Bacillus strains.
The
Bacillus strains used in this work are listed in
Table 1.
B. subtilis strains are derivatives of strain 168 and are isogenic derivatives of strain PS832. The
B. subtilis strain with a
gerB-lacZ fusion in the PS832 background was generated by transformation of strain PS832 to MLS
r (resistance to erythromycin plus lincomycin) with chromosomal DNA from strain AM1247 (
12), giving strain PS3709. A
B. subtilis strain with much of the
B. subtilis gerKD (hereafter termed
gerKDbs) coding sequence replaced by a tetracycline resistance (Tc
r) cassette was constructed as follows. The Tc
r cassette was PCR amplified from plasmid pFE149 (
7), and the upstream region of the
gerKDbs gene (−500 to +19 relative to the
gerKDbs translation start site [+1]) and the downstream region of the
gerKDbs gene (+222 to +721) were PCR amplified from chromosomal DNA of strain PS832. The latter two PCR products plus the amplified Tc
r cassette were used for a three-way overlap PCR, and the product was purified and ligated to pGEM-T Easy vector (Promega Corp., Madison, WI) to generate plasmid pXY1223 in
Escherichia coli. The
tet gene's promoter is oriented in the direction opposite that of the
B. subtilis gerK (hereafter termed
gerKbs) operon to minimize the effects of transcription of the
tet gene on transcription of the
gerKbs operon. Plasmid pXY1223 was used to transform
B. subtilis strain PS832 to Tc
r by a double-crossover event, giving strain PS4256; the expected chromosome structure in the
gerKDbs region of this strain was confirmed by PCR.
A
B. subtilis strain with a transcriptional fusion of the putative promoter region of
gerKDbs to the
E. coli lacZ gene was constructed as follows. The region between bp −447 to −1 relative to the putative
gerKDbs translation start site (+1) was PCR amplified from
B. subtilis PS832 DNA using primers containing EcoRI and BamHI sites (all primer sequences are available on request). The purified PCR product was digested with EcoRI and BamHI and then ligated to plasmid pDG268 (
13) that was digested with the same restriction enzymes. The resulting plasmid, pPS4289, was isolated in
E. coli and used to transform various
B. subtilis strains to a chloramphenicol-resistant (Cm
r) amylase-negative phenotype by a double-crossover event at the
amyE locus.
A
B. subtilis strain expressing
gerKDbs at
amyE in a Δ
gerKDbs background with the
gerKDbs gene expressed from its own likely promoter (see Results) was constructed as follows. A DNA fragment from −500 to +350 bp relative to the
gerKDbs translation start site (defined as +1), and encompassing the likely
gerKDbs promoter (see Results) and its coding sequence, was amplified from strain PS832 chromosomal DNA with primers containing an upstream EcoRI site and a downstream BamHI site. The PCR product was purified, digested with EcoRI and BamHI, and ligated to similarly cut plasmid pDG364 (
14). The recombinant plasmid was isolated in
E. coli and used to transform strain PS4256 (Δ
gerKDbs) to a Cm
r amylase-negative phenotype, giving strain PS4313. The double-crossover event leading to disruption of
amyE by the
gerKDbs gene plus the Cm
r cassette in strain PS4313 was confirmed by PCR.
To construct a
B. subtilis strain in which
gerKDbs expression was under the control of the strong forespore-specific P
sspB promoter (
4), the region from −500 to −1 bp relative to the translation start site of the
B. subtilis sspB gene (this region has P
sspB as well as a strong ribosome binding site [RBS]) was amplified from PS832 DNA using primers with an EcoRI site in the upstream primer and a region of overlap with the
gerKDbs translation start site in the downstream primer. The upstream primer for
gerKDbs amplification was complementary to the downstream primer for P
sspB, and the downstream
gerKDbs primer would amplify the complete
gerKDbs coding region plus 112 downstream bp that should include the likely
gerKDbs transcription terminator and a 3′ BamHI site. For overlap PCR, we amplified a product that has a P
sspB promoter plus RBS just upstream of the
gerKDbs coding region and between the EcoRI and BamHI sites, using the amplified P
sspB promoter and
gerKDbs fragments as the template. The overlap PCR product of the expected size was purified, digested with EcoRI and BamHI, and ligated to similarly cut plasmid pDG364 and the recombinant plasmid was isolated in
E. coli. This plasmid was used to transform
B. subtilis strain PS832 to a Cm
r amylase-negative phenotype by a double-crossover event, giving strain PS4314, and the expected genomic structure in the
amyE region of this strain was confirmed by PCR. Chromosomal DNA from strain PS4314 was also used to transform strains PS767 and PS3709 to a Cm
r MLS
r amylase-negative phenotype, and the resultant strains were termed PS4319 and PS4320, respectively.
B. megaterium strain GC614, which is isogenic with the plasmidless QM B1551 derivative PV361 and lacks all functional GRs due to insertion-deletions in the A-cistrons of four respective GR loci and excision of the
gerU-containing plasmid pBM700 (
36), was the host strain for plasmid-based complementation analyses. A plasmid containing the entire
B. megaterium gerK operon (hereafter referred to as
gerKbm) plus upstream promoter sequence was prepared by ligating a 4.7-kb PCR fragment, amplified from QM B1551 genomic DNA using primers with BamHI sites at the 5′ ends (all primer sequences are available on request), with plasmid pHT315t (modified from the plasmid in reference [
15]) digested with the same enzyme. The resulting plasmid, pHT-
gerKbm, was purified from
E. coli and used to transform
B. megaterium GC614 to Tc
r, giving strain GC635. The same plasmid served as the template for an inverse PCR using primers designed to remove the entire
gerKDbm open reading frame (ORF), which upon blunt-end religation would leave a 20-bp region between the stop codon of
gerKCbm and the predicted translational start site of
gerKBbm; the latter is preceded by an appropriately positioned RBS. The resultant plasmid, pHT-(
gerKbm Δ
gerKDbm), was isolated from
E. coli, verified by DNA sequencing, and used to transform strain GC614 to Tc
r, giving strain GC636.
Construction of a pHT315-based plasmid containing the
gerU* (
gerUA,
gerUC, and
gerVB) receptor operon has been described previously (
16). The cloned locus contains approximately 400 bp of sequence upstream of the predicted
gerUA translational start site, within which the putative
gerUD ORF is located. This entire region was amplified by PCR using primers with BamHI sites and ligated with plasmid pHT315t digested with the same enzyme to prepare plasmid pHT-(
gerUD gerU*), which was compatible with complementation experiments in the multiantibiotic-resistant GC614 genetic background. A plasmid in which the entire
gerUD ORF was deleted, but leaving the predicted
gerUA promoter intact, was prepared by PCR amplification of the region comprising −132 bp relative to the
gerUA translational start site to ∼200 bp downstream of the predicted
gerVB stop codon. This DNA fragment, which had flanking BamHI restriction sites, was digested and ligated with pHT315t digested with the same enzyme to prepare plasmid pHT-Δ
gerUD gerU* which was used to transform strain GC614 to Tc
r, giving strain GC630. Site-directed mutagenesis (SDM), conducted with a QuikChange Lightning SDM kit (Agilent Technologies, Wokingham, United Kingdom), was used to prepare plasmid pHT-
gerUDM10stop gerU*, in which the methionine encoded by predicted codon 10 of the
gerUD ORF, was changed to a stop codon. This plasmid, designed to result in the expression of a severely truncated GerUD protein while minimizing potential disruption to the
gerUA promoter, was used to transform strain GC614 to Tc
r, giving strain GC631.
The Gibson Assembly technique (New England BioLabs, Hitchin, United Kingdom) was used to prepare plasmids in which either gerUD or gerKDbm were located between the gerUC and gerVB ORFs within the gerU* operon. First, a PCR fragment spanning 444 bp with respect to the gerUA translational start site to position 2826, which included ORFs for gerUDM10stop, gerUA, and gerUC, was prepared using plasmid pHT-gerUDM10stop gerU* as the template DNA. A second PCR fragment comprising the gerVB ORF plus 104 bp upstream of the gerVB translational start site and 200 bp downstream of the stop codon was prepared from the same plasmid. Finally, PCR fragments spanning the gerUD and gerKDbm ORFs were prepared from genomic DNA. The appropriate fragments, including BamHI-linearized pHT315t vector backbone were subsequently purified and assembled using Gibson Assembly master mix (New England BioLabs, Hitchin, United Kingdom), and the reaction mixtures were used to transform E. coli. Purified and sequence-validated pHT-(gerUDM10stop gerUA gerUC gerUD gerVB) was used to transform strain GC614 to Tcr, giving strain GC632. Similarly, plasmid pHT-(gerUDM10stop gerUA gerUC gerKDbm gerVB) was used to construct strain GC634. A similar approach was used to prepare a plasmid in which the gerKDbm ORF was located between gerUC and gerVB in a gerU* operon with the intact gerUD gene, serving as the basis of strain GC633. The Gibson Assembly technique was also used to create a pHT315-based plasmid encoding a modified gerKbm operon, in which the gerKDbm ORF was replaced with the gerUD ORF. Essentially, PCR fragments encoding (i) gerKAbm and gerKCbm ORFs plus upstream promoter sequence, (ii) the gerUD ORF, and (iii) the gerKBbm ORF plus 200 bp of downstream sequence were amplified from B. megaterium QM B1551 genomic DNA. The PCR fragments, plus BamHI-linearized pHT315t vector, were purified and then assembled with Gibson Assembly master mix. Plasmid pHT-(gerKAbm gerKCbm gerUD gerKBbm) was isolated from E. coli, verified by PCR and sequencing, and used to transform B. megaterium GC614 to Tcr, giving strain GC637.
B. megaterium strains with transcriptional fusions between the
E. coli lacZ gene and either the
gerU or
gerKbm A and D genes (genes encoding the A and D proteins) were prepared essentially as described previously (
17). PCR was used to amplify ∼500-bp DNA fragments starting at the predicted translational start sites for
gerUA,
gerUD,
gerKAbm, and
gerKDbm using gene-specific primers with 5′ extensions to create
attB-flanked PCR products compatible with Gateway cloning (Life Technologies Ltd., Paisley, United Kingdom) into pDONRtet (
17) entry plasmids. Entry plasmids were isolated from
E. coli and used in a series of LR reactions to create appropriate receptor gene-pNFd13-derived (
17) destination plasmids. pNFd13-derived plasmids were then introduced into
B. megaterium QM B1551 via polyethylene glycol (PEG)-mediated protoplast transformation. Colonies that had undergone homologous recombination, integrating the pNFd13-derived plasmid at the cloned locus and placing
lacZ under the control of the promoter of the designated receptor gene, were isolated after incubation on solid medium at 42°C. The correct construction of the various strains was confirmed by PCR.
Spore germination.
B. subtilis spores were germinated following heat shock (30 min at 75°C) and cooling on ice. Spores at an OD
600 of 0.5 were germinated for 2.5 h at 37°C in 200 μl of 25 mM K-HEPES buffer (pH 7.4) with nutrient germinants added at various concentrations and with duplicate samples measured at all germinant concentrations.
B. subtilis spore germination was monitored by measuring the release of the spores' large depot of dipicolinic acid (DPA) by inclusion of 50 μM TbCl
3 in all germination mixtures and measuring Tb-DPA fluorometrically in a multiwell plate reader as described previously (
21). For all germination experiments, rates of germination were determined in arbitrary units (AU) by determining the maximum rates of increases in Tb-DPA fluorescence and correcting the values for slight differences in the DPA contents in different spore preparations; these corrections were always ≤10%. In a few experiments, spores were germinated as described above but without Tb
3+ present from the initiation of germination. Instead, at various times after germination was initiated, aliquots of the germinating culture were centrifuged, the supernatant fluid was made 50 μM in TbCl
3 and Tb-DPA fluorescence was measured as described previously (
22). Spore germination was also routinely monitored at the end of germination incubations by phase-contrast microscopy. The total amount of DPA present in spores was assessed by Tb-DPA fluorescence after DPA had been released from spores by boiling (
21).
In addition to nutrient germinants that trigger spore germination via GRs, there are also several agents that trigger spore germination without either GR involvement or a heat shock requirement, including the cationic surfactant dodecylamine and the 1:1 chelate of Ca
2+ and DPA (CaDPA) (
1,
3).
B. subtilis spores germinated in dodecylamine at 45°C in 25 mM K-HEPES buffer (pH 7.4) plus 50 μM TbCl
3 with 1.2 mM dodecylamine and spores at an OD
600 of 0.5, and germination was assessed by measuring the percentage of maximum Tb-DPA fluorescence obtained as described above. Germination of
B. subtilis spores with CaDPA was at 23°C and in 60 mM CaDPA made to pH 7.5 with Tris base and with spores at an OD
600 of 2. Spore germination with CaDPA was monitored by examining ∼100 individual spores at various times of germination by phase-contrast microscopy, since germinated spores become phase dark.
For B. megaterium spore germination, concentrated spore suspensions (OD600 of ∼50) were heat shocked (60°C for 10 min for gerU-associated experiments; 75°C for 30 min for gerKbm-associated experiments) and cooled on ice immediately prior to conducting germination assays. The physiological basis for apparent GR-specific heat shock regimens in B. megaterium has yet to be determined. Germination of B. megaterium spores was monitored by recording the decrease in optical density of 300-μl aliquots of spores suspended at an initial OD600 of 0.4 to 0.5 in 5 mM Tris-HCl (pH 7.5), supplemented with typically 5% (wt/vol) beef extract (for gerKbm-associated experiments) or 10 mM glucose or proline (for gerU-associated experiments), using a Perkin-Elmer EnVision-Xcite multilabel plate reader fitted with a 600-nm photometric filter. Germination assays were conducted typically for 90 min at 37°C, with 10 s of orbital shaking performed prior to OD600 measurements taken every minute. Experiments were conducted in triplicate, with at least two spore preparations for each strain.
Kinetic values associated with germination of B. megaterium spores were obtained by incubating heat-shocked spores at 37°C in 5 mM Tris-HCl (pH 7.5) containing various concentrations (0.1 mM to 1,000 mM) of glucose or proline. Germination data were plotted in SigmaPlot 11.0 (Systat Software Inc.), and the slope of the linear portion of curves was used to determine apparent Km and Vmax values using the program's ligand-binding macro. Plots of germinant concentration versus percent spore germination were analyzed with the same program to determine approximate concentrations of germinant required to stimulate the germination of 50% of the spore population in 60 min [K0.5 germ].
Analytical methods.
For analysis of the levels of germination proteins of
B. subtilis spores, spores of various strains were prepared, purified, decoated, ruptured with lysozyme, and subjected to brief sonication treatment, giving a spore lysate; in some cases, the spore inner membrane fraction was isolated by differential centrifugation of the spore lysate (all methods described previously [
23–26]). The levels of germination proteins were determined by Western blotting analyses on equal amounts of both spore lysate and inner membrane protein as determined by analysis of samples run on gels by SDS-polyacrylamide gel electrophoresis (PAGE), and staining the gels with Coomassie blue as described previously (
23–27). The antisera used have been shown to be specific for various germination proteins, including the various GR subunits, as well as the GerD protein essential for rapid GR-dependent spore germination and the SpoVAD protein essential for DPA movement into and out of spores (
2,
3).
For measurement of
gerKDbs-lacZ expression during sporulation,
B. subtilis strains were sporulated at 37°C by resuspension in Sterlini-Mandelstam medium (
28). At various times after resuspension (defined as time zero of sporulation), duplicate aliquots were centrifuged and washed with water, and pellets were frozen for subsequent analysis. β-Galactosidase activity in these samples was determined after lysozyme permeabilization, assaying equal aliquots of cell suspension using methylumbelliferyl-β-
d-galactoside (MUG) as the substrate in 1.2 ml, with incubation at 30°C and measuring methylumbelliferone production fluorometrically after 40 min as described previously (
4,
12). Note that this assay does not measure β-galactosidase in spores that have become lysozyme resistant, due to the assembly of the mature coat structure on the outer surfaces of the spores. DPA accumulation in the samples from the resuspension cultures was monitored by first boiling cells pelleted from 1 ml of culture in 1 ml of water, followed by centrifugation and measurement of DPA in supernatant fluids by its fluorescence with Tb as described previously (
21).
In a few cases, equal amounts (4 × 109) of dormant B. subtilis spores of either B. subtilis PS533 (wild type) or strains carrying gerA- or gerB-lacZ fusions with or without gerKDbs overexpression were prepared on plates. The spores were purified, decoated, and permeabilized with lysozyme, and 8 × 108 spores were assayed for β-galactosidase as described above.
With dormant B. megaterium spores, β-galactosidase activity was assayed in lysates from spores with lacZ transcriptional fusions to GR A and D genes. Typically, equal amounts of spores (1 ml at an OD600 of 10) were suspended in 0.6 ml of Z buffer (60 mM Na2HPO4, 40 mM NaH2PO4, 10 mM KCl, 1 mM MgSO4, and 50 mM β-mercaptoethanol [pH 7.0]) and then lysed by two rounds of shaking (instrument setting 6 for 30 s) in a FastPrep FP120 cell disrupter (Fisher Scientific, Loughborough, United Kingdom), with chilling on ice for 5 min between cycles. Spore lysates were recovered from the FastPrep tubes and centrifuged for 1 min at 15,000 × g, and 0.2 ml of 40 μg/ml MUG was added to the supernatant fluid. Reaction mixtures were incubated at 30°C for 40 min and then terminated by the addition of 0.4 ml of 1 M Na2CO3. Fluorescence associated with methylumbelliferone production was recorded in triplicate using a Tecan Infinite 200 series plate reader, using excitation and emission filters set at 365 nm and 450 nm, respectively.
For reverse transcriptase PCR (RT-PCR) to determine the time of expression of various GR A and D genes during B. megaterium sporulation, wild-type B. megaterium cells were cultured at 30°C in 200 ml SNB, and duplicate samples, adjusted to an OD600 of 10, were collected on an hourly basis following entry into stationary phase (designated time zero in sporulation), with this point determined by the plateauing of OD600 values recorded every 30 min during the growth phase of the culture. Samples were immediately centrifuged (15,000 × g for 1 min) and washed with RNAprotect bacterial reagent (Qiagen Ltd.), and the cell pellets were stored at −80°C until further analysis. RNA was subsequently extracted and purified from thawed cell pellets using an RNeasy minikit (Qiagen Ltd.), and then stored at −80°C. Approximately 1 μg of RNA from each sample was converted to cDNA using a QuantiTect reverse transcription kit (Qiagen Ltd., Manchester, United Kingdom), using random hexamers (Life Technologies Ltd., Paisley, United Kingdom) as primers for cDNA synthesis. Finally, cDNA samples for each time point served as templates for PCRs employing NovaTaq DNA polymerase (Merck Chemicals Ltd., Nottingham, United Kingdom), and gene-specific primers were designed to amplify ≤300-bp fragments of the genes of interest.