Skip to main content
  • Research article
  • Open access
  • Published:

Role of the gerA operon in L-alanine germination of Bacillus licheniformisspores

Abstract

Background

The genome of Bacillus licheniformis DSM 13 harbours three neighbouring open reading frames showing protein sequence similarities to the proteins encoded from the Bacillus subtilis subsp. subtilis 168 gerA operon, GerAA, GerAB and GerAC. In B. subtilis, these proteins are assumed to form a germinant receptor involved in spore germination induced by the amino acid L-alanine.

Results

In this study we show that disruption of the gerAA gene in B. licheniformis MW3 hamper L-alanine and casein hydrolysate-triggered spore germination, measured by absorbance at 600 nm and confirmed by phase contrast microscopy. This ability was restored by complementation with a plasmid-borne copy of the gerA locus. Addition of D-alanine in the casein hydrolysate germination assay abolished germination of both B. licheniformis MW3 and the complementation mutant. Germination of both B. licheniformis MW3 and the gerA disruption mutant was induced by the non-nutrient germinant Ca2+-Dipicolinic acid.

Conclusions

These results demonstrate that the B. licheniformis MW3 gerA locus is involved in germination induced by L-alanine and potentially other components present in casein hydrolysate.

Background

Germination of dormant Bacillus spores and subsequent outgrowth can be induced by various nutrients (amino acids, purine nucleosides, sugars, ions and combinations of these) recognised by receptor proteins encoded by the gerA family operons [1–3] and located in the inner membrane of the spore [4–7]. One or several germination receptor operons have been detected in the genomes of almost all spore formers, and supported by studies of different mutants it has been concluded that spores respond to germinants via receptors diverged from common ancestor(s) ([6] and references therein). Studies of receptor/germinant interactions have so far mainly been focusing on species belonging to Bacillus cereus, Bacillus subtilis, Bacillus megaterium and Bacillus anthracis [3, 8–16]. Bacillus licheniformis, another Gram-positive, spore forming soil bacterium closely related to B. subtilis [17], has on the other hand gained much less attention. B. licheniformis is a frequent contaminant of foods, and is a common spoilage organism of dairy products [18–20], bread [21, 22], packaged meats [23] and canned goods [24]. It has previously been considered non-pathogenic, and has been widely used in the industry for production of enzymes, antibiotics and biochemicals [25–27]. However, B. licheniformis-associated bovine abortion [28, 29], implant infection [30], corneal ulcer [31], bacteraemia sepsis [32] and food poisoning [33, 34] raise the question of its pathogenic potential. Some strains of B. licheniformis associated with human disease are capable of producing lichenysin A, a surfactin-like toxin [34, 35]. Due to its association with food-borne illness and spoilage, and its ability to undergo sporulation, [17, 36–38], extended knowledge about the germination apparatus of B. licheniformis is of general interest. To ensure microbiological safe food production of durable foods produced by relatively mild heat treatment, there is an obvious need for more information on spore forming bacteria.

Based on existing literature, B. subtilis could be considered as the model organism for germinant receptor studies. It was through early studies of germination defective mutants, that the theory of a L-alanine-induced germinant receptor was proposed [8]. Later studies identified the gerA locus as a tricistronic operon weakly expressed during sporulation, and that the polypeptide products of gerA probably formed a membrane associated complex [39–41]. The products of each of the three genes of gerA were later named GerAA, GerAB and GerAC, and were demonstrated to be simultaneously required for the spore to respond to L-alanine as sole germinant [2]. Genome sequence analysis and germination experiments of different mutants further identified four other tricistronic gerA homologs for B. subtilis; gerB, gerK, yndDEF and yfkQRT [10]. Receptors encoded by two of these operons, gerB and gerK, are confirmed functional when acting cooperatively with each other or with gerA [10, 15].

Homologous genes of germinant receptors belonging to the gerA family have been found in most spore formers, although the exact number, organisation and corresponding response germinant may vary for different species and even strains [3, 42, 43]. B. licheniformis ATCC 14580 is also predicted to possess potential germinant receptor proteins belonging to both the GerA and the GerK clades [44]. The GerAA, GerAB and GerAC protein sequences of B. licheniformis ATCC14580 are closely related to the protein sequences of the corresponding germinant receptor subunits of Bacillus subtilis subsp. subtilis 168. These are in B. subtilis encoded by the gerA operon, gerAA, gerAB and gerAC. Since B. subtilis gerA germination is triggered by L-alanine [2, 15], it is plausible that the B. licheniformis gerA operon also is involved in L-alanine germination. It has earlier been documented that spores of B. licheniformis from different strains actually respond to L-alanine as germinant [45–47], but to our knowledge, there are no functional studies of receptor/germinant interactions of strains belonging to B. licheniformis.

Mutational studies of B. licheniformis, including the fully sequenced B. licheniformis ATCC 14580/DSM 13 strain [48, 49], have long been a challenge, most likely due to their possession of a restriction apparatus destroying foreign DNA [48–50]. The construction of a more easily transformable mutant, B. licheniformis MW3, has largely overcome this challenge [50].

In order to facilitate the understanding of germinant/receptor interactions in B. licheniformis, we have constructed disruption and complementation mutants of the gerAA locus in B. licheniformis MW3. Spores of these mutants have been studied in germination assays with L-alanine, casein hydrolysate and the non-nutrient germinant Ca2+Dipicolinic acid (Ca2+DPA). These studies reveal that gerA is a main germinant receptor complex of B. licheniformis recognising amino acid(s), and supports the view that L-alanine is an important nutrient-germinant for this species.

Results and Discussion

Construction of the disruption and complementation mutants

To elucidate the role of the hypothetical GerA proteins during spore germination, a disruption mutant of the gerAA locus in B. licheniformis MW3 was constructed. B. licheniformis MW3 was used as target strain due to its superior transformability compared to its fully sequenced parent strain DSM 13 [50]. The gerAA mutant, NVH-1307, was constructed so that a part of the gerAA gene was substituted with a spectinomycin resistance cassette. This will cause the mutant to acquire spectinomycin resistance, and in addition, affect a potential phenotype related to the disrupted gene. If the target gene is part of an operon, which is the case of gerAA, downstream transcripted genes will also be affected, and the receptor non functional. Sequence analysis showed that in addition to harbouring the spectinomycin cassette in the gerAA locus, NVH-1307 also harboured two additional mutations (one base substitution and one base deletion) in the gerAA locus. These mutations were most likely acquired during PCR amplification of the fragments used to construct the disruption vector (pMAD_SpRΔgerAA). These mutations were "accepted" (not corrected) due to their location in the gene targeted for disruption. However, in construction of the plasmid used for gerAA complementation, a polymerase with a higher expected fidelity was applied to limit the risk of such mutations. Sequence analysis of the complementation plasmid pHT315_MW3gerA revealed no mutations in the amplified gerA operon when compared to the sequence of Veith et al.[48].

Genetic modification studies have shown that the germination rates could be significantly increased when specific germinant receptors are over-expressed in B. subtilis [51]. Thus, expression of germinant receptors is apparently not optimised for maximal spore germination, forwarded as a possible evolutionary strategy to prevent premature germination at nutrient conditions inadequate for sustained vegetative growth [3]. Very high levels of receptor expression could on the other hand have a negative effect on the sporulation process [51]. In such cells, the forespore lyses during the process of sporulation, perhaps as a result of premature forespore germination [51]. An appropriate evolutionary adaptation of germinant receptor expression/regulation is thus crucial to allow the cyclic transition between sporulation and germination upon environmental changes. In the construction of the complementation mutants in our study, certain precautions were therefore taken to avoid extensive over-expression of the complemented germinant receptor genes. By including some of the flanking regions of the gerAA, gerAB and gerAC fragment in the complementation plasmid, we wanted to maintain the native regulatory elements of this locus. In addition, a shuttle-vector with an expected low or moderate copy number was sought as a basis for the complementation plasmid. To our knowledge, there is no shuttle-vector available for B. licheniformis where the copy number is demonstrated to be low or moderate. However, Arantes and Lereclus [52] have constructed the pHT315 E. coli/B. thuringiensis shuttle-vector, with a copy number of ~ 15 per equivalent B. thuringiensis chromosome. This vector has successfully been used in germinant receptor complementation studies in B. megaterium [53], and was thus considered as a reasonable choice for B. licheniformis. Despite that this vector has shown to be stably maintained in B. thuringiensis and B. megaterium without a selective pressure [52, 54], the antibiotic erythromycin had to be included to ensure persistence of the complementation plasmid during sporulation of the B. licheniformis complementation mutant NVH-1311. This could be due to a different segregation stability of the vector in B. licheniformis. Another possibility is that there is a potential elevated risk of plasmid curing due to sporulation at a high temperature. Sporulation of B. licheniformis MW3, NVH-1307 and NVH-1311 were performed at 50 °C since a pilot study showed that sporulation at this temperature was faster, yielded more stable spores (less spontaneous germination) and a higher percentage of phase bright spores (results not shown).

Disruption of gerAA abolish L-alanine and casein hydrolysate induced germination

Decrease in absorbance at ~ 600 nm (A600) is used as a convenient method to monitor and compare germination of different spore populations [55, 56]. A fall in absorbance reflects a change in the refractive index (light scattering) of the multiple individual spores in a suspension, associated with germination events such as the excretion of spore's depot of Ca2+-DPA, followed by water influx, cortex degradation and core swelling [51, 56–59]. Figure 1 shows a representative experiment where different strains of heat activated (65 °C 20 min) spores (in Phosphate buffer) are supplemented with the germinant L-alanine. At these conditions, a clear change in absorbance was observed for spores of wild type (MW3) and wild type complementation mutant (NVH-1311) supplemented with L-alanine. Less than a 5%/h decrease in absorbance was observed for spores of the disruption mutant (NVH-1307). Phase-contrast images captured at the end of the germination assay (Figure 2), a technique where germinated and dormant spores appear dark and white/bright respectively [56, 59, 60], supports the absorbance measurements showing that spores of the wild type (MW3) and wild type complementation mutant (NVH-1311) germinate with L-alanine. No apparent increase in number of phase dark spores was observed for spores of the deletion mutant (NVH-1307) supplemented with L-alanine, or the negative controls. Together with the absorbance measurements, this shows that the introduced disruption of the gerAA gene abolishes the ability of B. licheniformis MW3 to use L-alanine as a germinant. The fact that the NVH-1311 complementation mutant showed a similar L-alanine triggered germination phenotype as the wild type spores, supports the hypothesis that an undisrupted copy of the gerAA, gerAB and gerAC genes, with flanking elements, are required for normal germination of B. licheniformis MW3 at these conditions. These findings were also supported by experiments performed with an alternative germination buffer; 50 mM Tris HCl pH 7.4 10 mM KCl (E. Klufterud, C. From; unpublished results).

Figure 1
figure 1

Germination of B. licheniformis with L-alanine. Germination is followed as a change in initial absorbance at 600 nm (A600) of phase bright spores in K-phosphate buffer pH 7.2 at 30 °C after addition of 100 mM L-alanine. Complete germination (>99% phase dark spores as observed by phase contrast microscopy) was observed at ~40% of initial A600. The results shown are representative of experiments performed in duplicate on two individual spore batches repeated at least twice.

Figure 2
figure 2

Phase contrast images of B. licheniformis spores following L-alanine germination. Phase contrast images (100 x) showing B. licheniformis spores after 3 hours germination at 30 °C with 100 mM L-alanine or negative control (MQ) in K-phosphatebuffer pH 7.2. The displayed images are representative of experiments performed in duplicate on two individual spore batches repeated at least twice.

An earlier study where germination in seven strains of B. licheniformis was investigated, showed that out of 24 amino acids tested, only L-alanine, L-cysteine and L-valine markedly stimulated germination [46]. In general, a greater germination response with L-alanine than with L-cysteine and L-valine was observed [46]. To assay the germination response of MW3, NVH-1307 and NVH-1311 to several amino acids, casein hydrolysate was used. Casein hydrolysate consists of a mixture of amino acids made from acid hydrolyzation of the milk protein casein and has been used as a germinant for Clostridium bifermentans and B. cereus in earlier studies [61–63]. In our study, casein hydrolysate proved to be a potent germinant for B. licheniformis, giving a rapid germination response (~70% phase dark spores as visualised by phase contrast microscopy) both for the wild type MW3 and the complementation mutant NVH-1311. The mutant NVH-1307, which most likely lacks a gerA receptor, did not show any germination response in casein hydrolysate (Figure 3).

Figure 3
figure 3

Germination of B. licheniformis with casein hydrolysate. Germination is followed as a change in initial absorbance at 600 nm (A600) of phase bright spores in Tris HCl buffer pH 7.4 at 30 °C after addition of 1% (w/v) casein hydrolysate. Complete germination (>99% phase dark spores as observed by phase contrast microscopy) was observed at ~40% of initial A600. The results shown are representative of experiments performed in duplicate on two individual spore batches repeated at least twice.

D-alanine is a well-known inhibitor of L-alanine germination of B. subtilis and B. licheniformis [64, 65, 46, 15, 66]. D-alanine has also been shown to reduce L-valine induced germination of B. subtilis [15, 66], but we are not aware of studies reporting the effect of D-alanine on L-valine induced germination of B. licheniformis. In order to abolish germination by L-alanine present in the casein hydrolysate, we added D-alanine in some of the above experiments. In these experiments, the germination response of both MW3 and NVH-1311 was hardly measurable (results not shown), indicating that L-alanine through its triggering of the gerA receptor is an important germinant of B. licheniformis. The contribution to germination of the remaining amino acids in the casein hydrolysate when D-alanine was present, appear to be minimal. Although one can not rule out that D-alanine also inhibits the effect of other amino acids present in casein hydrolysate (e.g. L-valine), all the findings support the view that gerA and L-alanine constitute one of the main germination pathways of B. licheniformis.

Germination of B. licheniformis with Ca2+-DPA

In order to by-pass the spore germination receptor apparatus, experiments using exogenous Ca2+-DPA to trigger germination of spores of B. licheniformis MW3 and the mutant strain NVH-1307 were performed. In B. subtilis spores, Ca2+-DPA induced germination is believed to act through activation of the cortex lytic enzyme CwlJ, without any requirement of functional germinant receptors [10, 67]. Bioinformatic analysis of complete genomes of different spore formers has shown that also B. licheniformis contains a B. subtilis homologous cwlJ gene [43]. If the germination apparatus of B. licheniformis spores is similar to that of its close relative B. subtilis, the wild type and disruption mutant of B. licheniformis should exhibit a similar germination response as B. subtilis to exogenous Ca2+-DPA. The DPA concentration needed to trigger germination in B. subtilis is ~ 20 - 60 mM, supplemented together with equal (or excess) amounts of Ca2+ (allowing formation of a 1:1 chelate of calcium and dipicolinic acid) [10]. Also spores of B. cereus and B. megaterium germinate when exposed to Ca2+-DPA [68, 69]. For B. cereus it has been shown that a final level of 60 mM Ca2+-DPA is sufficient to ensure germination [69]. In our experiments (Table 3), B. licheniformis spores of MW3, the mutant NVH-1307 and B. subtilis spores of strain B252 (used as a positive control) germinated effectively after 3 hours exposure in room temperature at a final concentration of 80 mM DPA and 100 mM CaCl2. Further, at 45 mM DPA 50 mM CaCl2 spores of B. cereus ATCC 14579 germinated effectively whilst spores of B. subtilis strain B252 showed a moderate germination response. B. licheniformis MW3 and NVH-1307 exhibited a weak germination response even after a prolonged exposure of ~21 h at these concentrations. At 20 mM DPA 30 mM CaCl2 B. cereus ATCC 14579 germinated moderately whilst spores of MW3, NVH-1307 and B. subtilis B252 did not germinate (Table 3). Earlier Ca2+-DPA germination studies with other B. licheniformis strains in our collection have yielded similar results with less effective Ca2+-DPA induced germination compared to B. cereus ATCC 14579 and spores of B. pumilus (results not shown). Reasons for a reduced sensitivity to Ca2+-DPA as a non-nutrient germinant in B. licheniformis MW3 spores compared to spores of some other spore forming bacteria is unknown. It might be that the relationship between Ca2+ and DPA or the concentration of the chelate is not ideal for B. licheniformis germination. Another possibility is that a so far uncharacterised non-nutrient inducing germinant or a mixture of DPA with other ions than Ca2+ is needed for effective CwlJ mediated germination of B. licheniformis. It has been shown in earlier studies that for instance strains of B. megaterium also germinate in mixtures with other ions than Ca2+ [70]. More information on CwlJ and other enzyme interactions with Ca2+-DPA is needed to get a clear view on which mechanisms form the basis for the different effects of Ca2+-DPA germination in B. licheniformis, B. cereus and B. subtilis. Further characterisation of Ca2+-DPA dependent germination of B. licheniformis is currently carried out by our group.

Conclusions

As demonstrated by genetic mutation and complementation analysis, this study reveals that the gerAA gene in B. licheniformis MW3 has a fundamental role in germination triggered by L-alanine and casein hydrolysate. We also show that D-alanine is an important inhibitor in B. licheniformis amino acid-induced germination. Further, both wild type and the gerAA disruption mutant germinated effectively when exposed to appropriate levels of the non-nutrient germinant Ca2+-DPA which by-pass the spore receptor apparatus. However, effective germination with Ca2+-DPA seems both strain and species specific. In order to understand and potentially control the germination behaviour of B. licheniformis spores, disclosure of factors involved in the transition from a dormant spore to a metabolically active proliferating cell is of prime importance. Although complete elucidation of the function and cooperation of the different germinant receptors are rather laborious, this study has taken a step in the direction of obtaining more knowledge about this less studied species. The availability of both disruption and complementation mutants will facilitate further research on the function of the GerA receptor of B. licheniformis MW3, as well as its potential involvement in germination triggered by alternative nutrients and cooperation with other germinant receptors. Further bioinformatic and phenotypic investigations are in progress in our laboratory and might eventually provide insight relevant for improved spore decimation techniques by the use of induced germination.

Methods

Bacterial strains and DNA extraction

The strains used in this study were B. licheniformis MW3 [50], B. subtilis B252 [71] and the B. cereus type-strain ATCC 14579 [72, 73] (Table 1). B. licheniformis MW3 is a mutant created from B. licheniformis DSM13 (isogenic to ATCC 14580) with targeted deletions of the hsdR loci of two type I restriction modification systems making the strain readily transformable. B. licheniformis MW3 was used as host for creating disruption and complementation mutants of the gerA locus. When not stated otherwise, bacteria were cultured at 37 °C on LB agar or broth containing appropriate selective antibiotics (Table 1). Genomic DNA for PCR amplifications and sequencing was extracted from B. licheniformis MW3 and B. licheniformis NVH-1307 by a method slightly modified from [71], as follows. An overnight culture was transferred to fresh growth medium and grown at 37 °C, 225 rpm (HT-Infors AG CH-4103, Bottmingen, Switzerland), to turbidity (4-5 h). Cells from 1 ml culture was harvested by centrifugation (3 min at 16.100 × g), and the pellet was frozen at -20 °C. Thawed pellet was resuspended in 495 µl SET buffer (75 mM NaCl, 25 mM EDTA, 20 mM Tris, pH 7.5) and 50 µl 10 mg/ml lysozyme before incubation at 37 °C for 1 h. Further, 50 µl 10% sodium dodecyl sulfate and 5 µl 25 mg/ml proteinase K was added, and the sample was incubated at 50 °C for 2 h. At room temperature (RT), the sample was mixed with 200 µl 5 M NaCl and 700 µl of chloroform-isoamyl alcohol (24:1), and incubated with frequent inversions for 30 min. The aqueous phase was separated by centrifugation (20-30 min at 16.100-20.800 × g), transferred to a fresh tube, and DNA was precipitated by addition of an equal volume of isopropanol followed by centrifugation (20 min at 16.100-20.800 × g). The precipitate was washed with 70% ethanol and centrifuged (15 min at 16.100-16.500 × g), and the supernatant was removed before the precipitate was left to air dry. DNA was resuspended in 100 µl 10 mM Tris HCl buffer (pH 8.5). Plasmid DNA was purified according to the manual provided with the Plasmid Mini/Midi kits (QIAGEN®).

Table 1 Strains and plasmids used in this study

Preparation and transformation of B. licheniformiselectrocompetent cells

Electrocompetent B. licheniformis was prepared and transformed by a modified version of the protocol described by Mahillion et al.[74] as follows. A preculture in Brain Heart Infusion broth (BHI) (Oxoid, Cambridge, United Kingdom) was grown overnight at 37 °C, and 1 ml was used to inoculate 200 ml pre-warmed BHI in a 1 l Erlenmeyer. The culture was incubated 4 to 5 h at 37 °C and 150 rpm (HT-Infors AG CH-4103, Bottmingen, Switzerland) until A600 of 0.9-1.0 was reached (Shimadzu UV-VIS 160A, Shimadzu Europa GMBH). Cells were pelleted and washed twice with 200 ml RT autoclaved MilliQ water (MQ) by 15 min centrifugations at 3.300 and 10.400 × g. The pellet was resuspended in a 10 ml filter sterilised solution of freshly prepared polyethylene glycol (PEG) 6000 (Merck, Darmstadt, Germany), made by dissolving 40 g PEG6000 in 100 ml MQ. Following 15 min centrifugation at 4.080 × g, cells were resuspended in 0.5-1 ml of the PEG6000/MQ solution, aliquoted (100 µl) and stored at -80 °C.

Transformation was conducted by adding 2 µl plasmid to 100 µl electro competent cells thawed on ice. Following ~1 min incubation on ice, electroporation was performed at 1.4 to 2.5 kV (Eppendorf Eporator, Eppendorf AG, Hamburg, Germany or MicroPulser™, Bio-Rad, Hercules, CA), using 0.2 cm gap width electroporation cuvettes (Bio-Rad Laboratories, Hercules, CA). Before plating on selective LB-agar plates, cells were recovered in LB or S. O. C. medium (Invitrogen) at 37 °C, 150 rpm, for 4 to 5 h.

Construction of B. licheniformis MW3ΔgerAA::spc

The shuttle vector used for construction of a spectinomycin resistant (SpR) insertion deletion in the gerAA was pMAD_SpR. This vector has been modified from pMAD [75] by insertion of a (SpR)-cassette in the restriction site SalI [76]. As selective antibiotics for the presence of pMAD_SpR or its derivative constructs, 100 µg/ml ampicillin and 100 µg/ml spectinomycin was used for E. coli TOP10 growth, and 3 µg/ml erythromycin and 250-300 µg/ml spectinomycin for B. licheniformis growth. This vector carries a constitutively expressed β-galactosidase gene, allowing blue-white screening on plates spread with X-Gal (40 µl 40 mg/ml 5-bromo-4-chloro-3-indolyl-β-D-galactopyranoside, VWR, BDH Prolabo). This screening was, however, not always unambiguous following long incubations of plates with B. licheniformis MW3 transformants, probably due to the natural precence of β-galactosidase in B. licheniformis DSM 13 [77]. To construct the gene replacement vector, primers (Table 2) were designed to amplify two DNA fragments, one homologous to upstream (709 bp) and one to downstream (696 bp) regions of the deletion target (567 bp) in the gerAA. Platinum Taq DNA Polymerase High Fidelity kit (Invitrogen) was used for PCR amplification with the following amplification procedure: initial denaturation for 2 min at 94°C, 30 cycles of 30 s at 94 °C, 30 s at 50 °C and 1 min at 68 °C, and final extension at 68 °C for 10 min. Primers of the upstream and downstream amplicons contained restriction sites BamHI and EcoRI respectively (Table 2), allowing a two_step ligation into the corresponding restriction sites on either side of the (SpR)-cassette in pMAD_SpR. The resulting gene replacement plasmid, pMAD_SpRΔgerAA, was controlled for correct orientation of the upstream and downstream fragments by PCR. pMAD_SpRΔgerAA was introduced into B. licheniformis MW3 by electroporation, and allelic exchange of internal parts of gerAA (567 bp) with the (SpR)-cassette of pMAD_SpRΔgerAA was allowed by double crossover. The protocol was performed as described by Arnaud et al.[75], except using growth temperatures of 37 °C following initial transformation, an incubation temperature of 45 °C and spectinomycin present during plasmid curing, and an incubation temperature of 37 °C when screening for the double crossover phenotype (spectinomycin resistant and erythromycin sensitive colonies). Chromosomal DNA was purified from a candidate colony and used in PCR amplifications (as described above) with primers hybridizing outside the cloned DNA fragment and inside the spectinomycin cassette (Table 2) to verify the deletion and insertion by sequencing. The disruption mutant was named B. licheniformis MW3ΔgerAA::spc (NVH-1307) and used in the following complementation, sporulation and germination assays.

Table 2 Primers used in this study
Table 3 Ca2+-DPA germination in selected Bacillus spp

Construction of gerAcomplementation mutant

The shuttle vector used as base for trans complementation of ΔgerAA::spc was pHT315 [52]. As selective antibiotics for presence of pHT315 constructs, 100 µg/ml ampicillin was used for E. coli TOP10 growth, while 1 µg/ml erythromycin for B. licheniformis growth. Spectinomycin (250 µg/ml) was also supplemented for the chromosomal antibiotic resistance of NVH-1307 mutants carrying pHT315 derivatives (Table 1). To construct pHT315 complementation plasmids that harbour the gerA operon, DNA isolated from B. licheniformis MW3 was used. Primers, with SalI and XbaI restriction sites (Table 2) were designed to amplify a 3982 bp fragment spanning from 151 bp upstream to 177 bp downstream gerA (gerAA, gerAB, gerAC). To ensure high fidelity, PCR amplification was performed with Phusion Hot Start II High-Fidelity DNA Polymerase kit (Finnzymes). The amplification protocol was as follows; initial denaturation for 30s at 98°C, 30 cycles of 10 s at 98 °C, 30 s at 58 °C and 2 min at 72 °C, and final extension at 72 °C for 10 min. The amplified fragments were cloned into the SalI/XbaI restriction site of pHT315, giving the complementation plasmid pHT315_MW3gerA. The purified plasmid was controlled by sequencing using primers hybridizing to pHT315 and internal gerA. The verified plasmid was introduced into the disruption mutant (NVH-1307) by electroporation as described earlier, giving the strain B. licheniformis MW3ΔgerAA::spcpHT315_MW3gerA (NVH-1311). The strain was used in sporulation and germination assays.

Sporulation

Sporulation was performed by a modified version of the sporulation protocol and medium described by van der Voort [42] as outlined below. Bacteria were pre-cultivated for 5 to 6 h in 50 ml LB-Broth with agitation (225 rpm) at 50 °C. Pre-culture of NVH-1307 was supplemented with 250 µg/ml spectinomycin, while the culture of NVH-1311 was supplemented with 250 µg/ml spectinomycin and 1 µg/ml erythromycin. Twenty µl of pre-culture was added to 100 ml sporulation medium, containing 8 g of nutrient broth (Difco, Becton, Dickinson and Company, NJ, USA) per liter, 1 μM FeSO4·7H2O (Merck KGaA, Darmstadt, Germany), 2.5 μM CuCl2·2H2O (Sigma-Aldrich, Steinheim, Germany), 12.5 μM ZnCl2 (Sigma-Aldrich, Steinheim, Germany), 66 μM MnSO4·4H2O (BDH Prolabo, VWR International AS, Oslo, Norway), 1 mM MgCl2·6H2O (J. T. Baker Chemicals B. V., Deventer, Holland), 5 mM (NH4)2SO4 (Merck KGaA, Darmstadt, Germany), 2.5 µM Na2MoO4·2H2O (Riedel-de Häen, Sigma-Aldrich, Seelze, Germany), 2.5 µM CoCl2·6H2O (Sigma-Aldrich, Steinheim, Germany) and 1 mM Ca(NO3)2·4H2O (Merck KGaA, Darmstadt, Germany). Filter sterilised Ca(NO3)2·4H2O, MnSO4·4H2O and FeSO4·7H2O were added to the medium after it had been autoclaved. pH was adjusted to 7.6 before autoclaving, and the pH of the final sporulation medium was 7.2. Sporulation medium of NVH-1311 was supplemented with 1 µg/ml erythromycin. The cultures were incubated with agitation (225 rpm) at 50 °C for 1 to 2 days for B. licheniformis strains MW3, NVH-1307 and NVH-1311, or for 2 days at 30 °C for B. subtilis B252 and B. cereus ATCC 14579 until ≥90% phase bright spores as judged by phase contrast microscopy. Spores were harvested by centrifugation for 10 min at 6000 × g at 4 °C, and resuspended in 10 ml cold autoclaved MQ. Washing of spores was done by centrifugation and resuspension in MQ a total of ten times. The resulting spore crops, < 10% germinated spores, were stored refrigerated in MQ. When used in the following germination studies, spore crops were between 2 and 7 months old.

Germination assays

The spore suspension was routinely washed by centrifugation at 16.100 × g for 3 min prior to each experiment [78]. Spores were heat activated in MQ at 65 °C for 20 min, chilled on ice, centrifuged (16.100 × g for 3 min) and resuspended in 2 × germination buffer (100 mM K-phosphate buffer pH 7.2) for L-alanine germination or 1 × germination buffer (50 mM Tris HCl pH 7.4 10 mM KCl) for germination with casein hydrolysate (Merck, Darmstadt, Germany). Casein hydrolysate consists of a mixture of different amino acids (Merck Microbiology Manual 12th Edition: typical amino acid content (% w/w); alanine (2.00), arginine (2.20), aspartic acid (4.40), glutamic acid (12.50), glycine (1.20), histidine (1.80), isoleucine (2.40), leucine (3.40), lysine (5.60), methionine (1.20), phenylalanine (2.50), proline (6.10), serine (2.70), threonine (2.20), tyrosine (0.60), valine (3.90)) made from acid hydrolyzation of the milk protein casein. Germination was followed as described by Hornstra et al.[13] by monitoring the reduction in absorbance at A600 as spores turn from phase-bright to phase dark at 30 °C in a 96-well microplate in a plate reader (Tecan Intinite M200, Grödig, Austria). The spore suspension was adjusted to an initial A600 of ~2 (Shimadzu UV-VIS 160A, Shimadzu Europa GMBH) prior to addition of germinant. Germinant (filter sterilised L-alanine dissolved in MQ or casein hydrolysate dissolved in 50 mM Tris HCl pH 7.4 10 mM KCl) or negative control (MQ for L-alanine germination and 50 mM Tris HCl pH 7.4 10 mM KCl for casein hydrolysate germination) was automatically injected, and the plate was shaken for 10 s prior to the first reading. A600 was recorded every 30 s for 142 to 170 min, with 10 s shaking in-between each measurement. The final concentration of germination buffer was 50 mM phosphatebuffer pH 7.2 or 50 mM Tris HCl pH 7.4 10 mM KCl, and final concentration of germinant was 100 mM L-alanine or 1% (w/v) casein hydrolysate. The final concentration of spores gave an initial A600 of ~0.7-0.8. To inhibit germination with L-alanine and potential other amino acids in the casein hydrolysate germination assay, 0.2% D-alanine (w/v, final concentration) was in some experiments added to each test well. The germination progress was described as the percentage of the initial A600 (% A600i) for each measurement point [13]. All experiments were performed in duplicates on two individual spore batches and repeated at least twice. Germination was routinely controlled by phase-contrast microscopy (Olympus BX51, Hamburg, Germany) [13].

Spore germination in Ca2+-DPA was performed as follows; spores were washed in cold autoclaved MQ and resuspended in germination buffer (125-250 mM Tris base, 25-100 mM DPA (2,6-Pyridinedicarboxylic acid 99%, Sigma-Aldrich, Steinheim, Germany) pH ~8) [79]. Germination was initiated by addition of excess CaCl2·2H2O (Riedel de Häen AG, Seelze, Germany), followed by incubation for 3 h with shaking at room temperature (~20°C). The final concentrations of Ca2+-DPA in the assay were 20-80 mM DPA 30-100 mM CaCl2, and the final concentration of spores gave an initial A600 of ~0.6-0.8. Germination was described as an approximate percentage of phase dark spores after screening of microscopic slides by phase contrast microscopy (100 x). Experiments were performed in duplicate on two individual spore batches and repeated at least twice.

DNA sequencing and bioinformatics

DNA sequencing was performed by GATC Biotech (Konstanz, Germany) or Source BioScience (Nottingham, United Kingdom). The genomic sequence of B. licheniformis DSM13 [48] was accessed at http://www.ncbi.nml.nih.gov [GenBank: AE017333].

References

  1. Setlow P: Spore germination. Curr Opin Microbiol. 2003, 6: 550-556. 10.1016/j.mib.2003.10.001.

    Article  PubMed  CAS  Google Scholar 

  2. Moir A, Smith DA: The genetics of bacterial spore germination. Ann Rev Microbiol. 1990, 44: 531-553. 10.1146/annurev.mi.44.100190.002531.

    Article  CAS  Google Scholar 

  3. Ross C, Abel-Santos E: The ger receptor family from sporulating bacteria. Curr Issues Mol Biol. 2010, 12: 147-157.

    PubMed  CAS  PubMed Central  Google Scholar 

  4. Hudson KD, Corfe BM, Kemp EH, Feavers IM, Coote PJ, Moir A: Localization of GerAA and GerAC germination proteins in the Bacillus subtilis spore. J Bact. 2001, 183: 4317-4322. 10.1128/JB.183.14.4317-4322.2001.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  5. Paidhungat M, Setlow P: Localization of a germinant receptor protein (GerBA) to the inner membrane of Bacillus subtilis spores. J Bact. 2001, 183: 3982-3990. 10.1128/JB.183.13.3982-3990.2001.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  6. Moir A: How do spores germinate?. J Appl Microbiol. 2006, 101: 526-530. 10.1111/j.1365-2672.2006.02885.x.

    Article  PubMed  CAS  Google Scholar 

  7. Griffiths KK, Zhang JQ, Cowan AE, Yu J, Setlow P: Germination proteins in the inner membrane of dormant Bacillus subtilis spores colocalize in a discrete cluster. Mol Microbiol. 2011, 81: 1061-1077. 10.1111/j.1365-2958.2011.07753.x.

    Article  PubMed  CAS  Google Scholar 

  8. Sammons RL, Moir A, Smith DA: Isolation and properties of spore germination mutants of Bacillus subtilis 168 deficient in the initiation of germination. J Gen Microbiol. 1981, 124: 229-241.

    Google Scholar 

  9. Clements MO, Moir A: Role of the gerI operon of Bacillus cereus 569 in the response of spores to germinants. J Bact. 1998, 180: 6729-6735.

    PubMed  CAS  PubMed Central  Google Scholar 

  10. Paidhungat M, Setlow P: Role of ger proteins in nutrient and nonnutrient triggering of spore germination in Bacillus subtilis. J Bact. 2000, 182: 2513-2519. 10.1128/JB.182.9.2513-2519.2000.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  11. Barlass PJ, Houston CW, Clements MO, Moir A: Germination of Bacillus cereus spores in response to L- alanine and to inosine: the roles of gerL and gerQ operons. Microbiology. 2002, 148: 2089-2095.

    Article  PubMed  CAS  Google Scholar 

  12. Ireland JAW, Hanna PC: Amino acid- and purine ribonucleoside-induced germination of Bacillus anthracis Delta Sterne endospores gerS mediates responses to aromatic ring structures. J Bact. 2002, 184: 1296-1303. 10.1128/JB.184.5.1296-1303.2002.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  13. Hornstra LM, de Vries YP, de Vos WM, Abee T, Wells-Bennik MHJ: gerR, a novel ger operon involved in L-alanine- and inosine-initiated germination of Bacillus cereus ATCC 14579. Appl Environ Microbiol. 2005, 71: 774-781. 10.1128/AEM.71.2.774-781.2005.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  14. Hornstra LM, de Vries YP, Wells-Bennik MHJ, de Vos WM, Abee T: Characterization of germination receptors of Bacillus cereus ATCC 14579. Appl Environ Microbiol. 2006, 72: 44-53. 10.1128/AEM.72.1.44-53.2006.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  15. Atluri S, Ragkousi K, Cortezzo DE, Setlow P: Cooperativity between different nutrient receptors in germination of spores of Bacillus subtilis and reduction of this cooperativity by alterations in the GerB receptor. J Bact. 2006, 188: 28-36. 10.1128/JB.188.1.28-36.2006.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  16. Christie G, Lowe CR: Role of chromosomal and plasmid-borne receptor homologues in the response of Bacillus megaterium QM B1551 spores to germinants. J Bact. 2007, 189: 4375-4383. 10.1128/JB.00110-07.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  17. Logan NA, De Vos P, et al: Genus I. Bacillus. Bergey's manual of systematic bacteriology. Edited by: De Vos P, Garrity GM, Jones D, Krieg NR, Ludwig W, Rainey FA. 2009, New York: Springer, 21-128.

    Google Scholar 

  18. Kalogridou-Vassiliadou D: Biochemical activities of Bacillus species isolated from flat sour evaporated milk. J Dairy Science. 1992, 75: 2681-2686. 10.3168/jds.S0022-0302(92)78030-8.

    Article  CAS  Google Scholar 

  19. Crielly EM, Logan NA, Anderton A: Studies on the Bacillus flora of milk and milk-products. J Appl Bacteriol. 1994, 77: 256-263. 10.1111/j.1365-2672.1994.tb03072.x.

    Article  PubMed  CAS  Google Scholar 

  20. Janstova B, Lukasova J: Heat resistance of Bacillus spp. spores isolated from cow's milk and farm environment. Acta Vet Brno. 2001, 70: 179-184. 10.2754/avb200170020179.

    Article  Google Scholar 

  21. Thompson JM, Waites WM, Dodd CER: Detection of rope spoilage in bread caused by Bacillus species. J Appl Microbiol. 1998, 85: 481-486. 10.1046/j.1365-2672.1998.853512.x.

    Article  Google Scholar 

  22. Sorokulova IB, Reva ON, Smirnov VV, Pinchuk IV, Lapa SV, Urdaci MC: Genetic diversity and involvement in bread spoilage of Bacillus strains isolated from flour and ropy bread. Lett Appl Microbiol. 2003, 37: 169-173. 10.1046/j.1472-765X.2003.01372.x.

    Article  PubMed  CAS  Google Scholar 

  23. Bell RG, Delacy KM: A note on the identity and properties of the spoilage microflora of chub-packed luncheon meat stored at ambient-temperature. Can J Microbiol. 1983, 29: 1220-1223. 10.1139/m83-187.

    Article  PubMed  CAS  Google Scholar 

  24. Fields ML, Zamora AF, Bradsher M: Microbiological analysis of home-canned tomatoes and green beans. J Food Science. 1977, 42: 931-934. 10.1111/j.1365-2621.1977.tb12640.x.

    Article  Google Scholar 

  25. Eveleigh DE: The microbiological production of industrial chemicals. Sci Am. 1981, 245: 120-130. 10.1038/scientificamerican0881-120.

    Article  Google Scholar 

  26. de Boer AS, Priest F, Diderichsen B: On the industrial use of Bacillus licheniformis - A review. Appl Microbiol Biotechnol. 1994, 40: 595-598. 10.1007/BF00173313.

    Article  Google Scholar 

  27. Schallmey M, Singh A, Ward OP: Developments in the use of Bacillus species for industrial production. Can J Microbiol. 2004, 50: 1-17. 10.1139/w03-076.

    Article  PubMed  CAS  Google Scholar 

  28. Agerholm JS, Krogh HV, Jensen HE: A retrospective study of bovine abortions associated with Bacillus licheniformis. J Vet Med Series B-Infectious Diseases and Veterinary Public Health. 1995, 42: 225-234. 10.1111/j.1439-0450.1995.tb00706.x.

    Article  CAS  Google Scholar 

  29. Syrjälä P, Anttila M, Dillard K, Fossi M, Collin K, Nylund M, et al: Causes of bovine abortion, stillbirth and neonatal death in Finland 1999-2006. Acta Vet Scan. 2007, 49: S3-10.1186/1751-0147-49-S1-S3.

    Article  Google Scholar 

  30. Santini F, Borghetti V, Amalfitano G, Mazzucco A: Bacillus licheniformis prosthetic aortic-valve endocarditis. J Clin Microbiol. 1995, 33: 3070-3073.

    PubMed  CAS  PubMed Central  Google Scholar 

  31. Tabbara KF, Tarabay N: Bacillus licheniformis corneal ulcer. Am J Ophthalmol. 1979, 87: 717-719.

    Article  PubMed  CAS  Google Scholar 

  32. Sugar AM, Mccloskey RV: Bacillus licheniformis sepsis. J Am Med Assoc. 1977, 238: 1180-1181. 10.1001/jama.1977.03280120072022.

    Article  CAS  Google Scholar 

  33. Kramer JM, Gilbert RJ: Bacillus cereus and other Bacillus species. Foodborne bacterial pathogens. Edited by: Doyle MP. 1989, New York: Marcel Dekker Inc, 21-70.

    Google Scholar 

  34. Salkinoja-Salonen MS, Vuorio R, Andersson MA, Kampfer P, Andersson MC, Honkanen-Buzalski T, et al: Toxigenic strains of Bacillus licheniformis related to food poisoning. Appl Environ Microbiol. 1999, 65: 4637-4645.

    PubMed  CAS  PubMed Central  Google Scholar 

  35. Mikkola R, Kolari M, Andersson MA, Helin J, Salkinoja-Salonen MS: Toxic lactonic lipopeptide from food poisoning isolates of Bacillus licheniformis. Eur J Biochem. 2000, 267: 4068-4074. 10.1046/j.1432-1033.2000.01467.x.

    Article  PubMed  CAS  Google Scholar 

  36. Errington J: Regulation of endospore formation in Bacillus subtilis. Nature Rev Microbiol. 2003, 1: 117-126. 10.1038/nrmicro750.

    Article  CAS  Google Scholar 

  37. Setlow P: Spores of Bacillus subtilis: their resistance to and killing by radiation, heat and chemicals. J Appl Microbiol. 2006, 101: 514-525. 10.1111/j.1365-2672.2005.02736.x.

    Article  PubMed  CAS  Google Scholar 

  38. Setlow P, Johnson EA: Spores and their significance. Food microbiology: fundamentals and frontiers. Edited by: Doyle MP, Beuchat LR. 2007, Washington, DC: ASM Press, 35-67.

    Chapter  Google Scholar 

  39. Zuberi AR, Feavers IM, Moir A: Identification of 3 complementation units in the gerA spore germination locus of Bacillus subtilis. J Bact. 1985, 162: 756-762.

    PubMed  CAS  PubMed Central  Google Scholar 

  40. Feavers IM, Miles JS, Moir A: The nucleotide sequence of a spore germination gene (gerA) of Bacillus subtilis 168. Gene. 1985, 38: 95-102. 10.1016/0378-1119(85)90207-0.

    Article  PubMed  CAS  Google Scholar 

  41. Zuberi AR, Moir A, Feavers IM: The nucleotide-sequence and gene organization of the gerA spore germination operon of Bacillus subtilis 168. Gene. 1987, 51: 1-11. 10.1016/0378-1119(87)90468-9.

    Article  PubMed  CAS  Google Scholar 

  42. van der Voort M, Garcia D, Moezelaar R, Abee T: Germinant receptor diversity and germination responses of four strains of the Bacillus cereus group. Int J Food Microbiol. 2010, 139: 108-115. 10.1016/j.ijfoodmicro.2010.01.028.

    Article  PubMed  CAS  Google Scholar 

  43. Paredes-Sabja D, Setlow P, Sarker MR: Germination of spores of Bacillales and Clostridiales species: mechanisms and proteins involved. Trends Microbiol. 2011, 19: 85-94. 10.1016/j.tim.2010.10.004.

    Article  PubMed  CAS  Google Scholar 

  44. Ross CA, Abel-Santos E: Guidelines for nomenclature assignment of Ger receptors. Res Microbiol. 2010, 161: 830-837. 10.1016/j.resmic.2010.08.002.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  45. Halmann M, Keynan A: Stages in germination of spores of Bacillus licheniformis. J Bact. 1962, 84: 1187-1193.

    PubMed  CAS  PubMed Central  Google Scholar 

  46. Martin JH, Harper WJ: Germination response of Bacillus licheniformis spores to amino acids. J Dairy Sci. 1963, 46: 663-667. 10.3168/jds.S0022-0302(63)89119-5.

    Article  CAS  Google Scholar 

  47. White CH, Chang RR, Martin JH, Loewenst M: Factors affecting L- Alanine induced germination of Bacillus spores. J Dairy Sci. 1974, 57: 1309-1314. 10.3168/jds.S0022-0302(74)85060-5.

    Article  PubMed  CAS  Google Scholar 

  48. Veith B, Herzberg C, Steckel S, Feesche J, Maurer KH, Ehrenreich P, et al: The complete genome sequence of Bacillus licheniformis DSM13, an organism with great industrial potential. J Mol Microbiol Biotechnol. 2004, 7: 204-211. 10.1159/000079829.

    Article  PubMed  CAS  Google Scholar 

  49. Rey MW, Ramaiya P, Nelson BA, Brody-Karpin SD, Zaretsky EJ, Tang M, et al: Complete genome sequence of the industrial bacterium Bacillus licheniformis and comparisons with closely related Bacillus species. Gen Biol. 2004, 5: R77-10.1186/gb-2004-5-10-r77.

    Article  Google Scholar 

  50. Waschkau B, Waldeck J, Wieland S, Eichstadt R, Meinhardt F: Generation of readily transformable Bacillus licheniformis mutants. Appl Microbiol Biotechnol. 2008, 78: 181-188. 10.1007/s00253-007-1278-0.

    Article  PubMed  CAS  Google Scholar 

  51. Cabrera-Martinez RM, Tovar-Rojo F, Vepachedu VR, Setlow P: Effects of overexpression of nutrient receptors on germination of spores of Bacillus subtilis. J Bact. 2003, 185: 2457-2464. 10.1128/JB.185.8.2457-2464.2003.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  52. Arantes O, Lereclus D: Construction of cloning vectors for Bacillus thuringiensis. Gene. 1991, 108: 115-119. 10.1016/0378-1119(91)90495-W.

    Article  PubMed  CAS  Google Scholar 

  53. Christie G, Gotzke H, Lowe CR: Identification of a receptor subunit and putative ligand-binding residues involved in the Bacillus megaterium QM B1551 spore germination response to glucose. J Bact. 2010, 192: 4317-4326. 10.1128/JB.00335-10.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  54. Kunnimalaiyaan M, Stevenson DM, Zhou YS, Vary PS: Analysis of the replicon region and identification of an rRNA operon on pBM400 of Bacillus megaterium QM B1551. Mol Microbiol. 2001, 39: 1010-1021. 10.1046/j.1365-2958.2001.02292.x.

    Article  PubMed  CAS  Google Scholar 

  55. Powell JF: Factors affecting the germination of thick suspension of Bacillus subtilis spores in L- alanine solution. J Gen Microbiol. 1950, 4: 330-339.

    Article  PubMed  CAS  Google Scholar 

  56. Paidhungat M, Setlow P: Spore germination and outgrowth. Bacillus subtilis and its closest relatives: From genes to cells. Edited by: Sonenshein AL, Hoch JA, Losick R. 2002, Washington, DC: American Society for Microbiology, 537-548.

    Chapter  Google Scholar 

  57. Setlow B, Peng L, Loshon CA, Li YQ, Christie G, Setlow P: Characterization of the germination of Bacillus megaterium spores lacking enzymes that degrade the spore cortex. J Appl Microbiol. 2009, 107: 318-328. 10.1111/j.1365-2672.2009.04210.x.

    Article  PubMed  CAS  Google Scholar 

  58. Zhang PF, Garner W, Yi XA, Yu J, Li YQ, Setlow P: Factors affecting variability in time between addition of nutrient germinants and rapid Dipicolinic acid release during germination of spores of Bacillus species. J Bact. 2010, 192: 3608-3619. 10.1128/JB.00345-10.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  59. Kong LB, Zhang PF, Setlow P, Li YQ: Characterization of bacterial spore germination using integrated phase contrast microscopy, Raman spectroscopy, and optical tweezers. Anal Chem. 2010, 82: 3840-3847. 10.1021/ac1003322.

    Article  PubMed  CAS  Google Scholar 

  60. Pulvertaft RJV, Haynes JA: Adenosine and spore germination; phase-contrast studies. J Gen Microbiol. 1951, 5: 657-662.

    Article  PubMed  CAS  Google Scholar 

  61. Waites WM, Wyatt LR: The outgrowth of spores of Clostridium bifermentans. J Gen Microbiol. 1974, 84: 235-244.

    Article  PubMed  CAS  Google Scholar 

  62. Patel DC, Dave JM, Sannabhadti SS: Effect of selected heat treatments and added amino acids on germination response of bacterial spores in buffalo milk. Indian J Dairy Sci. 1984, 37: 93-97.

    CAS  Google Scholar 

  63. Griffiths MW, Hurvois Y, Phillips JD, Muir DD: Elimination of spore-forming bacteria from double cream using sub-UTH temperatures. II. Effect of processing conditions on spores. Milchwissenschaft-Milk Sci Int. 1986, 41: 474-478.

    Google Scholar 

  64. Hills GM: Chemical factors in the germination of spore-bearing aerobes: observations on the influence of species, strain and conditions of growth. J Gen Microbiol. 1950, 4: 38-47.

    Article  PubMed  CAS  Google Scholar 

  65. Halmann M, Keynan A: Stages in germination of spores of Bacillus licheniformis. J Bact. 1962, 84: 1187-1193.

    PubMed  CAS  PubMed Central  Google Scholar 

  66. Yi XA, Setlow P: Studies of the commitment step in the germination of spores of Bacillus species. J Bact. 2010, 192: 3424-3433. 10.1128/JB.00326-10.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  67. Paidhungat M, Ragkousi K, Setlow P: Genetic requirements for induction of germination of spores of Bacillus subtilis by Ca2+-Dipicolinate. J Bact. 2001, 183: 4886-4893. 10.1128/JB.183.16.4886-4893.2001.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  68. Riemann H, Ordal ZJ: Germination of bacterial endospores with calcium and dipicolinic acid. Science. 1961, 133: 1703-1704. 10.1126/science.133.3465.1703.

    Article  PubMed  CAS  Google Scholar 

  69. Terry C, Shepherd A, Radford DS, Moir A, Bullough PA: YwdL in Bacillus cereus: Its role in germination and exosporium structure. Plos One. 2011, 6: e23801-10.1371/journal.pone.0023801.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  70. Jaye M, Ordal ZJ: Germination of spores of Bacillus megaterium with divalent metal-dipicolinate chelates. J Bact. 1965, 89: 1617-1618.

    PubMed  CAS  PubMed Central  Google Scholar 

  71. From C, Pukall R, Schumann P, Hormazábal V, Granum PE: Toxin-producing ability among Bacillus spp. outside the Bacillus cereus group. Appl Environ Microbiol. 2005, 71: 1178-1183. 10.1128/AEM.71.3.1178-1183.2005.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  72. Frankland GC, Frankland PF: Studies on some new micro-organisms obtained from air. Phil Trans R Soc London B. 1887, 178: 257-287. 10.1098/rstb.1887.0011.

    Article  Google Scholar 

  73. Ivanova N, Sorokin A, Anderson I, Galleron N, Candelon B, Kapatral V, et al: Genome sequence of Bacillus cereus and comparative analysis with Bacillus anthracis. Nature. 2003, 423: 87-91. 10.1038/nature01582.

    Article  PubMed  CAS  Google Scholar 

  74. Mahillon J, Chungjatupornchai W, Decock J, Dierickx S, Michiels F, Peferoen M, et al: Transformation of Bacillus thuringiensis by electroporation. FEMS Microbiol Lett. 1989, 60: 205-210. 10.1111/j.1574-6968.1989.tb03447.x.

    Article  CAS  Google Scholar 

  75. Arnaud M, Chastanet A, Debarbouille M: New vector for efficient allelic replacement in naturally nontransformable, low-GC-content, gram-positive bacteria. Appl Environ Microbiol. 2004, 70: 6887-6891. 10.1128/AEM.70.11.6887-6891.2004.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  76. Fagerlund A: Bacillus cereus Hbl, Nhe and CytK cytotoxins. PhD thesis. 2007, Norwegian School of Veterinary Science, Departement of Food Safety and Infection Biology

    Google Scholar 

  77. Juajun O, Nguyen TH, Maischberger T, Iqbal S, Haltrich D, Yamabhai M: Cloning, purification, and characterization of beta-galactosidase from Bacillus licheniformis DSM 13. Appl Microbiol Biotechnol. 2011, 89: 645-654. 10.1007/s00253-010-2862-2.

    Article  PubMed  CAS  Google Scholar 

  78. Nicholson WL, Setlow P: Sporulation, germination and outgrowth. Molecular biological methods for Bacillus. Edited by: Harwood CR, Cutting SM. 1990, Chichester: John Wiley and Sons Inc, 391-450.

    Google Scholar 

  79. Vepachedu VR, Setlow P: Analysis of interactions between nutrient germinant receptors and SpoVA proteins of Bacillus subtilis spores. FEMS Microbiol Lett. 2007, 274: 42-47. 10.1111/j.1574-6968.2007.00807.x.

    Article  PubMed  CAS  Google Scholar 

Download references

Acknowledgements and Funding

We would like to thank Kristin O'Sullivan (Norwegian School of Veterinary Science, Oslo, Norway) for technical assistance and Dr Graham Christie (University of Cambridge, England) for sharing the pHT315 vector. The pMAD plasmid was a gift from Michel Débarbouillé (Institut Pasteur, Centre National de la Recherche Scientifique, Paris, France). The work has been financially supported by the Research Council of Norway (grant 178299/I10).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Per Einar Granum.

Additional information

Authors' contributions

ISL assisted in experimental design, carried out the experiments, analysed data and drafted the manuscript. CF assisted in experimental design, carried out the experiments, analysed data and assisted in drafting the manuscript. EHM assisted in drafting the manuscript. EK and KCSR carried out experiments. JTR assisted in drafting the manuscript. PEG assisted in experimental design and drafting of the manuscript. All authors read and approved the final manuscript.

Authors’ original submitted files for images

Rights and permissions

Open Access This article is published under license to BioMed Central Ltd. This is an Open Access article is distributed under the terms of the Creative Commons Attribution License ( https://creativecommons.org/licenses/by/2.0 ), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Løvdal, I.S., From, C., Madslien, E.H. et al. Role of the gerA operon in L-alanine germination of Bacillus licheniformisspores. BMC Microbiol 12, 34 (2012). https://doi.org/10.1186/1471-2180-12-34

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1471-2180-12-34

Keywords