Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Implementation of a loss-of-function system to determine growth and stress-associated mutagenesis in Bacillus subtilis

  • Norberto Villegas-Negrete,

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Validation, Visualization, Writing – original draft, Writing – review & editing

    Affiliation Department of Biology, Division of Natural and Exact Sciences, University of Guanajuato, Guanajuato, Mexico

  • Eduardo A. Robleto,

    Roles Conceptualization, Data curation, Funding acquisition, Methodology, Resources, Supervision, Writing – review & editing

    Affiliation School of Life Sciences, University of Nevada, Las Vegas, Nevada, United States of America

  • Armando Obregón-Herrera,

    Roles Data curation, Formal analysis, Validation, Writing – review & editing

    Affiliation Department of Biology, Division of Natural and Exact Sciences, University of Guanajuato, Guanajuato, Mexico

  • Ronald E. Yasbin,

    Roles Data curation, Writing – review & editing

    Affiliation College of Arts and Sciences, University of Missouri—St. Louis, St. Louis, Missouri, United States of America

  • Mario Pedraza-Reyes

    Roles Conceptualization, Data curation, Formal analysis, Funding acquisition, Methodology, Resources, Supervision, Validation, Visualization, Writing – original draft, Writing – review & editing

    pedrama@ugto.mx

    Affiliation Department of Biology, Division of Natural and Exact Sciences, University of Guanajuato, Guanajuato, Mexico

Abstract

A forward mutagenesis system based on the acquisition of mutations that inactivate the thymidylate synthase gene (TMS) and confer a trimethoprim resistant (Tmpr) phenotype was developed and utilized to study transcription-mediated mutagenesis (TMM). In addition to thyA, Bacillus subtilis possesses thyB, whose expression occurs under conditions of cell stress; therefore, we generated a thyB- thyA+ mutant strain. Tmpr colonies of this strain were produced with a spontaneous mutation frequency of ~1.4 × 10−9. Genetic disruption of the canonical mismatch (MMR) and guanine oxidized (GO) repair pathways increased the Tmpr frequency of mutation by ~2–3 orders of magnitude. A wide spectrum of base substitutions as well as insertion and deletions in the ORF of thyA were found to confer a Tmpr phenotype. Stationary-phase-associated mutagenesis (SPM) assays revealed that colonies with a Tmpr phenotype, accumulated over a period of ten days with a frequency of ~ 60 ×10−7. The Tmpr system was further modified to study TMM by constructing a ΔthyA ΔthyB strain carrying an IPTG-inducible Pspac-thyA cassette. In conditions of transcriptional induction of thyA, the generation of Tmpr colonies increased ~3-fold compared to conditions of transcriptional repression. Further, the Mfd and GreA factors were necessary for the generation of Tmpr colonies in the presence of IPTG in B. subtilis. Because GreA and Mfd facilitate transcription-coupled repair, our results suggest that TMM is a mechanim to produce genetic diversity in highly transcribed regions in growth-limited B. subtilis cells.

Introduction

The ability of stressed microbial subpopulations to acquire genetic alterations in response to a persistent non-lethal pressure allowing them to escape from growth-limiting conditions has been termed adaptive or stationary-phase mutagenesis (SPM) [13]. This phenomenon promotes genetic diversity and is conserved in prokaryotes and eukaryotes [47].

In Bacillus subtilis, the mechanisms underlying SPM have been successfully investigated in the strain YB955 bearing the chromosomal auxotrophies hisC952, leuC427 and metB5 [6]. Using this gain-of-function (reversion) mutagenesis system it has been shown that adaptive mutations arise from error-prone processing of mismatched and chemically modified DNA bases [811]. Additional evidence revealed a direct correlation between transcriptional derepression and SPM in non-dividing B. subtilis cells [12, 13]. Recent results showed that in growth-limited B. subtilis cells, the transcription repair-coupling factor (Mfd) promotes mutagenic events in transcriptionally active genes coordinating error-prone repair events that required nucleotide excision (NER) and base excision (BER) repair components as well as low-fidelity DNA synthesis [14]. Notably, a mutagenic pathway dependent on Mfd, the NER system and the error prone polymerase PolY1 that prevents conflicts between the replicative and transcriptional machineries has been recently described in growing B. subtilis cells [15]. In Escherichia coli, the elongation factor of the RNA polymerase NusA has been found to be necessary for stress-induced mutagenesis [16]. In addition to Mfd, B. subtilis possesses the transcriptional factors NusA and GreA [17, 18]; however, a possible contribution of these factors in modulating transcriptional-mediated mutagenic events in nutritionally stressed, non-growing B. subtilis, remains to be elucidated.

Two types of genetic alterations affect organism's physiology, namely, the gain-of- (i.e., reversion mutagenesis) and the loss-of-function (i.e., forward mutagenesis) mutations; the former may generate a product with an enhanced or a novel function whereas the latter one leads to reducing or abolishing protein function [19]. As noted above, SPM frequencies in B. subtilis have commonly calculated from mutation events occurring in strain YB955 containing point mutations in the chromosomal genes, metB5 (ochre), leuC427 (missense) and hisC952 (amber) [6]. However, evolutionary experiments conducted in distinct microorganisms have revealed that loss-of- or modification-of- are by far more frequent than gain-of-function mutational events [19].

Thymidine synthesis plays an essential role in DNA metabolism. In both, prokaryotes and eukaryotes, thymidylate synthase (TMS) converts dUMP to dTMP using N5N10-methylenetetrahydrofolate as cofactor [20]. Thus, tetrahydrofolate analogs such as aminopterin or trimethoprim that inhibit dihydrofolate reductase (DHFR) also inhibit thymidine synthesis [21, 22]. B. subtilis possesses thyA and thyB encoding TMSs [23, 24]. In this bacterium, thymine auxotrophs incorporate this metabolite much more efficiently than the wild type strain and are able to grow in the presence of aminopterin or trimethoprim [2022]. Thus, loss of TMS function allows selection of Tmpr mutants that requires exogenous thymine for growth [2022].

Here, we developed a loss-of-function mutagenesis system based on the production of trimethoprim resistant colonies (Tmpr) resulting from mutations that inactivate TMS as an efficient and more direct method to analyze growth and SPM in B. subtilis. The use of this system showed that a null thyB B. subtilis strain proficient for thyA accumulated a high frequency of adaptive Tmpr colonies. Furthermore, under growth-limited conditions, derepression of a wild type thyA gene in a ΔthyA ΔthyB background revealed a positive correlation between transcription and accumulation of Tmpr colonies. Interestingly, the generation of transcription-associated Tmpr mutants was dependent not only on Mfd but also on GreA, two proteins known to process RNA polymerase (RNAP) pausing. Thus, our results suggest that under conditions of nutritional stress RNAP backtracking and/or RNAP pausing promote mutagenesis in non-growing B. subtilis cells.

Materials and methods

Bacterial strains and culture conditions

The bacterial strains used in this study are listed in Table 1. B. subtilis YB955 is a prophage-“cured” strain that contains the hisC952, metB5, and leuC427 alleles [6, 25]. The procedures for transformation and isolation of chromosomal and plasmid DNA were as described previously [2628]. Liquid cultures of B. subtilis strains were routinely grown in Penassay broth (PAB) (antibiotic A3 medium; Difco Laboratories, Sparks, MD). When required, neomycin (Neo; 10 μg ml-1); tetracycline (Tet; 10 μg ml-1); spectinomycin (Sp; 100 μg ml-1); erythromycin (Em; 1 μg ml-1 or 5 μg ml-1); chloramphenicol (Cm; 5 μg ml-1); kanamycin (Kan; 50 μg ml-1); trimethoprim (Tmp; 10 μg ml-1), or IPTG (0.07 mM) was added to the medium. E. coli cultures were grown in Luria-Bertani (LB) medium supplemented with ampicillin to a final concentration of 100 μg ml-1. The PCR products were generated with Vent DNA polymerase (New England BioLabs, Ipswich, MA) and the set of homologous oligonucleotide primers described in Table 2.

thumbnail
Table 2. Oligonucleotide primers employed in the PCR reactions.

https://doi.org/10.1371/journal.pone.0179625.t002

Strains construction

Knockouts in the genes of interest were constructed by marker exchange between the chromosome- and plasmid-borne alleles. A plasmid to disrupt thyA was constructed as follows, a 493-bp fragment from the 5'-thyA region (-300 to +164 relative to thyA start codon) and a 473-bp fragment from the 3'-thyA region (+600 to +1073 relative to thyA start codon) were PCR amplified using chromosomal DNA from B. subtilis 168. The set of oligonucleotide primers 3, 4 and 5, 6 (Table 2) were used for amplification of the 5'-thyA and 3'-thyA fragments, respectively. The amplified 5'-thyA and 3'-thyA fragments were cloned between the SphI/SalI and BamHI/SalI sites of pBEST502 [29]. The resulting construction pPERM1014 was replicated in E. coli XL10-GOLD Kanr (Stratagene, Cedar Creek, TX) and employed to transform the strain B. subtilis YB955 to neomycin resistance, thus generating strain B. subtilis PERM1000 (Table 1).

A ΔthyB strain was generated by amplifying a fragment from position 38 to 675 from the thyB open reading frame was amplified by PCR using chromosomal DNA from B. subtilis YB955 and the oligonucleotide primers 5 and 6 (Table 2). The PCR product was cloned between the HindIII and BamHI sites of the integrative plasmid pMUTIN4 [30]. The resulting construct was designated pPERM1013 and used to transform B. subtilis YB955 and PERM1000, thus generating strains PERM1037 (ΔthyB; Emr) and PERM1074 (ΔthyA ΔthyB; Neor Emr), respectively (Table 1).

Genetic inactivation of error prevention GO system (ytkD, mutM and yfhQ) from B. subtilis YB955 was achieved by transforming the strain ΔthyB (PERM1037; Table 1) with genomic DNA isolated from B. subtilis PERM573 (ΔytkD::Neo ΔmutM::Tet ΔmutY::Spc). This procedure generated the strain B. subtilis PERM1491 (ΔthyB GO; Emr Cmr Tetr Spcr), respectively (Table 1). The thyB mutSL mutant in the YB955 background was generated by transforming strain PERM151 (ΔmutSL::Neo) [8] with genomic DNA isolated from B. subtilis PERM1037 (ΔthyB), thus generating strain PERM1565 (ΔthyB mutSL; Emr Neor) (Table 1). Disruptions of the appropriate chromosomal genes were confirmed by PCR.

The strain containing the thyA gene under the control of an IPTG-inducible promoter was constructed as follows. The open reading frame (ORF) of thyA was PCR amplified from B. subtilis 168 chromosomal DNA and the oligonucleotide primers 7 and 8 (Table 2). The PCR product (1,194 bp) was purified from a low-melting-point agarose gel and cloned between the SalI and SphI sites of the amyE integrative vector pdr111 (a gift from David Rudner), immediately downstream of the IPTG-inducible Phyperspank promoter (Phs). The resulting plasmid (pPERM1099) was replicated in E. coli XL10-GOLD (Stratagene). This construct was transformed and integrated into the amyE locus of B. subtilis PERM1074 (ΔthyA thyB) to generate the strain B. subtilis PERM1100 (ΔthyA ΔthyB; amyE::Phs-thyA; Neor Emr Spcr) (Table 1). As an experimental control, the empty pdr111 vector was also recombined in the amyE locus thus generating the strain B. subtilis PERM1075 (ΔthyA ΔthyB; amyE::Phs; Neor Emr Spcr) (Table 1).

To disrupt mfd, competent cells of strain B. subtilis PERM1100 were transformed with chromosomal DNA of B. subtilis YB9800 (Δmfd::Cm) [12] to generate the strain B. subtilis PERM1171 (ΔthyA ΔthyB Δmfd; amyE::Phs-thyA; Neor Emr Cmr Spcr) (Table 1).

An integrative plasmid designed to inactivate greA was constructed as follows. A fragment of greA was first amplified by PCR using chromosomal DNA from B. subtilis 168 and the set of oligonucleotide primers 9 and 10 (Table 2). The 222-bp PCR product, extending from position 115-bp to 318-bp downstream of the greA start codon was cloned between the EcoRI-BamHI sites of the integrative plasmid pMUTIN4-Cat [11]. The resulting construct designated pPERM1191was used to transform B. subtilis PERM1100, generating strain PERM1192 (ΔthyA ΔthyB ΔgreA; amyE::Phs-thyA; Neor Emr Cmr Spcr) (Table 1). The single- or double-crossover events leading to inactivation of the appropriate genes and integration of transcriptional cassettes were corroborated by PCR with specific oligonucleotide primers.

Stationary-phase mutagenesis soft-agar overlay assays

Cultures were grown in flasks containing antibiotic A3 medium with aeration (250 rpm) at 37°C to 90 min after the cessation of exponential growth (designated T90). Growth was monitored with a spectrophotometer measuring the optical density at 600nm (OD600). The cultures were centrifuged at 10, 000 × g for 10 min and resuspended in 10 ml of 1X Spizizen Minimal Salts (SMS) [31]. Aliquots of 0.1 ml were then spread plated on Spizizen minimal medium (SMM) containing 1X SMS, 0.5% (w/v) dextrose, 50 μg isoleucine ml-1, 50 μg glutamate ml-1, 1.5% (w/v) agar (BD, Bioxom), 50 μg histidine ml-1 of and a trace amount (200 ng ml-1) of leucine and methionine (amino acids from Sigma-Aldrich, St. Louis, MO), with or without 0.07 mM IPTG. The initial titer was determined from the 10-ml culture. Starting from 48 h of incubation, a set of plates was overlaid with soft agar (0.7% (w/v) agar and prewarmed at 42°C lacking histidine and containing 0.07 mM IPTG, 50 μg thymine ml-1, 10 μg Tmp ml-1, 50 μg methionine and leucine ml-1. Of note, adjustment of the final concentrations for IPTG, methionine and leucine considered the volume and IPTG concentration in the medium dispensed previous to performing the overlay. The plates were incubated for two days, and the number of Tmpr colonies was scored. The initial titers were used to normalize the cumulative number of resistant colonies per day to the number of CFU plated. Assays were replicated three times, and the experiment was repeated at least twice. The viability of the non-revertant cell background was assessed every other day as follows. Using a sterile Pasteur pipette, a plug of agar was removed from the non-revertant background of each of five plates corresponding to one type of selection (no leucine and methionine without IPTG or no leucine and methionine with IPTG). The plugs were combined in 0.4 ml of 1X SMS, serially diluted, and plated in triplicate on SMM containing 50 μg ml-1 of the required amino acids. Colonies were counted after 48 h of incubation.

Phenotypic analysis of B. subtilis thy mutants

Trimethoprim resistance of the strains of interest was analyzed on solid SMM containing 1X SMS, 0.5% (w/v) dextrose, 50 μg ml-1 isoleucine, glutamate, histidine, methionine, leucine and thymine. All plates contained thymine at 50 μg ml-1 and the concentration (μg/ml) of trimethoprim indicated in Table 3. When necessary, the media was supplemented with IPTG to a final concentration of 0.1 mM. Growth was scored after 24 hr at 37°.

Analysis of mutation frequencies to Tmpr and Rifr

Essentially, the appropriate strains were propagated for 12 h at 37°C in A3 medium with proper antibiotics. For Tmpr, mutation frequencies were determined by plating aliquots on six LB plates containing 10 μg ml-1 trimethoprim and 50 μg ml-1 thymidine, and the trimethoprim-resistant (Tmpr) colonies were counted after 2 days of incubation at 37°C. The same procedure was employed to determine the Rifr phenotype, except that mutant colonies were selected in LB plates containing 10 μg ml-1 rifampin. The number of cells used to calculate the frequency of mutation to Tmpr or Rifr was determined by plating aliquots of appropriate dilutions on LB plates without antibiotics and incubating the plates for 24 to 48 h at 37°C. These experiments were repeated at least three times.

Total RNA extraction

Cultures were grown to saturation in 1X SMS containing 0.5% (w/v) dextrose; 50 μg ml-1 of Ile, Glu, methionine, histidine, and leucine; Ho-Le trace elements; 5 mM MgSO4; and 1 mM IPTG. Samples were removed at mid-exponential (approximately OD600nm = 0.5) and stationary (150 min after onset of the stationary phase; T150) growth phases, centrifuged at 5, 000 × g and 4°C for 10 min and frozen at -20°C. RNA was extracted from the pellets using a Tri reagent (Molecular Research Center, Inc. Cincinnati, OH). After DNAse I treatment, the samples were analyzed by PCR to confirm the absence of genomic DNA with Vent DNA polymerase and the set of primers 11 and 12 described in Table 2. The RNA content in the samples was quantitated using a Nanodrop spectrophotometer (Thermo Fisher Scientific, Dubuque, IA).

Reverse transcription and quantitative real-time PCR

One microgram of RNA was reverse transcribed using an ImPromII reverse transcriptase kit (Promega), as directed, with random hexamers (0.5-μg final concentration). No-reverse transcriptase (NRT) controls were included for all examples. Master mixes for real-time PCR contained Absolute QPCR SYBR green Mix (Thermo Fisher Scientific) and a 70 mM final concentration of the oligonucleotide primers thyART FW and thyART RV (191-bp amplicon) or of veg FW and veg RV (82-bp amplicon) described in Table 2. Three replicates from each culture condition containing 4 μl of cDNA were assayed and normalized to the expression of the internal control gene veg [32, 33]. Two replicates were assessed for NRT and no-template controls. Quantitative real-time PCR was run on a Bio-Rad iCycler iQ Real-Time PCR Detection System (Bio-Rad, Hercules, CA), using the manufacturer´s suggested protocol and an annealing temperature of 57°C, followed by a melting profile and assessment of amplicon size on an agarose gel. Results were calculated by the 2-ΔΔC(T) (where CT is threshold cycle) method for relative fold expression [34].

DNA sequencing

Colonies with a Tmpr phenotype were independently propagated in liquid A3 medium supplemented with 10 μg/ml trimethoprim and 50 μg ml-1 thymidine and subjected to DNA isolation [27]. The thyA open reading frame of each colony was amplified by PCR using high fidelity and specific oligonucleotide primers (Table 2). Sequencing services were carry out by Functional Biosciences, Inc. (Madison, WI).

Results and discussion

Spontaneous mutation frequencies to trimethoprim resistance in growing B. subtilis cells

To implement a loss-of function system to study mutagenesis in B. subtilis, the thyA gene that codes for TMS was considered as a possible target [23, 24]. In B. subtilis the presence of ThyA prevents the incorporation of exogenous thymine into DNA synthesis [20]; therefore, strains deficient for this activity incorporate this metabolite more efficiently than ThyA-proficient strains [20]. Of note, thyA mutants of B. subtilis are able to grow in medium supplemented with DHFR inhibitors such as trimethoprim; if thymine is present in the culture medium [20], spontaneous colonies with a Tmpr phenotype can be selected [20, 24, 35]. This phenomenon has been described in several bacteria, including Escherichia coli [3537]. B. subtilis possesses two TMS encoding genes; i.e., thyA and thyB respectively [20]. It has been reported that thyB expression takes place when B. subtilis is grown under temperatures below 37°C contributing with only 5–8% of the total TMS activity present in this bacterium [23]. Since thyB is the first cistron of the thyB-dfrA-ypkP operon, we employed a gene construct (pPERM1013; Table 1) to only disrupt thyB and left dfrA-ypkP under control of the IPTG-inducible Pspac promoter. Under our experimental conditions, no significant differences were observed in the growth rates of this mutant with respect to those of the parental strain YB955. Thus, in A3 medium, in the absence or presence of 0.1 mM IPTG the thyB strain grew with similar doubling time values as those of its parental strain (i.e., 31 ± 2, 30 ± 2.5 and 31 ± 1.8 min, respectively); furthermore, this strain was incapable of growing in minimal medium supplemented with trimethoprim and thymidine (Table 3).

The ΔthyB strain was first used to analyze the spontaneous appearance of colonies resistant to trimethoprim in B. subtilis cells. As shown in Fig 1, colonies with a Tmpr phenotype appeared, in solid minimal medium supplemented with thymidine, with a spontaneous mutation frequency of ~1.4 ± 0.11 × 10−9. These results are in good agreement with the mutation frequencies reported for the strain B. subtilis YB955 using the rpoB gene as a marker of mutagenesis [10, 14]. Previous reports, employing the Rifr phenotype demonstrated that inactivation of the canonical mismatch (MMR) and guanine-oxidized (GO) repair systems increased, two and three orders of magnitude, respectively, the spontaneous mutagenesis in growing B. subtilis cells [8, 9]. As shown in Fig 1, the genetic inactivation of the GO system, which in B. subtilis is composed of the MutY, MutM and YtkD proteins [9, 38] increased the mutation frequency to Tmpr in the ThyB-deficient B. subtilis strain around 1000 times. Furthermore, our results revealed that disruption of the mutSL operon, encoding the MMR system, increased the frequency of mutation to trimethoprim resistance in the ΔthyB strain around 30 times. As shown in Fig 1, the mutation frequencies as determined with the rpoB marker in the mutSL and GO-deficient strains were similar to those determined with the thyA gene. Therefore, the loss-of-function mutagenesis system described in this report is robust and can be successfully employed to study the spectrum of mutagenic events that occur in thyA in a population of B. subtilis cells.

thumbnail
Fig 1. Frequencies of spontaneous mutation to Tmpr (gray bars) and Rifr (black bars) of different B. subtilis strains.

B. subtilis PERM1037 (ΔthyB), PERM1491 (ΔthyB GO system deletion), PERM1565 (ΔthyB mutSL) were grown overnight in PAB medium, and frequencies of mutation to Rifr or Tmpr were determined as described in Materials and Methods. Each bar represents the mean of data collected from three independent experiments done in sixtuplicate, and the error bars represent standard errors of the means (SEM).

https://doi.org/10.1371/journal.pone.0179625.g001

Mutational spectra of thyA in colonies exhibiting a Tmpr phenotype

We further investigated the mutagenic events associated with the loss of TMS function in the thyB-deficient strain. It was recently shown, that the heterologous expression of thyP3, a homologue of thyA, from the ITPG-inducible Pspac promoter, resulted in the production of mutations in the -5 region of Pspac conferring trimethoprim resistance in actively growing B. subtilis cells [39]. In this study, we were interested in identifying mutations affecting the function of ThyA under conditions of native thyA expression; therefore, the thyA gene of 40 colonies exhibiting a Tmpr phenotype was PCR amplified with high-fidelity DNA polymerase. All the samples produced amplification DNA bands of the expected size for the thyA ORF (namely, 837 bp), which were subsequently subjected to DNA sequencing to identify the type of mutations conferring the Tmpr phenotype. A wide spectrum of frameshift mutations and base substitutions were detected in the sequenced thyA mutants, predominating the insertions/deletions over the base substitutions in a proportion of 60% to 40% (Table 4 and Figs 2 and 3). Among the substitutions, 63% corresponded to transversions and 37% to transitions events, predominating the A→T (~29%) and A→G (~17%) changes (Table 4). Of note, three of the base changes generated non-sense mutations producing truncated ThyA proteins (Table 4 and Fig 2). It was found that most of the missense mutations in thyA resulted in non-conserved amino acid changes, moreover, two of the substitutions changed an alanine residue for the secondary structure-disrupting amino acid proline and one of them switched an isoleucine for the bulky aromatic residue phenylalanine (Table 4 and Fig 2). Importantly, 14 of the 17 missense mutations took place in codons encoding amino acids conserved in both thymidylate synthases (Table 4 and Fig 2). However, additional mutations in residues non-conserved in E. coli ThyA, were identified in the enzyme from B. subtilis, including, Ala68→Asp, Trp31→Arg and Ala68→Pro (Fig 2). Thus, in addition to Arg181 previously reported as essential for ThyA activity in E. coli [40], our results identified additional residues necessary for the proper function of ThyA in B. subtilis.

thumbnail
Fig 2. Base substitutions and predicted amino acid changes in the thyA ORF from Tmp-resistant B. subtilis colonies.

The deduced amino acid sequences of thyA from B. subtilis and E. coli were aligned with the Clustal Omega program (http://www.ebi.ac.uk/Tools/msa/clustalo/). The position of the amino acid mutated is indicated with a blue arrow. The amino acid change (in the single letter format) is shown in bold red, above the arrow, the number of changes in a particular position is shown between parentheses. The amino acids involved in the mutation invariably conserved in both proteins are marked in black bold. The mutations that generated a termination codon are indicated as STOP, in bold red letters.

https://doi.org/10.1371/journal.pone.0179625.g002

thumbnail
Fig 3. Frame-shift mutations in the thyA ORF from B. subtilis colonies with a Tmpr phenotype.

The black arrow depicts the open reading frame of thyA in kbp. Nucleotide insertions or deletions at each position are shown above or below the thyA ORF, respectively. Oligonucleotide deletions and insertions as well as its positions are shown between brackets. I, insertions; D, deletions.

https://doi.org/10.1371/journal.pone.0179625.g003

thumbnail
Table 4. Base substitutions in thyA alleles from Tmpr colonies.

https://doi.org/10.1371/journal.pone.0179625.t004

As noted above, insertion/deletions were also detected in the thyA sequence of colonies exhibiting a Tmpr phenotype. As shown in Fig 3, the insertion events took place between the positions 553–562 of the thyA ORF and consisted of single base additions, including, 4 adenines 2 thymines or 1 guanine. On the other hand, the frameshift deletions that took place between the positions 455–458 of the thyA ORF, involved the single loss of 7 thymines or the dinucleotide guanine-adenine (Fig 3). Notably, the insertion and deletions occurred in the base repeats GGGAAA and GAATT, respectively, and gave rise to truncated non-functional ThyA proteins. Thus, replication errors in these sequences that escaped the action of the mismatch repair machinery may be involved in generating these types of mutations. In support of this notion, the genetic inactivation of the MMR system (MutSL) in the ΔthyB strain increased by about two orders of magnitude the mutation frequency to trimethoprim resistance (Fig 1). Altogether, our results revealed that a wide spectrum of mutagenic events may lead to loss of ThyA function corroborating that the Tmpr system informs on the genetic events underlying stress-associated mutagenesis in B. subtilis.

Analysis of stationary-phase associated mutagenesis to trimethoprim resistance in nutritionally stressed B. subtilis cells

We employed the ΔthyB strain to investigate whether loss-of thymidylate synthase (ThyA) function leading to Tmpr resistance can be employed to measure adaptive mutagenesis in B. subtilis. To this end, B. subtilis cells of the ThyB-deficient strain collected from the stationary-phase of growth were extensively washed to eliminate residual nutrients and plated in selective Spizizen Minimal Medium (SMM) as described in Materials and Methods. The number of colonies that acquired a Tmpr phenotype under growth-limited conditions was scored every two days for a ten days period. The results revealed that Tmpr colonies from the ΔthyB thyA+ accumulated over a period of ten days with a frequency of ~65 ± 11 × 10−7 (Fig 4A), demonstrating the feasibility of employing this loss-of-function system to study mutagenesis in non-growing B. subtilis cells. Notably, the increase in the number of Tmpr colonies took place without significant changes in the viable cell count number providing thus additional support for this assumption (Fig 4B).

thumbnail
Fig 4.

(A). Frequencies of stationary-phase accumulation of Tmpr colonies of strain B. subtilis PERM1037 (thyA+ thyB-) were determined as described in Materials and Methods. Data were normalized to initial cell titers ± SD and represent counts averaged from three separate tests. (B). Ability of strain PERM1037 to survive Met-Leu-. Data were collected from plugs removed from three plates and titers were plated on media containing all essential amino acids every other day for testing of viability of non-revertant background cells (see Materials and Methods for details). Data is represented as the number of CFU per plug. Error bars represent 1 standard error of the mean.

https://doi.org/10.1371/journal.pone.0179625.g004

Transcription-associated mutagenesis analysis in growth-limited B. subtilis cells acquiring a Tmpr phenotype

Previous studies have reported on the role of transcription in stationary-phase mutagenesis in B. subtilis [12]. It has been shown that a combination of transcriptional derepression and error-prone repair events in stress conditions can modulate the generation of mutations in highly transcribed DNA regions [13, 41].

To investigate whether acquisition of Tmpr resistance can be used to study transcription-associated mutagenesis (TAM) in nutritionally stressed bacteria we engineered a double thyA thyB knock out strain that overexpressed a wild type copy of thyA from the IPTG-inducible Phyperspank (Phs) promoter; this strain was termed B. subtilis PERM1100 (Table 1). As specified in Table 3, the thyA thyB mutant was able to grow in SMM supplemented with Tmp and thymidine. However, IPTG-induction of thyA expression abrogated the growth of the strain PERM1100 in this medium. Further, the lack of two amino acids (methionine and leucine) in the incubation agar media prevented the growth of cells. Therefore, the strain's properties and the double amino acid starvation allowed us to inquire whether derepression of thyA in growth-limited cells promote mutagenic events influencing the production of Tmpr colonies. As shown in Fig 5A in reference to the non-induced condition, derepression of the Phs-thyA construct resulted in a 3-fold increase in the production of colonies with a Tmpr phenotype.

thumbnail
Fig 5.

Frequencies of stationary-phase accumulation of Tmpr colonies of strains B. subtilis PERM1100 (A), PERM1171 (B) and PERM1192 (C) under repressed (–IPTG; white symbols) or induced (+ IPTG: black symbols) were determined as described in Materials and Methods. Data represent counts from three plates averaged from three separate tests, normalized to initial cell titers.

https://doi.org/10.1371/journal.pone.0179625.g005

Our loss-of function mutagenesis system indicate that increased transcription levels of thyA correlated with an increased production of colonies with a Tmpr phenotype in the ΔthyAthyB amyE::Phs-thyA strain. Of note, these results are in good agreement with former studies showing a positive correlation between derepression of a leuC allele and production of Leu+ prototrophs in B. subtilis [13]. Furthermore, viability of the strain PERM1100 did not significantly change in the absence and/or presence of the inducer IPTG (Fig 6A); therefore, the increase in the number of Tmpr can be uncoupled from the growth of the sensitive strain in the plate. To confirm that accumulation of Tmpr mutants were due to increases in transcription levels of thyA, the mRNAs of this gene were quantified by qRT-PCR in stationary-phase cells of strain PERM1100 under induced and non-induced conditions. The results of this analysis showed that the mRNA levels of thyA increased over 3-fold (i.e. from 0.5 ± 0.06 to 4.5 ± 0.4) when IPTG was added to the medium. Therefore, the loss-of-function system implemented here can be successfully employed to investigate the mutagenic processes occurring in transcriptionally active DNA regions that allow cells to escape from non-proliferating conditions. Of note, the discovery of missense and nonsense mutations affecting B. subtilis ThyA function (Fig 2), will allow to adapt this system to determine the production of thyA+ revertants in nutritionally stressed thyA- thyB- cells overexpressing point-mutated alleles of thyA.

thumbnail
Fig 6.

Ability of strain B. subtilis PERM1100 (A), PERM1171 (B) and PERM1192 (C) to survive Met-Leu- starvation in selective media supplemented (black bars) or not (white bars) with IPTG. Data were collected from plugs removed from three plates and titers were plated on media containing all essential amino acids every other day for testing of viability of non-revertant background cells (see Materials and Methods for details). Data is represented as the number of CFU per plug. Error bars represent 1 standard error of the mean.

https://doi.org/10.1371/journal.pone.0179625.g006

A recent study revealed that Mfd, a protein that couples DNA repair with transcription, promotes mutagenic events that increase the production of His+, Met+ and Leu+ revertants in starved cells of strain B. subtilis YB955 [12, 13]. Here, we inquired whether the Mfd requirement for transcription-associated SPM can be detected with the loss-of-function Tmpr mutagenesis system. To this end, the gene mfd was genetically inactivated in the ΔthyA thyB strain bearing the Phs-thyA construct. The number of Tmpr colonies produced by this strain under starving conditions was determined over a period of ten days under conditions where the thyA gene was repressed or derepressed for transcription. As shown in Fig 5B, in reference to the strain harboring an intact copy of Mfd, the number of adaptive Tmpr colonies produced by the Mfd-deficient strain did not significantly differ under conditions of repression or derepression for the thyA gene. Therefore, employing a system that overexpresses a different gene; namely, thyA, our results provide further support for the concept that Mfd is a factor required for TMM in growth-limited B. subilis cells. As noted above, NusA, the elongation factor of the RNA polymerase has been found to be necessary for stress-induced mutagenesis in Escherichia coli [16]. Remarkably, disruption of greA in the thyA thyB strain abrogated the production of Tmpr colonies under conditions that induced thyA expression. Thus, as shown in Fig 5C, the levels of Tmpr mutagenesis in this strain were almost similar under conditions that induced or repressed thyA. Importantly, these results took place in the absence of growth since viability of the GreA-deficient strain did not significantly change in the absence and/or presence of the inducer IPTG (Fig 6C). Based on these results is feasible to speculate that in nutritionally stressed B. subtilis cells, processing of paused or backtracked RNAP-DNA complexes promote mutagenic events that allow cells to escape from growth-limited conditions. Interestingly, in E. coli, both Mfd and GreA are speculated to prevent UvrD-dependent RNAP backtracking and repair during transcription [42]. However, the consequences of such transcriptional events on growth and stress-associated mutagenesis remain to be elucidated.

In summary, the mutagenesis method implemented here and the contribution of GreA to TMM in B. subtilis cells add further elements to our understanding on how bacteria develop beneficial mutations, including antibiotic resistance, under stressful conditions.

Acknowledgments

The authors are grateful for the excellent technical assistance of Karina Olmos-López and Norma Ramírez-Ramírez.

References

  1. 1. Bridges BA. The role of DNA damage in stationary phase (“adaptative”) mutation. Mutat Res. 1998; 408: 1–9. pmid:9678058
  2. 2. Rosenberg SM, Thulin C, Harris RS. Transient and heritable mutators in adaptative evolution in the lab and in nature. Genetics. 1998; 148: 1559–1566. pmid:9560375
  3. 3. Robleto EA, Yasbin R, Ross C, Pedraza-Reyes M. Stationary phase mutagenesis in Bacillus subtilis: a paradigm to study genetic diversity programs in cells under stress. Crit Rev Biochem Mol Biol. 2007; 42:327–339. pmid:17917870
  4. 4. Cairns J, Overbaugh J, Miller S. The origin of mutants. Nature. 1988; 335: 142–145. pmid:3045565
  5. 5. Kasak L, Horak R, Kivisaar M. Promoter-creating mutations in Pseudomonas putida: a model system for the study of mutation in starving bacteria. Proc Natl Acad Sci USA. 1997; 94: 3134–3139. pmid:9096358
  6. 6. Sung HM, Yasbin RE. Adaptive, or stationary-phase, mutagenesis, a component of bacterial differentiation in Bacillus subtilis. J Bacteriol. 2002; 184: 5641–5653. pmid:12270822
  7. 7. Halas A, Baranowska H, Policinska Z. The influence of the mismatch repair system on stationary phase mutagenesis in the yeast Saccharomyces cerevisiae. Curr Genet. 2002; 42:140–146. pmid:12491007
  8. 8. Pedraza-Reyes M, Yasbin RE. Contribution of the mismatch DNA repair system to the generation of stationary-phase-induced mutants of Bacillus subtilis. J Bacteriol. 2004; 186: 6485–6491. pmid:15375129
  9. 9. Vidales LE, Cardenas LC, Robleto E, Yasbin RE, Pedraza-Reyes M. Defects in the error prevention oxidized guanine system potentiate stationary-phase mutagenesis in Bacillus subtilis. J Bacteriol. 2009; 191:506–513. pmid:19011023
  10. 10. Debora BN, Vidales LE, Ramirez R, Ramirez M, Robleto EA, Yasbin RE, et al. Mismatch repair modulation of MutY activity drives Bacillus subtilis stationary-phase mutagenesis. J Bacteriol. 2011; 193:236–245. pmid:20971907
  11. 11. Barajas-Ornelas RDC, Ramirez-Guadiana FH, Juarez-Godinez R, Ayala-Garcia VM, Robleto EA, Yasbin RE, et al. Error-prone processing of apurinic/apyrimidinic (AP) sites by PolX underlies a novel mechanism that promotes adaptive mutagenesis in Bacillus subtilis. J Bacteriol. 2014; 196:3012–3022. pmid:24914186
  12. 12. Ross C, Pybus C, Pedraza-Reyes M, Sung HM, Yasbin RE, Robleto E. Novel role of mfd: effects on stationary-phase mutagenesis in Bacillus subtilis. J Bacteriol. 2006; 188:7512–7520. pmid:16950921
  13. 13. Pybus C, Pedraza-Reyes M, Ross CA, Martin H, Ona K, Yasbin RE, et al. Transcription-associated mutation in Bacillus subtilis cells under stress. J Bacteriol. 2010; 192:3321–3328. pmid:20435731
  14. 14. Gómez-Marroquín M, Vidales LE, Debora BN, Obregón-Herrera A, Robleto EA, Pedraza-Reyes M. Role of Bacillus subtilis DNA glycosylase MutM in counteracting oxidative-induced DNA damage and in stationary-phase-associated mutagenesis. J Bacteriol. 2015; 197:1963–1971. pmid:25825434
  15. 15. Millon-Weaver S, Samadpour NA, Moreno-Hable DA, Nugen P, Briinacher MJ, Weiss E, et al. An underlying mechanism for the increased mutagenesis of lagging-strand genes in Bacillus subtilis. Proc Natl Acad Sci USA. 2015; 112:E1096–E1105. pmid:25713353
  16. 16. Cohen SE, Walker GC. The transcription elongation factor NusA is required for stress-induced mutagenesis in Escherichia coli. Curr Biol. 2010; 20:80–85. pmid:20036541
  17. 17. Ishikawa S, Oshima T, Kurokawa K, Kusuya Y, Ogasawara N. RNA polymerase trafficking in Bacillus subtilis cells. J Bacteriol. 2010; 192: 5778–5787. pmid:20817769
  18. 18. Kusuya Y, Kurokawa IK, Ishikawa S, Ogasawara N, Oshima T. Transcription factor GreA contributes to resolving promoter-proximal pausing of RNA polymerase in Bacillus subtilis cells. J Bacteriol. 2011; 193: 3090–3099. pmid:21515770
  19. 19. Behe MJ. Experimental evolution, loss-of-function mutations, and “The first rule of adaptative evolution”. Q Rev Biol. 2010; 85; 419–445. pmid:21243963
  20. 20. Wilson MC, Farmer JL, Rothman F. Thymidylate synthesis and amonipterin resistance in Bacillus subtilis. J Bacteriol. 1966; 92:186–196. pmid:4957432
  21. 21. Quinlivan EP, McPartlin J, Weir DG, Scott J. Mechanism of the antimicrobial drug trimethoprim revisited. FASEB J. 2000; 14:2519–2524. pmid:11099470
  22. 22. Hawser S, Lociuro S, Islam K. Dihydrofolate reductase inhibitors as antibacterial agents. Biochem Pharma. 2006; 71:941–948.
  23. 23. Fox KM, Maley F, Garibian A, Changchien LM, Van Roey P. Crystal structure of thymidylate synthase A from Bacillus subtilis. Protein Sci. 1999; 8:538–544. pmid:10091656
  24. 24. Neuhard J, Price AR, Schack L, Thomassen E. Two thymidylate synthetases in Bacillus subtilis. Pro Natl Acad Sci USA. 1978; 75:1194–1198.
  25. 25. Yasbin RE, Fields PI, Andersen BJ. Properties of Bacillus subtilis 168 derivatives freed of their natural prophages. Gene. 1980; 12:155–159. pmid:6783474
  26. 26. Boylan RJ, Mendelson NH, Brooks D, Young FE. Regulation of the bacterial cell wall: analysis of a mutant of Bacillus subtilis defective in biosynthesis of teichoic acid. J Bacteriol. 1972; 110:281–290. pmid:4622900
  27. 27. Cutting SM, Vander Horn PB. Genetic analysis. In Harwood CR, Cutting SM (ed), Molecular biological methods for Bacillus. John Wiley & Sons, Sussex, England; 1990. pp.27–24.
  28. 28. Sambrook J, Fristsch EF, Maniatis T. Molecular cloning: a laboratory manual. 2nd ed. Cold Spring Harbor Laboratory. Cold Spring Harbor, N.Y; 1989.
  29. 29. Itaya M, Kondo K, Tanaka T. A neomycin resistance gene cassette selectable in a single copy state in the Bacillus subtilis chromosome. Nucleic Acids Res. 1989; 17:4410. pmid:2500645
  30. 30. Vagner V, Dervyn E. Ehrlich SD. A vector for systematic gene inactivation in Bacillus subtilis. Microbiology. 1998; 144:3097–3104. pmid:9846745
  31. 31. Spizizen J. Transformation of biochemically deficient strains of Bacillus subtilis by deoxyribonucleate. Proc Natl Acad Sci USA. 1958; 44:1072–1078. pmid:16590310
  32. 32. Ollington JF, Haldenwang WG, Huynh TV, Losick R. Developmentally regulated transcription in a cloned segment of the Bacillus subtilis chromosome. J Bacteriol. 1981; 147:432–442. pmid:6790515
  33. 33. Fukushima T, Ishikawa S, Yamamoto H, Ogasawara N, Sekiguch J. Transcriptional, functional and cytochemical analyses of the veg gene in Bacillus subtilis. J Biochem. 2003; 133:475–483. pmid:12761295
  34. 34. Livak KJ, Schmittgen TD. Analysis of relative gene expression data using real-time quantitative PCR and the 2-ΔΔC(T) method. Methods. 2001; 2:402–408.
  35. 35. Okada T, Yanagisawa K, Ryan FJ. Elective production of thymine-less mutants. Nature. 1960; 188:340–341. pmid:13730568
  36. 36. Okada T, Yanagisawa K, Sonohara M. Improved method for obtaining thymineless mutants of Escherichia coli and Salmonella typhimurium. J Bacteriol. 1962; 84:602–603. pmid:13939760
  37. 37. Farmer JL, Rothman F. Transformable thymine-requiring mutant of Bacillus subtilis. J Bacteriol. 1965; 89:262–263. pmid:14255676
  38. 38. Castellanos-Juárez FX, Alvarez-Alvarez C, Yasbin RE, Setlow B, Setlow P, Pedraza-Reyes M. YtkD and MutT protect vegetative cells but not spores of Bacillus subtilis from oxidative stress. J Bacteriol. 2006; 188:2285–2289. pmid:16513759
  39. 39. Sankar ST, Wastuwidyaningtyas BD, Dong Y, Lewis SA, Wang JD. The nature of mutations induced by replication-transcription collisions. Nature. 2016; 535:178–181. pmid:27362223
  40. 40. Michaels ML, Kim CW, Matthews DA, Miller JH. Escherichia coli thymidylate synthase: Amino acid substitutions by suppression of amber nonsense mutations. Proc Natl Acad Sci USA. 1990; 87: 3957–3961. pmid:2187197
  41. 41. Ambriz-Aviña V, Yasbin RE, Robleto EA, Pedraza-Reyes M. Role of base excision repair (BER) in transcription-associated-mutagénesis of nutritionally stressed non-growing Bacillus subtilis cell subpopulations. Curr Microbiol. 2016; 73:721–726. pmid:27530626
  42. 42. Kamarthapu V, Nudler E. Rethinking transcription coupled DNA repair. Curr Opin Microbiol. 2015; 24:15–20. pmid:25596348