YdcI_pH. All other data are included in the manuscript and/or SI Appendix. Previously published data were used for this work (18).

Acknowledgments

We would like to thank the Advanced Computing Center for Research and Education at Vanderbilt and Vanderbilt Cell Imaging Shared Resource (supported by NIH Grants CA68485, DK20593, DK58404, DK59637, and EY08126), specifically the Nikon Multi-Excitation TIRF Microscope acquired under S10-OD018075-01A1. This work was supported by Army Research Office Grant W911NF-21-1-0161 (M.G.B.), NIH Grant R35GM150625 (M.G.B.), and French Agence Nationale de la Recherche Grant ANR-19-CE44-0009-01 (M.B.) Additional funds were provided by Vanderbilt University (startup funds: M.G.B.), Vanderbilt University Medical Center (startup funds: B.P.B), Evolutionary Studies Initiative at Vanderbilt (M.G.B and B.P.B), and the Vanderbilt University Medical Center Discovery Scholars in Health and Medicine Program (B.P.B).

Author contributions

S.B.W., B.P.B., M.B., and M.G.B. designed research; S.B.W., R.D.P.M., M.D., R.S., B.P.B., M.B., and M.G.B. performed research; R.S. contributed new reagents/analytic tools; S.B.W., R.D.P.M., M.D., B.P.B., M.B., and M.G.B. analyzed data; and S.B.W., R.D.P.M., B.P.B., M.B., and M.G.B. wrote the paper.

Competing interests

The authors declare no competing interest.

Supporting Information

Appendix 01 (PDF)

References

1
A. Aertsen, C. W. Michiels, Stress and how bacteria cope with death and survival. Crit. Rev. Microbiol. 30, 263–273 (2004).
2
G. Lambert, E. Kussell, Memory and fitness optimization of bacteria under fluctuating environments. PLoS Genet. 10, e1004556 (2014).
3
J. Nguyen, J. Lara-Gutiérrez, R. Stocker, Environmental fluctuations and their effects on microbial communities, populations and individuals. FEMS Microbiol. Rev. 45, fuaa068 (2021).
4
R. Y. Morita, Bioavailability of energy and its relationship to growth and starvation survival in nature. Can. J. Microbiol. 34, 436–441 (1988).
5
A. T. Reese et al., Microbial nitrogen limitation in the mammalian large intestine. Nat. Microbiol. 3, 1441–1450 (2018).
6
J. Hobbie, E. Hobbie, Microbes in nature are limited by carbon and energy: The starving-survival lifestyle in soil and consequences for estimating microbial rates. Front. Microbiol. 4, 324 (2013).
7
R. Kolter, D. A. Siegele, A. Tormo, The stationary phase of the bacterial life cycle. Annu. Rev. Microbiol. 47, 855–874 (1993).
8
M. M. Zambrano, D. A. Siegele, M. Almirón, A. Tormo, R. Kolter, Microbial competition: Escherichia coli mutants that take over stationary phase cultures. Science 259, 1757–1760 (1993).
9
M. M. Zambrano, R. Kolter, GASPing for life in stationary phase. Cell 86, 181–184 (1996).
10
S. E. Finkel, Long-term survival during stationary phase: Evolution and the GASP phenotype. Nat Rev Microbiol 4, 113–120 (2006).
11
S. Katz et al., Dynamics of adaptation during three years of evolution under long-term stationary phase. Mol. Biol. Evol. 38, 2778–2790 (2021).
12
N. R. Ratib, F. Seidl, I. M. Ehrenreich, S. E. Finkel, Evolution in long-term stationary-phase batch culture: Emergence of divergent Escherichia coli lineages over 1,200 days. mBio 12, e03337-20 (2021).
13
W. R. Shoemaker et al., Microbial population dynamics and evolutionary outcomes under extreme energy limitation. Proc. Natl. Acad. Sci. U.S.A. 118, e2101691118 (2021).
14
T. Aouizerat et al., Eukaryotic adaptation to years-long starvation resembles that of bacteria. iScience 19, 545–558 (2019).
15
M. G. Behringer et al., Escherichia coli cultures maintain stable subpopulation structure during long-term evolution. Proc. Natl. Acad. Sci. U.S.A. 115, E4642–E4650 (2018).
16
M. G. Behringer et al., Complex ecotype dynamics evolve in response to fluctuating resources. mBio 13, e03467-21 (2022).
17
K. E. Kram et al., Adaptation of Escherichia coli to long-term serial passage in complex medium: Evidence of parallel evolution. mSystems 2 (2017), https://doi.org/10.1128/msystems.00192-16.
18
M. G. Behringer et al., Trade-offs, trade-ups, and high mutational parallelism underlie microbial adaptation during extreme cycles of feast and famine. Curr. Biol. 34, 1403–1413.e5 (2024).
19
C. Ratzke, J. Gore, Modifying and reacting to the environmental pH can drive bacterial interactions. PLoS Biol. 16, e2004248 (2018).
20
M. J. Farrell, S. E. Finkel, The growth advantage in stationary-phase phenotype conferred by rpoS mutations is dependent on the pH and nutrient environment. J. Bacteriol. 185, 7044–7052 (2003).
21
G. Sezonov, D. Joseleau-Petit, R. D’Ari, Escherichia coli physiology in Luria–Bertani broth. J. Bacteriol. 189, 8746–8749 (2007).
22
R. Sánchez-Clemente, M. I. Guijo, J. Nogales, R. Blasco, Carbon source influence on extracellular pH changes along bacterial cell-growth. Genes 11, 1292 (2020).
23
S. B. Worthan, R. D. P. McCarthy, M. G. Behringer, Case studies in the assessment of microbial fitness: Seemingly subtle changes can have major effects on phenotypic outcomes. J. Mol. Evol. 91, 311–324 (2023).
24
T. A. Krulwich, G. Sachs, E. Padan, Molecular aspects of bacterial pH sensing and homeostasis. Nat. Rev. Microbiol. 9, 330–343 (2011).
25
V. C. Santos, R. N. Araujo, L. A. D. Machado, M. H. Pereira, N. F. Gontijo, The physiology of the midgut of Lutzomyia longipalpis (Lutz and Neiva 1912): pH in different physiological conditions and mechanisms involved in its control. J. Exp. Biol. 211, 2792–2798 (2008).
26
R. Lande, Evolution of phenotypic plasticity and environmental tolerance of a labile quantitative character in a fluctuating environment. J. Evol. Biol. 27, 866–875 (2014).
27
P. A. O’Brien, K. M. Morrow, B. L. Willis, D. G. Bourne, Implications of Ocean acidification for marine microorganisms from the free-living to the host-associated. Front. Mar. Sci. 3, 47 (2016).
28
X. Yi, A. M. Dean, Phenotypic plasticity as an adaptation to a functional trade-off. eLife 5, e19307 (2016).
29
M.-P. Castanié-Cornet et al., Acid stress response in Escherichia coli: Mechanism of regulation of gadA transcription by RcsB and GadE. Nucleic Acids Res. 38, 3546–3554 (2010).
30
P. A. Hoskisson, M. I. Hutchings, MtrAB–LpqB: A conserved three-component system in actinobacteria? Trend. Microbiol. 14, 444–449 (2006).
31
I. Haneburger, A. Eichinger, A. Skerra, K. Jung, New insights into the signaling mechanism of the pH-responsive, membrane-integrated transcriptional activator CadC of Escherichia coli. J. Biol. Chem. 286, 10681–10689 (2011).
32
A. Battesti, N. Majdalani, S. Gottesman, The RpoS-mediated general stress response in Escherichia coli. Annu. Rev. Microbiol. 65, 189–213 (2011).
33
I. Hamdallah et al., Experimental evolution of Escherichia coli K-12 at high pH and with RpoS induction. Appl. Environ. Microbiol. 84, e00520-18 (2018).
34
J. Rebek, On the structure of histidine and its role in enzyme active sites. Struct. Chem. 1, 129–131 (1990).
35
O. Rötzschke, J. M. Lau, M. Hofstätter, K. Falk, J. L. Strominger, A pH-sensitive histidine residue as control element for ligand release from HLA-DR molecules. Proc. Natl. Acad. Sci. U.S.A. 99, 16946–16950 (2002).
36
D. M. Williamson, J. Elferich, U. Shinde, Mechanism of fine-tuning pH sensors in proprotein convertases: Identification of a pH-sensing histidine pair in the propeptide of proprotein convertase 1/3. J. Biol. Chem. 290, 23214–23225 (2015).
37
Y. Vercoulen et al., A Histidine pH sensor regulates activation of the Ras-specific guanine nucleotide exchange factor RasGRP1. eLife 6, e29002 (2017).
38
Y. Gerchman et al., Histidine-226 is part of the pH sensor of NhaA, a Na+/H+ antiporter in Escherichia coli. Proc. Natl. Acad. Sci. U.S.A. 90, 1212–1216 (1993).
39
B. A. Webb et al., A histidine cluster in the cytoplasmic domain of the Na-H exchanger NHE1 confers pH-sensitive phospholipid binding and regulates transporter activity. JBC 291, 24096–24104 (2016).
40
Z. A. Szpiech et al., Prominent features of the amino acid mutation landscape in cancer. PLoS ONE 12, e0183273 (2017).
41
K. A. White et al., Cancer-associated arginine-to-histidine mutations confer a gain in pH sensing to mutant proteins. Sci. Signal. 10, eaam9931 (2017).
42
S. F. Elena, R. E. Lenski, Evolution experiments with microorganisms: The dynamics and genetic bases of adaptation. Nat. Rev. Genet. 4, 457–469 (2003).
43
M. J. McDonald, Microbial experimental evolution—A proving ground for evolutionary theory and a tool for discovery. EMBO Rep. 20, e46992 (2019).
44
G. Saxer et al., Mutations in global regulators lead to metabolic selection during adaptation to complex environments. PLoS Genet. 10, e1004872 (2014).
45
D. T. Fraebel, K. Gowda, M. Mani, S. Kuehn, Evolution of generalists by phenotypic plasticity. iScience 23, 101678 (2020).
46
C. J. Cardinale et al., Termination factor rho and its cofactors NusA and NusG silence foreign DNA in E. coli. Science 320, 935–938 (2008).
47
J. M. Peters et al., Rho and NusG suppress pervasive antisense transcription in Escherichia coli. Genes Dev. 26, 2621–2633 (2012).
48
N. Sedlyarova et al., sRNA-mediated control of transcription termination in E. coli. Cell 167, 111–121.e13 (2016).
49
M. Delaleau, E. Eveno, I. Simon, A. Schwartz, M. Boudvillain, A scalable framework for the discovery of functional helicase substrates and helicase-driven regulatory switches. Proc. Natl. Acad. Sci. U.S.A. 119, e2209608119 (2022).
50
R. S. Washburn, M. E. Gottesman, Transcription termination maintains chromosome integrity. Proc. Natl. Acad. Sci. U.S.A. 108, 792–797 (2011).
51
M. A. Kriner, A. Sevostyanova, E. A. Groisman, Learning from the leaders: Gene regulation by the transcription termination factor rho. TIBS 41, 690–699 (2016).
52
L. Bossi, N. Figueroa-Bossi, P. Bouloc, M. Boudvillain, Regulatory interplay between small RNAs and transcription termination factor Rho. BBA-Gene Regul. Mech. 1863, 194546 (2020).
53
E. J. Steinmetz, T. Platt, Evidence supporting a tethered tracking model for helicase activity of Escherichia coli Rho factor. Proc. Natl. Acad. Sci. U.S.A. 91, 1401–1405 (1994).
54
V. Molodtsov, C. Wang, E. Firlar, J. T. Kaelber, R. H. Ebright, Structural basis of Rho-dependent transcription termination. Nature 614, 367–374 (2023).
55
T. Kishimoto et al., Transition from positive to neutral in mutation fixation along with continuing rising fitness in thermal adaptive evolution. PLoS Genet. 6, e1001164 (2010).
56
O. Tenaillon et al., The molecular diversity of adaptive convergence. Science 335, 457–461 (2012).
57
A. Rodríguez-Verdugo, D. Carrillo-Cisneros, A. González-González, B. S. Gaut, A. F. Bennett, Different tradeoffs result from alternate genetic adaptations to a common environment. Proc. Natl. Acad. Sci. U.S.A. 111, 12121–12126 (2014).
58
D. E. Deatherage, J. L. Kepner, A. F. Bennett, R. E. Lenski, J. E. Barrick, Specificity of genome evolution in experimental populations of Escherichia coli evolved at different temperatures. Proc. Natl. Acad. Sci. U.S.A. 114, E1904–E1912 (2017).
59
S. M. Boyd et al., Experimental evolution of copper resistance in Escherichia coli produces evolutionary trade-offs in the antibiotics chloramphenicol, bacitracin, and sulfonamide. Antibiotics 11, 711 (2022).
60
P. L. Freddolino, H. Goodarzi, S. Tavazoie, Fitness landscape transformation through a single amino acid change in the rho terminator. PLoS Genet. 8, e1002744 (2012).
61
R. J. F. Haft et al., Correcting direct effects of ethanol on translation and transcription machinery confers ethanol tolerance in bacteria. Proc. Natl. Acad. Sci. U.S.A. 111, E2576–E2585 (2014).
62
T. M. Conrad et al., Whole-genome resequencing of Escherichia coli K-12 MG1655 undergoing short-term laboratory evolution in lactate minimal media reveals flexible selection of adaptive mutations. Genome Biol. 10, R118 (2009).
63
M. D. Herron, M. Doebeli, Parallel evolutionary dynamics of adaptive diversification in Escherichia coli. PLoS Biol. 11, e1001490 (2013).
64
M. G. Behringer et al., Trade-offs, trade-ups, and high mutational parallelism underlie microbial adaptation to extreme feast/famine. Curr. Biol. 34, 1403–1413.e5 (2024). https://doi.org/10.1101/2023.10.04.560893.
65
Y. Gao et al., Systematic discovery of uncharacterized transcription factors in Escherichia coli K-12 MG1655. Nucleic Acids Res. 46, 10682–10696 (2018).
66
A. J. Wolfe, The acetate switch. Microbiol. Mol. Biol. Rev. 69, 12–50 (2005).
67
S. Cesar et al., Bacterial evolution in high-osmolarity environments. mBio 11, e01191-20 (2020).
68
P. Mitra, G. Ghosh, M. Hafeezunnisa, R. Sen, Rho protein: Roles and mechanisms. Annu. Rev. Microbiol. 71, 687–709 (2017).
69
M. Damaghi, J. W. Wojtkowiak, R. J. Gillies, pH sensing and regulation in cancer. Front. Physiol. 4, 370 (2013).
70
E. L. DiGiammarino et al., A novel mechanism of tumorigenesis involving pH-dependent destabilization of a mutant p53 tetramer. Nat. Struct. Mol. Biol. 9, 12–16 (2002).
71
C. Frantz et al., Cofilin is a pH sensor for actin free barbed end formation: Role of phosphoinositide binding. JCB 183, 865–879 (2008).
72
C.-H. Choi, B. A. Webb, M. S. Chimenti, M. P. Jacobson, D. L. Barber, pH sensing by FAK-His58 regulates focal adhesion remodeling. JCB 202, 849–859 (2013).
73
V. Romiyo, J. W. Wilson, Phenotypes, transcriptome, and novel biofilm formation associated with the ydcI gene. Antonie van Leeuwenhoek 113, 1109–1122 (2020).
74
S. E. Maddocks, P. C. F. Oyston, Structure and function of the LysR-type transcriptional regulator (LTTR) family proteins. Microbiology 154, 3609–3623 (2008).
75
R. Shashni, M. Z. Qayyum, V. Vishalini, D. Dey, R. Sen, Redundancy of primary RNA-binding functions of the bacterial transcription terminator Rho. Nucleic Acids Res. 42, 9677–9690 (2014).
76
K. Tsujimoto, M. Semadeni, M. Huflejt, L. Packer, Intracellular pH of halobacteria can be determined by the fluorescent dye 2’,7’-bis(carboxyethyl)-5(6)-carboxyfluorescein. BBRC 155, 123–129 (1988).
77
D. Molenaar, H. Bolhuis, T. Abee, B. Poolman, W. N. Konings, The efflux of a fluorescent probe is catalyzed by an ATP-driven extrusion system in Lactococcus lactis. J. Bacteriol. 174, 3118–3124 (1992).
78
C. Riondet, R. Cachon, Y. Waché, G. Alcaraz, C. Diviès, Measurement of the intracellular pH in Escherichia coli with the internally conjugated fluorescent probe 5- (and 6-)carboxyfluorescein succinimidyl ester. Biotechnol. Tech. 11, 735–738 (1997).
79
J. Heo, K. J. Thomas, G. H. Seong, R. M. Crooks, A microfluidic bioreactor based on hydrogel-entrapped E. coli: Cell viability, lysis, and intracellular enzyme reactions. Anal. Chem. 75, 22–26 (2003).
80
E. R. Zinser, R. Kolter, Mutations enhancing amino acid catabolism confer a growth advantage in stationary phase. J. Bacteriol. 181, 5800–5807 (1999).
81
S. Banerjee, J. Chalissery, I. Bandey, R. Sen, Rho-dependent transcription termination: More questions than answers. J. Microbiol. 44, 11–22 (2006).
82
S. Festa, B. M. Coppotelli, I. S. Morelli, Bacterial diversity and functional interactions between bacterial strains from a phenanthrene-degrading consortium obtained from a chronically contaminated-soil. Int. Biodeterior. Biodegrad. 85, 42–51 (2013).
83
P. Ileana et al., “Soil quality problems associated with horticulture in the southern urban and peri-urban area of Buenos Aires, Argentina” in Urban Horticulture—Necessity of the Future (IntechOpen, 2020), https://doi.org/10.5772/intechopen.90351.
84
R. Ruvindy, R. A. White III, B. A. Neilan, B. P. Burns, Unravelling core microbial metabolisms in the hypersaline microbial mats of Shark Bay using high-throughput metagenomics. ISME J. 10, 183–196 (2016).
85
R. A. White, H. L. Wong, R. Ruvindy, B. A. Neilan, B. P. Burns, Viral communities of Shark Bay modern stromatolites. Front. Microbiol. 9, 1223 (2018).
86
A. Dostálová, P. Volf, Leishmania development in sand flies: Parasite–vector interactions overview. Parasites Vectors 5, 276 (2012).
87
D. R. Speth et al., Microbial communities of Auka hydrothermal sediments shed light on vent biogeography and the evolutionary history of thermophily. ISME J. 16, 1750–1764 (2022).
88
K. Hama, T. Kunimaru, R. Metcalfe, A. J. Martin, The hydrogeochemistry of argillaceous rock formations at the Horonobe URL site, Japan. Phys. Chem. Earth, Parts A/B/C 32, 170–180 (2007).
89
A. W. Hernsdorf et al., Potential for microbial H2 and metal transformations associated with novel bacteria and archaea in deep terrestrial subsurface sediments. ISME J. 11, 1915–1929 (2017).
90
J. M. Peters et al., Rho directs widespread termination of intragenic and stable RNA transcription. Proc. Natl. Acad. Sci. U.S.A. 106, 15406–15411 (2009).
91
E. T. Mohamed et al., Generation of a platform strain for ionic liquid tolerance using adaptive laboratory evolution. Microbial. Cell Fact. 16, 204 (2017).
92
B. Du et al., Adaptive laboratory evolution of Escherichia coli under acid stress. Microbiology (Reading, MA) 166, 141–148 (2020).
93
Md. Hafeezunnisa, R. Sen, The rho-dependent transcription termination is involved in broad-spectrum antibiotic susceptibility in Escherichia coli. Front. Microbiol. 11, 605305 (2020).
94
K. Bhardwaj et al., Rho-dependent termination enables cellular pH homeostasis. J. Bacteriol. 206, e0035623 (2024).
95
L. M. Stancik et al., pH-dependent expression of periplasmic proteins and amino acid catabolism in Escherichia coli. J. Bacteriol. 184, 4246–4258 (2002).
96
T. S. Gunasekera, L. N. Csonka, O. Paliy, Genome-wide transcriptional responses of Escherichia coli K-12 to continuous osmotic and heat stresses. J. Bacteriol. 190, 3712–3720 (2008).
97
S. M. Hug, B. S. Gaut, The phenotypic signature of adaptation to thermal stress in Escherichia coli. BMC Evol. Biol. 15, 177 (2015).
98
C. H. Waddington, Genetic assimilation of an acquired character. Evolution 7, 118–126 (1953).
99
M. Hafeezunnisa, P. I. R. Chhakchhuak, J. Krishnakumar, R. Sen, Rho-dependent transcription termination regulates the toxin–antitoxin modules of cryptic prophages to silence their expression in Escherichia coli. FEBS Lett. 595, 2057–2067 (2021).
100
M. G. L. Elferink, K. J. Hellingwerf, W. N. Konings, The role of the proton motive force and electron flow in solute transport in Escherichia coli. Eur. J. Biochem. 153, 161–165 (1985).
101
M. Sharma et al., pH gradient reversal: An emerging hallmark of cancers. Recent Pat. Anticancer Drug Discov. 10, 244–258 (2015).
102
T. Koltai et al., “Chapter 5—How pH deregulation favors the hallmarks of cancer” in pH Deregulation as the Eleventh Hallmark of Cancer, T. Koltai et al., Eds. (Academic Press, 2023), pp. 101–121, https://doi.org/10.1016/B978-0-443-15461-4.00008-4.
103
H. Lee, E. Popodi, H. Tang, P. L. Foster, Rate and molecular spectrum of spontaneous mutations in the bacterium Escherichia coli as determined by whole-genome sequencing. Proc. Natl. Acad. Sci. U.S.A. 109, E2774–E2783 (2012).
104
J. A. Mosberg, C. J. Gregg, M. J. Lajoie, H. H. Wang, G. M. Church, Improving lambda red genome engineering in Escherichia coli via rational removal of endogenous nucleases. PLoS ONE 7, e44638 (2012).
105
T. Baba et al., Construction of Escherichia coli K-12 in-frame, single-gene knockout mutants: The Keio collection. Mol. Syst. Biol. 2, 2006.0008 (2006).
106
A. Saragliadis, T. Trunk, J. C. Leo, Producing gene deletions in Escherichia coli by P1 transduction with excisable antibiotic resistance cassettes. J. Vis. Exp. 139, 58267 (2018), https://doi.org/10.3791/58267.
107
B. R. Levin, F. M. Stewart, L. Chao, Resource-limited growth, competition, and predation: A model and experimental studies with bacteria and bacteriophage. Am. Nat. 111, 3–24 (1977).
108
M. A. Schell, Molecular biology of the LysR family of transcriptional regulators. Annu. Rev. Microbiol. 47, 597–626 (1993).
109
E.-A. Giannopoulou et al., Crystal structure of the full-length LysR-type transcription regulator CbnR in complex with promoter DNA. FEBS J. 288, 4560–4575 (2021).
110
M. Boudvillain, C. Walmacq, A. Schwartz, F. Jacquinot, “Simple enzymatic assays for the in vitro motor activity of transcription termination factor Rho from Escherichia coli” in Helicases: Methods and Protocols, M. Abdelhaleem, Eds. (Humana Press, Totowa, NJ, 2010), pp. 137–154, https://doi.org/10.1007/978-1-60327-355-8_10.
111
E. Soares, A. Schwartz, M. Nollmann, E. Margeat, M. Boudvillain, The RNA-mediated, asymmetric ring regulatory mechanism of the transcription termination Rho helicase decrypted by time-resolved nucleotide analog interference probing (trNAIP). Nucleic Acids Res. 42, 9270–9284 (2014).
112
C. Nadiras, E. Eveno, A. Schwartz, N. Figueroa-Bossi, M. Boudvillain, A multivariate prediction model for Rho-dependent termination of transcription. Nucleic Acids Res. 46, 8245–8260 (2018).
113
F. D’Heygère, A. Schwartz, F. Coste, B. Castaing, M. Boudvillain, ATP-dependent motor activity of the transcription termination factor Rho from Mycobacterium tuberculosis. Nucleic Acids Res 43, 6099–6111 (2015).
114
H. Z. Low et al., Fast and easy phage-tagging and live/dead analysis for the rapid monitoring of bacteriophage infection. Front. Microbiol. 11, 602444 (2020).
115
B. Roeder, M. Wagner, P. Rossmanith, Autonomous growth of isolated single Listeria monocytogenes and Salmonella enterica serovar typhimurium cells in the absence of growth factors and intercellular contact. Appl. Environ. Microbiol. 76, 2600–2606 (2010).
116
R. E. Lenski et al., Adaptation and divergence during 2,000 generations. Am. Nat. 138, 1315–1341 (1991).
117
Z. Zhang et al., Protein sequence similarity searches using patterns as seeds. Nucleic Acids Res. 26, 3986–3990 (1998).
118
J. S. Papadopoulos, R. Agarwala, COBALT: Constraint-based alignment tool for multiple protein sequences. Bioinformatics 23, 1073–1079 (2007).

Information & Authors

Information

Published in

The cover image for PNAS Vol.121; No.39
Proceedings of the National Academy of Sciences
Vol. 121 | No. 39
September 24, 2024
PubMed: 39298488

Classifications

Data, Materials, and Software Availability

All raw data and code needed to reproduce the figures in this study are available at https://github.com/BehringerLab/Rho_YdcI_pH. All other data are included in the manuscript and/or SI Appendix. Previously published data were used for this work (18).

Submission history

Received: March 17, 2024
Accepted: August 19, 2024
Published online: September 19, 2024
Published in issue: September 24, 2024

Keywords

  1. phenotypic plasticity
  2. experimental evolution
  3. alkaline stress
  4. permissive mutation
  5. transcription termination

Acknowledgments

We would like to thank the Advanced Computing Center for Research and Education at Vanderbilt and Vanderbilt Cell Imaging Shared Resource (supported by NIH Grants CA68485, DK20593, DK58404, DK59637, and EY08126), specifically the Nikon Multi-Excitation TIRF Microscope acquired under S10-OD018075-01A1. This work was supported by Army Research Office Grant W911NF-21-1-0161 (M.G.B.), NIH Grant R35GM150625 (M.G.B.), and French Agence Nationale de la Recherche Grant ANR-19-CE44-0009-01 (M.B.) Additional funds were provided by Vanderbilt University (startup funds: M.G.B.), Vanderbilt University Medical Center (startup funds: B.P.B), Evolutionary Studies Initiative at Vanderbilt (M.G.B and B.P.B), and the Vanderbilt University Medical Center Discovery Scholars in Health and Medicine Program (B.P.B).
Author contributions
S.B.W., B.P.B., M.B., and M.G.B. designed research; S.B.W., R.D.P.M., M.D., R.S., B.P.B., M.B., and M.G.B. performed research; R.S. contributed new reagents/analytic tools; S.B.W., R.D.P.M., M.D., B.P.B., M.B., and M.G.B. analyzed data; and S.B.W., R.D.P.M., B.P.B., M.B., and M.G.B. wrote the paper.
Competing interests
The authors declare no competing interest.

Notes

This article is a PNAS Direct Submission.

Authors

Affiliations

Sarah B. Worthan
Department of Biological Sciences, Vanderbilt University, Nashville, TN 37232
Evolutionary Studies Initiative, Vanderbilt University, Nashville, TN 37232
Vanderbilt Institute for Infection, Immunology and Inflammation, Nashville, TN 37232
Robert D. P. McCarthy
Department of Biological Sciences, Vanderbilt University, Nashville, TN 37232
Centre de Biophysique Moléculaire, CNRS UPR4301, affiliated with Université d’Orléans, Orléans Cedex 2 45071, France
Biodesign Center for Mechanisms of Evolution, Arizona State University, Tempe, AZ 85281
Evolutionary Studies Initiative, Vanderbilt University, Nashville, TN 37232
Vanderbilt Institute for Infection, Immunology and Inflammation, Nashville, TN 37232
Department of Pathology, Microbiology and Immunology, Vanderbilt University Medical Center, Nashville, TN 37232
Department of Cell and Developmental Biology, Vanderbilt University, Nashville, TN 37232
Centre de Biophysique Moléculaire, CNRS UPR4301, affiliated with Université d’Orléans, Orléans Cedex 2 45071, France
Department of Biological Sciences, Vanderbilt University, Nashville, TN 37232
Evolutionary Studies Initiative, Vanderbilt University, Nashville, TN 37232
Vanderbilt Institute for Infection, Immunology and Inflammation, Nashville, TN 37232
Department of Pathology, Microbiology and Immunology, Vanderbilt University Medical Center, Nashville, TN 37232

Notes

1
To whom correspondence may be addressed. Email: [email protected]

Metrics & Citations

Metrics

Note: The article usage is presented with a three- to four-day delay and will update daily once available. Due to ths delay, usage data will not appear immediately following publication. Citation information is sourced from Crossref Cited-by service.


Altmetrics




Citations

Export the article citation data by selecting a format from the list below and clicking Export.

Cited by

    Loading...

    View Options

    View options

    PDF format

    Download this article as a PDF file

    DOWNLOAD PDF

    Login options

    Check if you have access through your login credentials or your institution to get full access on this article.

    Personal login Institutional Login

    Recommend to a librarian

    Recommend PNAS to a Librarian

    Purchase options

    Purchase this article to access the full text.

    Single Article Purchase

    Evolution of pH-sensitive transcription termination in Escherichia coli during adaptation to repeated long-term starvation
    Proceedings of the National Academy of Sciences
    • Vol. 121
    • No. 39

    Figures

    Tables

    Media

    Share

    Share

    Share article link

    Share on social media