Free access
15 December 1999

The Genetic Switch Regulating Activity of Early Promoters of the Temperate Lactococcal Bacteriophage TP901-1

ABSTRACT

A functional analysis of open reading frame 4 (ORF4) and ORF5 from the temperate lactococcal phage TP901-1 was performed by mutant and deletion analysis combined with transcriptional studies of the early phage promoters pR andpL. ORF4 (180 amino acids) was identified as a phage repressor necessary for repression of both promoters. Furthermore, the presence of ORF4 confers immunity of the host strain to TP901-1. ORF5 (72 amino acids) was found to be able to inhibit repression of the lytic promoter pL by ORF4. Upon transformation with a plasmid containing both ORF4 and ORF5 and their cognate promoters, clonal variation is observed: in each transformant, either pL is open andpR is closed or vice versa. The repression is still dependent on ORF4, and the presence of ORF5 is needed for the clonal variation. Induction of a repressed pLfusion containing orf4 and orf5 was obtained by addition of mitomycin C, and the induction was also shown to be dependent on the presence of the RecA protein, even though ORF4 does not contain a recognizable autocleavage site. Our results suggest that the relative amounts of the two proteins ORF4 and ORF5 determine the decision between lytic or lysogenic life cycle after phage infection and that a protein complex consisting of ORF4 and ORF5 may constitute a new type of genetic switch in bacteriophages.
A temperate bacteriophage has a choice between two very different life cycles after infection of a sensitive host bacterium: it can enter either a lytic cycle or a silent state, transforming the host into a lysogenic bacterium. During the lytic cycle, phage genes are expressed in a temporal manner leading to production of new phages and cell lysis. In a lysogenic bacterium, the genome of the phage is present but only very few genes are expressed. The almost silent phage DNA is transferred to new daughter cells whether it is integrated in the genome, which seems to be the prevailing mechanism, or is present in the form of a plasmid, thus maintaining the lysogenic state of the host. The existence of these two life cycles argues for the presence of tightly regulated promoters and their cognate regulatory proteins in the phage genome. The paradigm for the early expressed promoters is represented by thepL and the pR promoters of bacteriophage λ from Escherichia coli. They are recognized by the host polymerase, and repression is established through the cI repressor protein, while the Cro protein, by competing for binding to the same operator sites as cI, first represses synthesis of cI from the pRM promoter and later in infection also represses transcription from thepR and the pL promoters. The cro gene is the first gene transcribed from the λpR promoter, while the adjacent cI gene is divergently transcribed from the pRMpromoter (for a review, see reference 11).
TP901-1 is a temperate lactococcal phage belonging to the P335 group of phages infecting lactic acid bacteria, including lytic phage species (1). Analysis of the temporal transcriptional pattern during the lytic cycle resulted in localization of consecutive regions of early, middle, and late expressed genes on the phage genome (19). The early region was found to cover 13 kb of the 38.4-kb phage genome, and 6.4 kb of this region was sequenced and found to encode 11 ORFs as well as the phage attachment site attP(5, 19). The region was found to be divergently transcribed in accordance with the orientation of the ORFs, and two consensus promoters (pL and pR) were proposed from the sequence data. The start site for initiation of transcription for both promoters was identified by primer extension (19). The pR promoter is oriented facing attP, and four ORFs were identified as being transcribed from pR, the last of which (orf1) encodes the phage integrase (5). The lytic promoter pL is located upstream ofpR, directing transcription in the opposite direction; the longest mRNA produced from this promoter reaches a size of 10 kb. The sequenced region covers ORF5 to ORF11 and part of ORF12 of the promoter-distal part of this mRNA. In this report, we present experimental evidence for the presence and regulation of these promoters obtained from promoter fusions and primer extension studies. The reported data identify ORF4 as a repressor protein for both promoters and furthermore show that ORF4 is sufficient for achievement of phage immunity. Finally, the data demonstrate that both ORF4 and ORF5 together with the divergent promoter region can function as a genetic switch, resulting in either repression or derepression of the lytic pL promoter.

MATERIALS AND METHODS

DNA technology.

Recombinant plasmid DNA from E. coli was isolated by the alkaline lysis technique. Plasmids were isolated from Lactococcus lactis subsp. cremorisby the alkaline lysis technique after incubation with lysozyme (20 mg/ml) in 20 min at 37°C. Preparative portions were further purified by passage through Qiagen columns as recommended by the supplier (Qiagen, Hilden, Germany). Pharmacia Biotech supplied restriction endonuclease enzymes, DNA polymerase Klenow fragment, T4 DNA ligase, and buffer systems. All enzymes were used as recommended by the supplier. PCRs were performed in a DNA thermal cycler (Perkin-Elmer Cetus) using Amplitaq polymerase and buffer supplied by Perkin-Elmer. DNA sequencing was performed by the method of Sanger and coworkers (24) as instructed by the protocol for the Sequenase version 2.0 DNA sequencing kit (U.S. Biochemical Corp., Cleveland, Ohio).

Construction of plasmids.

The plasmids used in this study are listed in Table 1. Plasmid pPM92 was constructed by cloning a 2,787-bp XhoI fragment of pG5f4 (4) into the XhoI site of the E. colivector pIC19R (20). This plasmid contains orf4 toorf9 of TP901-1. The following plasmids were constructed by cloning PCR products obtained by using pPM92 as a template, a vector primer (1211; see below), and primers located either in the region between orf4 and orf5 (24BamHI, P5A, and erm2ABamHI), within orf5 (pextPstI), or downstream oforf5 (orf4orf5PstI and porf7PstI). All clonings were performed in L. lactis subsp. cremoris MG1363. Plasmids pAJ93 and pAJ94 were constructed by cloning aBamHI-digested PCR product (primers 24BamHI and erm2ABamHI) into pAK80 (13) also digested with BamHI. The resulting plasmids contain pL (pAJ93) andpR (pAJ94) fused to the lacLM genes. PCR products obtained by using primer 1211 and pextPstI were cloned into pAK80 after digestion with PstI, giving rise to plasmids pPM126 and pPM137 carrying orf4 as well aspL and pR fused to thelacLM genes, respectively. Plasmids pPM127 and pPM132 were constructed by cloning a HindIII-digested PCR product (primers 1211 and porf7PstI) into pAK80. The plasmids thus containorf4 and orf5 as well aspL (pPM127) or pR(pPM132) fused to the lacLM genes. A frameshift mutation was introduced in the beginning of the orf4 gene by filling in the SpeI site of pPM92 with Klenow polymerase. After ligation and redigestion with SpeI, two PCRs were performed on the ligation mixture. In the first PCR, primers 1211 and pextPstI were used, and a PCR product containing the mutated orf4gene was cloned into pAK80 after digestion with PstI, giving rise to plasmids pPM138 and pPM136, carrying only the mutatedorf4 gene as well as pL andpR fused to the lacLM genes, respectively. In the second PCR, primers 1211 and orf7PstI were used. After digestion with HindIII, a PCR product containingorf5 in addition to the mutated orf4 was cloned in pAK80, and plasmids pPM131 and pPM139 were constructed. Plasmids pPM131 and pPM139 thus contain pL andpR, respectively, fused to the lacLMgenes. In all cases, several independent clones giving rise to the same β-galactosidase specific activity were isolated. Furthermore, the inserts of plasmid pAJ93, pAJ94, pPM127, pPM126, and pPM132 were sequenced, as well as the mutation within the orf4 gene in plasmids pPM131, pPM136, pPM138, and pPM139. The PCR product generated by use of primer 1211 and erm2ABamHI was digested with BamHI and cloned into pCI372 (10) also digested withBamHI, resulting in plasmid pAJ80, carryingpL, pR, andorf4. The PCR product produced by using primer 1211 and P5A was cloned into pNZ8010 (7) after digestion withBamHI, giving rise to pAJ115 containing orf4 andpL fused to the gusA gene. PCR products produced by using primer 1211 and orf4orf5PstI or Orf5A and Orf5B were cloned into pNZ8010 (7) after digestion withPstI and BamHI-PstI, respectively. Thereby plasmids pAJ98 and pAJ142 were constructed so as to containorf4 and orf5 transcribed frompR and pL on a high-copy-number plasmid and orf5 under the control of the nisin promoter, respectively. The inserts of pAJ80, pAJ98, pAJ115, and pAJ142 were sequenced.
Table 1.
Table 1. Plasmids used in this work
Plasmid Descriptiona Promoter fusion Relevant genotype Reference
pIC19R       20
pCI372       10
pCI3340       10
pG5f4 pGEM5::EcoRV4 of TP901-1   orf2–orf10b 4
pAK80     lacLM 13
pNZ8010   pnisA-gusA   7
pPM92 pIC19R::2787 bp XhoI pG5f4   orf4–orf9b This work
pPM126 pAK80::PstI PCR (2622–3458) pL-lacLM orf4b This work
pPM127 pAK80::HindIII PCR (2622–3900) pL-lacLM orf4 orf5b This work
pPM131 pAK80::HindIII PCR (2622–3900) pL-lacLM orf5b This work
pPM132 pAK80::HindIII PCR (2622–3900) pR-lacLM orf4 orf5b This work
pPM136 pAK80::PstI PCR (2622–3458) pR-lacLM   This work
pPM137 pAK80::PstI PCR (2622–3458) pR-lacLM orf4b This work
pPM138 pAK80::PstI PCR (2622–3458) pL-lacLM   This work
pPM139 pAK80::HindIII PCR (2622–3900) pR-lacLM orf5b This work
pAB223 pCI3340::XhoI-HindIII (2621–3900)   orf4 orf5b 2
pAJ80 pCI372::BamHI PCR (2622–3332)   orf4b This work
pAJ93 pAK80::BamHI PCR (3221–3332) pL-lacLMc   This work
pAJ94 pAK80::BamHI PCR (3221–3332) pR-lacLMc   This work
pAJ98 pNZ8010::PstI PCR (2622–3900)   orf4 orf5b This work
pAJ115 pNZ8010::BamHI PstI PCR (2622–3370) pnisA-pR-orf4 gusA orf4b This work
pAJ142 pNZ8010::BamHI PstI PCR (3338–3574) pnisA-orf5 gusA orf5b This work
a
Coordinates in parentheses refer to the sequence Y14232 .
b
Open reading frame(s) of TP901-1 present.
c
The −10 region of thepR promoter is disrupted.

Primers used in this study.

For construction of plasmids, the following primers were used: 1211 (5′-GTAAAACGACGGCCAGT-3′), 24BamHI (5′-GCGCGGATCCTATAGCGCATCTTGAAC-3′), P5A (5′-GCGCGGATCCTAATCATAACTCATTTATGT-3′), erm2ABamHI (5′-GCGCGGATCCCACGTTTCATGAACTTT-3′), orf4orf5PstI (5′-GCGCCTGCAGCCCAAGCTTCATCAGTTC-3′), Orf5A (5′-GCGCGGATCCAAGAAAGGAGAAACATAAAT-3′), Orf5B (5′-GCGCCTGCAGTTAATGAACTTTTGCATTAA-3′), pextPstI (5′-AAAACTGCAGAGACACAGTTCTCTCTG-3′), and porf7PstI (5′-AAAAAACTGCAGATCCCCCATGTGCTTTCC-3′). For primer extensions, primer PE (5′-AGACACAGTTCTCTCTGAAAGCCCC-3′) was used for mapping of the pL promoter, while primer 23 (5′-GGTCAGGAGATTGAACG-3′) was used forpR.

Transformation and selection in E. coli and L. lactis subsp. cremoris.

E. coli XL1-Blue was made competent with CaCl2 and was transformed as described by Sambrook et al. (23). Selection was performed with 100 μg of ampicillin per ml. L. lactis subsp.cremoris MG1363 was transformed by electroporation according to the method described by Holo and Nes (12), using 0.03 to 0.5 μg of DNA per electroporation. When L. lactis subsp.cremoris 3107 was made competent, the growth medium (12) contained only 0.2 M sucrose, and 25 μg DNA was used per transformation. Transformants were selected on plates containing 5 μg of erythromycin per ml. Screening on 5-bromo-4-chloro-3-indolyl-β-d-galactopyranoside (X-Gal) plates was performed at a concentration of 90 μg/ml.

Bacteria and phages.

Bacterial strains used in this study are listed in Table 2. The temperate bacteriophage TP901-1 was induced from L. lactis subsp.cremoris 901-1 (1) by the use of UV light as previously described (4).
Table 2.
Table 2. Bacterial strains used in this study
Strains Relevant description or genotype Source or reference
E. coliXL1-Blue   Stratagene
L. lactis subsp. cremoris
 3107 Indicator strain for TP901-1 1
 901-1 Lysogenic for TP901-1 1
 MG1363   9
 NZ3900 MG1363pepN::nisRK 6
 VEL1122 MG1363 ΔrecA::tet 8
 PM75 MG1363/pPM126 This work
 PM80 MG1363/pPM131 This work
 PM81 MG1363/pPM132a This work
 PM85 MG1363/pPM136 This work
 PM86 MG1363/pPM137 This work
 PM87 MG1363/pPM138 This work
 PM88 MG1363/pAK80 This work
 PM92 MG1363/pPM139 This work
 PM93 MG1363/pPM127b This work
 PM94 MG1363/pPM127c This work
 PM95 MG1363/pPM132d This work
 PM97 3107/pPM138 This work
 PM99 3107/pPM131 This work
 LKH48 3107/pPM127c This work
 LKH49 3107/pPM127d This work
 LKH51 3107/pPM126 This work
 LKH68 VEL1122/pPM127c This work
 LKH148 NZ3900/pPM127b This work
 LKH149 NZ3900/pPM127c This work
 LKH169 NZ3900/pNZ8010 pPM127c This work
 LKH206 NZ3900/pAJ142 pPM127c This work
a
White strain on plates containing X-Gal (90 μg/ml) after transformation with pPM132.
b
Very blue strain on plates containing X-Gal (90 μg/ml) after transformation with pPM127.
c
White strain on plates containing X-Gal (90 μg/ml) after transformation with pPM127.
d
Light blue strain on plates containing X-Gal (90 μg/ml) after transformation with pPM132.

Cell growth and enzyme assay.

E. coli strains were grown with agitation at 37°C in Luria-Bertani broth (23) (Difco Laboratories, Detroit, Michigan). Bacto Agar (Difco Laboratories) was used at 1.5% (wt/vol) in solid media. L. lactis strains were propagated at 30°C without shaking in GM17 (M17 broth [Oxoid Limited, Basingstoke, Hampshire, United Kingdom] containing 0.5% [wt/vol] glucose) (26). For determination of β-galactosidase activity, exponential growing cells were diluted in fresh medium to an optical density at 600 nm (OD600) of 0.05. At an OD600 of 0.8, 70 ml of culture was harvested and resuspended in 7 ml of Z-buffer, and cell extracts were obtained with a French press. The activity of β-galactosidase was determined according to Miller (21), and specific activity was calculated in Miller units.

Induction of transcription from the nisin promoter.

Exponentially growing cells of strains LKH169 and LKH206 were diluted in fresh medium (GM17 containing 5 μg of erythromycin per ml and 5 μg of chloroamphenicol per ml) to an OD600 of 0.05. At an OD600 of 0.2, nisin was added to the culture to a final concentration of 1 ng/ml. Samples were withdrawn before and after addition of nisin, diluted, and plated on plates containing X-Gal but no nisin. After overnight incubation at 30°C, blue and white colonies were counted.

Induction by mitomycin C.

Strains were grown in GM17 containing 5 μg of erythromycin per ml at 28°C, due to the temperature sensitivity of the recA strain. Exponentially growing overnight cultures were diluted to an OD600 of 0.05 in fresh medium, and growth was monitored. At an OD600 of 0.4, each culture was diluted into two separate cultures to an OD600 of 0.1. Subsequently, mitomycin C was added to a final concentration of 2.5 μg/ml to one of the cultures, while the other served as a control. Samples were removed at different time points, harvested, extracted, and assayed as described above for β-galactosidase activity.

Plaque assay.

TP901-1 phages were diluted in A buffer (0.8% [wt/vol] NaCl, 10 mM Tris-HCl [pH 7.5], 5 mM MgCl2, 1 mM CaCl2) and plated on lawns of L. lactissubsp. cremoris 3107 transformed with either pPM126, pPM138, pPM127, or pPM131. As a control, L. lactis subsp.cremoris 3107 was used as an indicator strain. After incubation at 30°C overnight, plaques were counted. For each strain, the efficiency of plaquing (EOP) was calculated as the number of plaques formed on a lawn of the strain of interest divided by the number of plaques formed on L. lactis subsp.cremoris 3107.

Extraction of RNA.

RNA was isolated from strains PM93, PM94, and PM80 growing exponentially in defined medium SA (14) (1.0% glucose [wt/vol], 5 μg of erythromycin per ml), using the method described by Johnsen et al. (15) as modified by Madsen and Hammer (19).

Primer extensions.

Primers PE and 23 were phosphorylated by using T4 polynucleotide kinase (Gibco BRL) and [32P]ATP (Amersham) as described by Sambrook et al. (23); 25 μg of total RNA isolated from strain PM93, PM94, or PM80 was mixed with 1 pmol of phosphorylated primer in a volume of 10 μl and incubated at 70°C for 10 min. The mixture was allowed to cool to 37°C over a 1-h period. Then 1 μl of Superscript reverse transcriptase (Gibco), 4 μl of 5× first-strand buffer (Gibco), 2 μl of 0.1 M dithiothreitol (Gibco), and dATP, dGTP, dCTP, and dTTP to a final concentration of 500 μM were added, and the mixture was incubated at 37°C for 1 h.

RESULTS

ORF4 is a negative regulator of pL andpR.

Previously, we identified two divergently orientated promoters (pL andpR) located between orf4 andorf5 of the temperate bacteriophage TP901-1 (19). To investigate the expression and regulation of the promoters, we cloned a phage DNA fragment of 837 bp, containing both promoters andorf4, in both orientations in the promoter probe vector pAK80. The lacLM genes contained in this plasmid give rise to β-galactosidase activity when a functional promoter is inserted in the cloning region, and stop codons in all three reading frames ensure that a transcriptional fusion is constructed. WhenpL (pPM126) and pR(pPM137) are located upstream of the lacLM genes, low β-galactosidase activities (1.0 and 0.5 Miller units, respectively) can be measured when the plasmids are present in L. lacticsubsp. cremoris MG1363 (Fig.1). These enzyme levels are, however, significantly different from the basal level of 0.02 Miller units measured in MG1363 containing the pAK80 vector, which is in agreement with the existence of the divergently orientated promoterspL and pR on this phage DNA fragment. The effect of the presence of orf4 on the activity of the promoters was investigated by introduction of an early frameshift mutation at codon position 6. The β-galactosidase activities of the resulting plasmids, carrying thepL fusion (pPM138) and thepR fusion (pPM136) were measured as 62 and 15 Miller units, respectively, in MG1363. These results demonstrate the presence of divergently orientated promoter activities on the cloned phage DNA fragment and clearly show that the presence of a functional ORF4 protein represses the expression of both phage promoters. The levels of repression are 62- and 31-fold for pLand pR, respectively (Fig. 1).
Fig. 1.
Fig. 1. Effects of orf4 and orf5 on expression from pL (A) andpR (B). Fragments cloned upstream of thelacLM genes of the pAK80 vector are indicated by lines. ThelacLM genes are located downstream ofpL (A) or pR (B). Small vertical lines indicate a frameshift mutation in orf4. Plasmid numbers and specific activities of β-galactosidase (β-gal) of the corresponding strains are indicated. For details of cell growth and the β-galactosidase assay, see Materials and Methods. For PM80 and PM95, only two independent experiments were performed; the remaining measurements are averages of three to five experiments.
Since the spacer between the initiation sites of thepL and pR transcripts is only 101 bp, it is possible that binding of the RNA polymerase to one of them affects transcription of the other. To exclude this possibility, a 111-bp PCR fragment carrying thepL promoter but only the −35 region of thepR promoter was cloned in both orientations upstream of the lacLM genes in the promoter probe vector pAK80. The specific activity of β-galactosidase of L. lactis subsp. cremoris MG1363 containing either pAJ94 (pR fusion) or pAJ93 (pLfusion) was measured. The presence of pAJ94 (pRfusion) gave rise to only a background level of β-galactosidase activity (0.14 Miller units), as expected due to the absence of the −10 region of the pR promoter. MG1363 containing pAJ93 (pL fusion), in which thepR promoter is not active, showed the same specific β-galactosidase activity (66.5 Miller units) as when the larger pL fusion (pPM138) carrying an activepR promoter was present. Thus, the activity of the pL promoter is the same regardless of whether the pR promoter is present, showing that the activity of the pR promoter does not affect the efficiency of the pL promoter.

Antirepressing activity of ORF5.

In bacteriophage λ, the Cro and the cI repressors compete for the λ operator sites and thereby regulate expression of the pRM,pR, and pL promoters. To investigate the influence of both ORF4 and ORF5 on thepL and pR promoters of TP901-1, we cloned a 1,279-bp phage DNA fragment carrying bothorf4 and orf5 in both orientations in pAK80, resulting in plasmids pPM127 (pL fusion) and pPM132 (pR fusion). The transformants arising from both plasmids exhibited clonal variation when screened on GM17 plates containing X-Gal. Of 200 transformants arising from pPM127 (pL fusion), strong blue colonies were observed in 191 cases and the remaining 9 colonies were white, giving a frequency of 4.5% of white colonies. This frequency is comparable to the frequency of lysogenization of approximately 1 to 3% by TP901-1 (3). When plasmid pPM132 (pR fusion) was used, the same type of clonal variation was observed but with opposite frequencies: 95% white colonies and only 5% light blue transformants were observed. The phenotype was stable for both plasmids when restreaked to single colonies more than three consecutive times and also after freezing and restreaking. When plasmid DNA was isolated from either blue or white transformants and used for retransformation of MG1363, segregation into white and blue transformants was observed again with the same frequencies as seen before. The phenotype of a transformant was thus independent of the phenotype of the strain from which the plasmid was prepared.
In light of the observed variation of phenotype when bothorf4 and orf5 were present, we repeated the transformation of MG1363 with plasmids pPM126 (pL fusion) and pPM137 (pR fusion), containing only orf4. Screening of the transformants on 90 μg of X-Gal per ml made it possible to visualize the difference in enzyme levels of 0.5, 1.0, and 62 Miller units as light blue, blue, and very blue colonies, respectively, all being distinctively different from the white colonies. Again 200 transformants containing either pPM126 (pL fusion) or pPM137 (pRfusion) were obtained and all were either light blue (pPM137) or blue (pPM126), corresponding to the measured enzyme levels of 0.5 and 1.0 Miller units (Fig. 1). Clonal variation was therefore not observed in the presence of only ORF4, suggesting that ORF5 is necessary for this phenomenon.
To assess the effect of the repression mediated by ORF4 in the presence of ORF5, orf4 mutants were constructed in both plasmids by filling in the SpeI site in the N-terminal region oforf4. β-Galactosidase activities of the transformants containing the clonal variants of pPM127 (pLfusion) and of the clonal variants of pPM132 (pRfusion) as well as the plasmids carrying the orf4 mutation were measured (Fig. 1). First, let us consider expression from the lytic promoter pL. In the presence oforf4 and orf5, the expression frompL observed in PM94 is about 1,000-fold repressed compared to the activity found in the derepressed state (PM93) (Fig. 1). Furthermore, the pL promoter activity in PM93 (containing orf4 and orf5) is as high as in the orf4 mutant (PM80); thus, PM93 is phenotypically an orf4 mutant (Fig. 1). Expression frompR is also subject to clonal variation. However, in the presence of both orf4 and orf5, the difference between the repressed state of the pRpromoter in strain PM81 and the derepressed state in strain PM95 is only fourfold. When a mutation was introduced into orf4, the expression from pR becomes even more derepressed, 439-fold compared to the expression found in strain PM81 (Fig. 1). Thus, also in the presence of ORF5, ORF4 is needed for repression of both pR andpL.

ORF4 and ORF5 are trans acting.

The effect of ORF4 and ORF5 donated from a plasmid in trans to the promoter fusions contained on the compatible pAK80 plasmid was investigated. As observed in Table 3, the presence oforf4 on the donor plasmid results in repression of both apL fusion (pPM138) and apR fusion (pPM136) in the recipient strain, whereas the vector plasmid itself (pCI372) does not change the constitutive expression monitored as blue colonies. Thus, ORF4 is able to repress both pL and pRwhen given in trans. Subsequently, the effect of donation of a plasmid containing both orf4 and orf5 (pAB223) to either the pR or thepL fusion strain was investigated. The result clearly showed clonal variation, with the majority of thepL fusions turning blue and the majority of thepR fusions turning white. Thus, the same effect of ORF4 and ORF5 was observed whether orf4 andorf5 were donated in trans or were located on the same plasmid as the promoters (pPM127 and pPM132). Regulation ofpL and pR thus seems to be dependent on ORF4 and ORF5, both being able to function intrans.
Table 3.
Table 3. Effect of donation of orf4 and/ororf5 in trans to promoter fusion plasmids measured as the color of transformants on X-Gal plates
Recipient strain Color of transformantsa
Strain Plasmid Genotype pCI372b pAJ80 (low copy no.; orf4) pAJ115 (high copy no.;orf4) pAB223 (low copy no.; orf4 orf5)
PM85 pPM136 pR-lacLM B W ND 94% W, 6% B
PM87 pPM138 pL-lacLM B W ND 1.2% W, 98.8% B
LKH148 pPM127 orf4 orf5 pL-lacLMc B B 92% W, 8% B ND
LKH149 pPM127 orf4 orf5 pL-lacLMd W ND ND We
a
B, blue colonies on plates containing X-Gal (90 μg/ml) after transformation with the indicated donor plasmid; W, white colonies on plates containing X-Gal (90 μg/ml) after transformation with the indicated donor plasmid; ND, not determined.
b
Donor plasmid for transformation.
c
Very blue strain on plates containing X-Gal (90 μg/ml) after transformation with pPM127.
d
White strain on plates containing X-Gal (90 μg/ml) after transformation with pPM127.
e
Transformants were also white when the high-copy-number plasmid pAJ98 (orf4 orf5) was used as the donor plasmid.

The state of pL can be changed.

The clonal variation mediated by the presence of both ORF4 and ORF5 was found to be stably inherited. Particularly, expression frompL is affected by the choice of configuration, since an approximately 1,000-fold difference in the specific activities of β-galactosidase of strains PM93 (pL open) and PM94 (pL closed) was found (Fig. 1). The ratio between the amount of ORF4 and ORF5 could determine the choice of configuration of pL, and we therefore examined whether the state (repressed or open) of the pLpromoter could be changed by introduction of additional amounts of either ORF4 or ORF5. For this purpose, strain LKH148 carrying pPM127 (orf4 orf5), in which pL is open, was transformed with plasmids expressing orf4 from thepR promoter, and the color of the transformants was observed on plates containing X-Gal. In the presence oforf4 on a low-copy-number plasmid (pAJ80), only blue transformants were seen, whereas 92% white and 8% blue colonies were found when orf4 was present on a high-copy-number plasmid (pAJ115) (Table 3). It is thus possible to change an open state of thepL promoter to a repressed state by introduction of a sufficient amount of additional ORF4.
In the situation where pL is closed (strain LKH149 carrying pPM127 [orf4 orf5]) we investigated the effect of increasing the expression of orf5 first by introducing a plasmid containing both orf5 andorf4 transcribed from pL andpR, respectively (pAB223). However, this did not change the repressed state of the pL promoter located on pPM127 in LKH149, even when a high-copy-number plasmid was used (Table 3). This was expected since a surplus of ORF4 is able to shut down expression of the pL promoter andorf5 expression is dependent on pL in pAB223. To be able to produce a surplus of ORF5, a plasmid containingorf5 under the control of the inducible promoterpnisA was therefore constructed (pAJ142). Addition of the inducer nisin will lead to a rapid increase inorf5 expression. For this experiment we used strain LKH206, which contains pAJ142 in addition to the pLfusion on pPM127 (orf4 orf5). The advantage of this experimental setup is that the influence of a short induction period with nisin in the liquid culture on the final choice of the phenotype of a colony originating from a cell diluted from this culture and plated without nisin can be tested. Thus, after dilution, the color of LKH206 was examined on plates containing X-Gal but no nisin before and after nisin was added to the culture (Fig.2). Before addition of nisin, 4 to 6% blue colonies were seen, probably due to a low frequency of initiation of transcription of the nisin promoter also in the absence of nisin. After 5 min of growth in the presence of nisin, 51% of the colonies were blue if samples were removed and plated. This frequency of blue colonies was increased further to 84% after 10 min, and after 40 min all colonies were blue. The phenotype of the blue colonies obtained in this experiment was stable after restreaking to single colonies. This shows that a repressed state of pL can be changed into an open state if the amount of ORF5 is increased only momentarily (10 min) and that once established, this configuration can be maintained without further nisin-induced ORF5 production.
Fig. 2.
Fig. 2. Change of the pL promoter from the closed to the open configuration by a burst of induction of synthesis of ORF5 from the nisin promoter. Strains LKH169 and LKH206 both contain pPM127 carrying the pL promoter in the closed configuration. At time zero, nisin (1 ng/ml) was added to exponentially growing cultures. At the indicated times, samples were withdrawn, diluted, and plated on GM17 plates containing X-Gal. After incubation at 30°C, the number and color of the colonies were determined and the frequency of blue colonies was calculated. Circles, strain LKH206 containing pPM127 and pAJ142 (orf5 under the control of the nisin promoter); triangles, strain LKH169 containing pPM127 and pNZ8010 (the vector of pAJ142).
Next we wanted to show that overproduction of ORF5 is also able to inhibit the ORF4-mediated repression of the pLpromoter in the short fusion pPM126, which contains only 145 bp of thepL transcript, in contrast to the 587 bp in pPM127. Strains containing pPM126 (orf4 pLfusion) and pAJ142 (carrying orf5 under the control of thepnisA promoter) or pNZ8010 (vector of pAJ142) were grown in the presence of nisin, and the specific activity of β-galactosidase was measured. An 11-fold increase in expression frompL in the presence of expressed orf5was found. Thus, ORF5 is also able to inhibit the ORF4-mediated repression of the pL promoter in the short fusion contained in pPM126.

Study of the regulation of promoters by primer extension.

In a previous study, locations of the divergent promoterspL and pR in the region between the divergent reading frames in the early expressed region of TP901-1 were determined by primer extension (19). The results from the transcriptional fusions presented in the present report have also demonstrated the presence as well as the regulation of divergent promoter activity. By primer extension analysis, we were furthermore able to simultaneously monitor expression from thepR and pL promoters in the same strain. Therefore, we decided to map the transcriptional start sites from both promoters by primer extension using mRNA prepared from three strains: PM93 and PM94, both containing pPM127 but carrying thepL promoter in the derepressed state and the repressed state, respectively, as well as PM80, containing the corresponding plasmid with a mutation in orf4 (pPM131). In PM93 (orf4 and orf5 present), expression from thepL promoter is high, as shown by the β-galactosidase measurements. The mRNA prepared from this strain shows an mRNA start site from the pL promoter but none from the pR promoter (Fig.3, lanes 1). On the other hand, when thepL promoter is in the repressed state as in the clonal variant PM94 containing the same plasmid (pPM127) as PM93, transcription from pR but not frompL is observed (lanes 2). For comparison, we investigated the primer extension results from the strain containing the orf4 mutant plasmid, pPM131. Since the promoter fusions showed that ORF4 represses transcription from both promoters, we would expect to find transcriptional start sites from both promoters in this strain. As shown in lanes 3, this was also observed. The transcripts from pR are 4 bp longer due to filling in of theSpeI site, which is located between the primer used and the promoter. Therefore, the same transcriptional start site is used for expression from pR in strains PM94 and PM80, just as in the case of pL expression in both PM93 and PM80. The results for strains containing pPM127 (orf4 and orf5 present) demonstrate that the two different phenotypes which can be obtained with this plasmid result from a strain where pL is open andpR is simultaneously closed, and vice versa: a strain where pL is closed andpR is open. However, as judged from the enzyme measurements in Fig. 1, pR is only partially derepressed in this situation.
Fig. 3.
Fig. 3. Primer extension analysis of regulation ofpL and pR. Total RNA used in the analysis was isolated from the following strains, all derivatives of MG1363: lanes 1, PM93 containing pPM127 (pL open); lanes 2, PM94 containing pPM127 (pL closed); lanes 3, PM80 containing pPM131 (pL and pR open). Lanes G, A, T, and C contain the sequence reactions performed with the primers used for the analysis. (A) Mapping of thepL promoter; (B) mapping of thepR promoter. To the right of each lane, the sense strand sequence is given; the −10 and +1 positions of the promoters are also indicated.

Effect of ORF4 and ORF5 on phage immunity.

The effect of ORF4 and ORF5 on the plating capacity of phage TP901-1 was investigated after transformation of the relevant plasmids into the phage indicator strain L. lactis subsp. cremoris 3107. We tested the presence of ORF4 alone by using pPM126 (pLfusion) and, as a control, the corresponding orf4 mutant plasmid pPM138. A strain containing orf4 (LKH51) showed full resistance to TP901-1 and an EOP of less than 10−6, compared to the EOP of 1 for both the orf4 mutant strain (PM97) and the indicator strain 3107 (Fig.4). The effect of the presence of bothorf4 and orf5 was tested with pPM127 (pL fusion). Use of this plasmid also gave rise to clonal variation in the indicator strain 3107. LKH48 and LKH49 are representatives of clones containing repressed and derepressedpL promoters, respectively. No plaques were detected with LKH48, while a few pinpoint plaques were obtained with LKH49. For comparison, we also used PM99 containing the orf4mutant plasmid pPM131, which should produce ORF5 frompL since orf5 is located downstream of pL, which in this strain has a high activity (plasmid pPM131 in strain PM80 [Fig. 1]). As reported in Fig. 4, the presence of this plasmid is inhibitory to the plaque-forming capacity of TP901-1, giving an EOP of 10−4. This is only about fivefold higher than the EOP on LKH49, indicating that the inhibitory effect in LKH49 is mainly due to the production of ORF5.
Fig. 4.
Fig. 4. Effect of orf4 and orf5 on phage TP901-1 immunity. The lines indicate fragments cloned upstream of thelacLM genes of the pAK80 vector. In all cases, thelacLM genes are located downstream of thepL promoter, and the state ofpL is indicated as either repressed or open, as judged from the color of the strains on plates containing X-Gal. A small horizontal line indicates a frameshift mutation inorf4. For each strain, the transformed pAK80 plasmid derivative and the EOP are noted. For details, see Materials and Methods.

Induction by mitomycin C is dependent on RecA and ORF5.

For bacteriophage λ, it is known that during the maintenance of lysogeny the cI repressor is bound to the pR promoter, preventing transcription of the genes involved in lytic development. In the presence of UV light or mitomycin C, which both induce the SOS response, the cI repressor is inactivated by autocleavage induced by the RecA protein (17). To investigate whether transcription can be induced from the clone containing the lyticpL promoter of TP901-1 by treatment with mitomycin C, β-galactosidase activity was monitored after addition of mitomycin C to strain PM94, carrying both orf4 andorf5. Strain PM94 is the clonal variant carrying thepL promoter in the repressed state. As seen in Fig. 5, the β-galactosidase activity was induced 28-fold by mitomycin C. Furthermore, mitomycin C induction of the pL promoter fusion was found to be dependent on a functional recA allele, since no induction was seen in a recA mutant (LKH68 in Fig. 5). However when PM75, the recA+ strain containing thepL fusion plasmid pPM126 with onlyorf4, was exposed to mitomycin C, no induction of β-galactosidase activity was observed (data not shown).
Fig. 5.
Fig. 5. Induction of pL by mitomycin C in strains containing the pL promoter fused to thelacLM genes of pAK80. At time 5 min, mitomycin C (2.5 μg/ml) was added. Triangles, strain PM94 containing pPM127 (orf4 orf5 with repressed pL); circles, strain LKH68 recA derivative containing pPM127 (orf4 orf5 with repressed pL).

DISCUSSION

By primer extension studies and use of transcriptional fusions, the early transcribed region of phage TP901-1 has been shown to contain two divergently orientated promoters. The untranscribed DNA between the mRNA start sites located at positions 3212 and 3314 has a length of 101 bp. Both promoters have close to consensus sequences with 17 bp between the −10 and −35 regions; furthermore, pL has the TG motif 1 bp upstream of the −10 region. By deletion of the −10 region of the pR, promoter it was shown that the efficiency of pL is independent of the activity of pR. Furthermore, it was shown that both promoters are active in the absence of expressed phage TP901-1-encoded genes.
In this study, we investigated the functions of ORF4 and ORF5 that are located topologically equivalent to the cI and Cro proteins of phage λ. The transcriptional fusions demonstrated that ORF4 is necessary for negative regulation of both pR andpL, and the primer extensions (Fig. 3) strongly indicate that the negative effect of ORF4 is at the level of transcription initiation. As shown by the phenotype of the strains containing plasmids pPM126 and pPM137, ORF4 alone is furthermore sufficient for repression of both promoters (Fig. 1). This repression is also seen when ORF4 is donated in trans to the promoter fusions (Table 3). The role of ORF4 as the putative phage repressor was further substantiated by proving that the orf4-containing plasmids conferred immunity to TP901-1 when transferred to the phage indicator strain 3107, while introduction of a frameshift mutation inorf4 resulted in abolishment of the immunity of strain 3107 (Fig. 4). The experiments furthermore showed that large amounts of ORF5 are inhibitory to formation of plaques.
We investigated the effect of the presence of both ORF4 and ORF5 onpL and pR promoter activities by analysis of plasmids pPM127 and pPM132 and found that both plasmids gave rise to clonal variation among the transformants. The presence of ORF5 was found to be necessary for the observed clonal variation, but ORF4 was still responsible for repression of the promoters. It was demonstrated that the pLpromoter was either open or closed in the obtained transformants and that the clonal variation was stably inherited. However, an open configuration of pL could be changed to the closed configuration by introduction of additional amounts of ORF4 donated from a high-copy-number plasmid, while no change was seen when ORF4 was derived from a plasmid with moderate copy number (Table 3). Furthermore, a burst of ORF5 production from the nisin promoter was sufficient to alter the configuration of the pLpromoter from closed to open, and the phenotype was stably inherited when the cells were plated without nisin. These results show that the amount of the regulatory proteins ORF4 and ORF5 as well as the ratio determines the choice of the open or repressed state of thepL promoter. We have also shown that ORF5 is able to inhibit the ORF4-mediated repression ofpL in the short fusion containing only 145 bp of the pL transcript.
By primer extension analysis, we found that whenpL is open pR is totally repressed, while pR transcripts can be found when pL is closed. However, the enzyme levels of the transcriptional pR fusions demonstrated thatpR is only partially derepressed under these conditions, 0.5 Miller units in PM95 compared to 0.1 in PM81 (Fig. 1).
When the pL promoter is open (43 Miller units in PM93), this results in a large production of pLmRNA encoding ORF5 and therefore presumably rather high levels of ORF5 in the cell (Fig. 1). In these conditions pR is repressed (0.1 Miller units), but some ORF4 must be produced since ORF4 is necessary for the repression of pR, as verified by the orf4 mutation in strain PM92 (Fig. 1). The presence of the high level of ORF5 must somehow preferentially interfere with the binding of ORF4 to the pLpromoter but not with the binding of ORF4 to thepR promoter. Thus, initiation of transcription from pR is repressed by ORF4, whereas the presence of large amounts of ORF5 prevents repression ofpL by ORF4. This resembles the commitment to lytic growth where the pL transcripts are needed (19).
In the situation where pL is repressed, stimulating the lysogenic state, activity from thepL promoter in PM94 is only 0.04 Miller units, and therefore only small amounts of ORF5 could be produced in these cells (Fig. 1). The pR fusion in PM95 shows activity of 0.5 Miller units; thus, considerably more ORF4 is produced than when pL is open. When both ORF4 and ORF5 are present in the cell, pL is repressed 25-fold more than in the situation where only ORF4 is present, as seen by comparison of PM94 (0.04 Miller units) with PM75 (1.0 Miller units) in Fig. 1. This does not seem to be due to different levels of ORF4 produced in these cells since the output from pRis the same under both conditions (compare PM86 [0.47 Miller units] with PM95 [0.48 Miller units] in Fig. 1). Therefore, we suggest that some ORF5 is produced when pL is repressed and that ORF5 is able to enhance the repression ofpL in these cells. In summary, these results suggest that during commitment to lysogenic growth, sufficient ORF4 is produced from the pR promoter to repress initiation of transcription from both pR andpL and that the repression ofpL is further enhanced by the presence of small amounts of ORF5.
We assume that the ORF4 protein binds somewhere in the +10 to −45 region of the promoters in order to repress initiation of transcription. ORF5 may also be a DNA binding protein since it contains a helix-turn-helix motif (from Gln19 to Asn38) (19). In bacteriophage λ, the regulatory proteins cI and Cro regulates expression of the pR andpRM promoters by competition for the same operator sites in the region of the two promoters. But a similar regulation is not likely for TP901-1 since ORF5, in the presence of ORF4, both enhances (large amounts) and inhibits (small amounts) the level of initiation of transcription of the pLpromoter. In addition to this, ORF5 alone is not able to repress transcription from the TP901-1 early promoters. Instead, ORF5 may exert its action by binding to the phage DNA in a complex with ORF4 or just by binding to ORF4 and thereby somehow prevent repression ofpL by ORF4. The latter case would thus resemble the situation for the antirepressor Ant from E. coli phage P22 (for a review, see reference 22). In the first case, the high amount of ORF5 present during commitment to lytic growth may result in formation of an ORF4-ORF5 protein complex which is able to repress initiation of transcription of pR and not that of pL. On the other hand, during commitment to lysogenic growth ORF4 is free to bind to the target sites in both promoters, since no inhibition occurs from small amounts of ORF5. This could explain why the amounts of the regulatory proteins ORF4 and ORF5 as well as their relative abundance have a significant effect on the choice of the open or repressed state of thepL promoter.
The expression of the pL fusion present in pPM127 can be induced from the repressed state to the derepressed state by addition of mitomycin C. This induction was found to be dependent on the presence of a functional RecA protein. When the shorter fusion pPM126 was used, no mitomycin C induction was observed. This could indicate that ORF5 is needed for mitomycin C induction; however, we cannot completely rule out the possibility that mitomycin C acts on a separate promoter present in the extra DNA found in pPM127 between bp 3458 and 3900. RecA has been shown to act by stimulating repressor self-cleavage in the investigated cases of LexA and λ repressor and is therefore mechanistically a coprotease (17; for a review, see reference 18). A consensus AG autodigestion site has not been found in ORF4 (19), but ORF4 could belong to a different protein family than the repressors analyzed so far. Furthermore, the possibility exists that RecA acts on a protein complex consisting of both ORF4 and ORF5.
A homology search between ORF4 and ORF5 of TP901-1 identified a region of ORF4 (from Tyr59 to Asp76) that is similar to the proposed helix-turn-helix motif in ORF5 and thus might constitute a DNA binding motif. This notion is further supported by the very high similarity of this region of ORF4 to the suggested helix-turn-helix motif of ORF4 (from Leu64 to Asp82) of Streptococcus thermophilus phage φO1205 (25). A high degree of similarity on the level of protein sequences is found between ORF4 and ORF5 of TP901-1 and φO1205, the degrees of identity being 48% in 122 amino acids for ORF4 and 47% in 67 amino acids for ORF5. Interestingly, ORF4 of φO1205 is considerably shorter than ORF4 from TP901-1 (134 versus 180 amino acids) and also does not contain a suggested RecA binding site. Furthermore, ORF4 and ORF5 show extensive homology to the repressor (50% identity in 168 amino acids) and protein 31a (75% identity in 44 amino acids), respectively, of Staphylococcus aureus phage phiPVL (16). In addition to this, the intergenic region located between orf4 and orf5 of TP901-1 and the repressor and 31a genes of phiPVL show 62% identity at the nucleotide level in a 66-bp region. The similarity between the TP901-1, φO1205, and phiPVL phages in this region indicates that they may share a common mechanism for the switch determining lysis or lysogeny, but this has to be proven experimentally. Future studies may enable us to determine whether a protein complex consisting of ORF4 and ORF5 is responsible for the features of this new type of phage switch exemplified by TP901-1.

ACKNOWLEDGMENTS

We acknowledge Patrick Duwat for donation of the recAstrain. We sincerely appreciate the expert technical assistance of Lise Sørensen and Lotte Bredahl.
This work is part of the FØTEK program supported by the Danish Dairy Research Foundation (Danish Dairy Board) and the Danish government through The Center of Advanced Food Studies. We also thank the Carlsberg Foundation for support.

REFERENCES

1.
Braun V., Hertwig S., Neve H., Geis A., and Teuber M. Taxonomic differentiation of bacteriophages of Lactococcus lactis by electron microscopy, DNA-DNA hybridization, and protein profiles.J. Gen. Microbiol.13519892551-2560
2.
Breüner A., Brøndsted L., and Hammer K. Novel organization of genes involved in prophage excision identified in the temperate lactococcal bacteriophage TP901-1.J. Bacteriol.18119997291-7297
3.
Brøndsted, L. Unpublished data.
4.
Christiansen B., Johnsen M. G., Stenby E., Vogensen F. K., and Hammer K. Characterization of the lactococcal temperate phage TP901-1 and its site-specific integration.J. Bacteriol.17619941069-1076
5.
Christiansen B., Brøndsted L., Vogensen F. K., and Hammer K. A resolvase-like protein is required for the site-specific integration of the temperate lactococcal bacteriophage TP901-1.J. Bacteriol.17819965164-5173
6.
de Ruyter P. G. G. A., Kuipers O. P., Beerthuyzen M. M., van Alen-Boerrigter I. J., and de Vos W. M. Functional analysis of promoters in the nisin cluster of Lactococcus lactis.J. Bacteriol.17819963434-3439
7.
de Ruyter P. G. G. A., Kuipers O. P., and de Vos W. M. Controlled gene expression systems for Lactococcus lactis with the food-grade inducer nisin.Appl. Environ. Microbiol.6219963662-3667
8.
Duwat P., Ehrlich S. D., and Gruss A. The recA gene of Lactococcus lactis: characterization and involvement in oxidative and thermal stress.Mol. Microbiol.1719951121-1131
9.
Gasson M. J. Plasmid complements of Streptococcus lactis NCDO 712 and other lactic streptococci after protoplast-induced curing.J. Bacteriol.15419831-9
10.
Hayes F., Daly C., and Fitzgerald G. F. Identification of the minimal replicon of Lactococcus lactis subsp. lactis UC317 plasmid pCI305.Appl. Environ. Microbiol.561990202-209
11.
Herskowitz I. and Hagen D. The lysis-lysogeny decision of phage λ: explicit programming and responsiveness.Annu. Rev. Genet.141980399-445
12.
Holo H. and Nes I. F. High-frequency transformation, by electroporation, of Lactococcus lactis subsp. cremoris grown with glycine in osmotically stabilized media.Appl. Environ. Microbiol.5519893119-3123
13.
Israelsen H., Madsen S. M., Vrang A., Hansen E. B., and Johansen E. Cloning and partial characterization of regulated promoters from Lactococcus lactis Tn917-lacZ integrants with the new promoter probe vector, pAK80.Appl. Environ. Microbiol.6119952540-2547
14.
Jensen P. R. and Hammer K. Minimal requirements for exponential growth of Lactococcus lactis.Appl. Environ. Microbiol.5919934363-4366
15.
Johnsen M. G., Appel K. F., Madsen P. L., Vogensen F. K., Hammer K., and Arnau J. A genomic region of lactococcal temperate bacteriophage TP901-1 encoding major virion proteins.Virology2181996306-315
16.
Kaneko J., Kimura T., Narita S., Tomita T., and Kamio Y. Complete nucleotide sequence and molecular characterization of the temperate staphylococcal bacteriophage phiPVL carrying Panton-Valentine leukocidin genes.Gene215199857-67
17.
Little J. W. Autodigestion of LexA and phage lambda repressors.Proc. Natl. Acad. Sci. USA8119841375-1379
18.
Little J. W. LexA cleavage and other self-processing reactions.J. Bacteriol.17519934943-4950
19.
Madsen P. L. and Hammer K. Temporal transcription of the lactococcal temperate phage TP901-1 and DNA sequence of the early promoter region.Microbiology14419982203-2215
20.
Marsch J. L., Erfle M., and Wykes E. J. The pIC plasmid and phage vectors with versatile cloning sites for recombinant selection by insertional inactivation.Gene321984481-485
21.
Miller J. H. Experiments in molecular genetics. 1972 Cold Spring Harbor Laboratory Press Cold Spring Harbor, N.Y
22.
Poteete A. R. Bacteriophage P22 The bacteriophages. Calendar R. 1988 647 -682 Plenum Press New York, N.Y
23.
Sambrook J., Fritsch E. F., and Maniatis T. Molecular cloning: a laboratory manual 2nd ed. 1989 Cold Spring Harbor Laboratory Cold Spring Harbor, N.Y
24.
Sanger F., Nicklen S., and Coulson A. R. DNA sequencing with chain-terminating inhibitors.Proc. Natl. Acad. Sci. USA7419775463-5467
25.
Stanley E., Fitzgerald G. F., Le Marrec C., Fayard B., and van Sinderen D. Sequence analysis and characterization of φO1205, a temperate bacteriophage infecting Streptococcus thermophilius CNRZ1205.Microbiology14319973417-3429
26.
Terzaghi B. E. and Sandine W. E. Improved medium for lactic streptococci and their bacteriophages.Appl. Microbiol.291975807-813

Information & Contributors

Information

Published In

cover image Journal of Bacteriology
Journal of Bacteriology
Volume 181Number 2415 December 1999
Pages: 7430 - 7438
PubMed: 10601198

History

Received: 7 June 1999
Accepted: 17 September 1999
Published online: 15 December 1999

Permissions

Request permissions for this article.

Contributors

Authors

Peter Lynge Madsen
Department of Microbiology, Technical University of Denmark, DK-2800 Lyngby, Denmark
Present address: Carlsberg Laboratory, Department of Yeast Genetics, DK-2500 Valby, Denmark.
Annette H. Johansen
Department of Microbiology, Technical University of Denmark, DK-2800 Lyngby, Denmark
Karin Hammer
Department of Microbiology, Technical University of Denmark, DK-2800 Lyngby, Denmark
Lone Brøndsted
Department of Microbiology, Technical University of Denmark, DK-2800 Lyngby, Denmark

Metrics & Citations

Metrics

Note:

  • For recently published articles, the TOTAL download count will appear as zero until a new month starts.
  • There is a 3- to 4-day delay in article usage, so article usage will not appear immediately after publication.
  • Citation counts come from the Crossref Cited by service.

Citations

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. For an editable text file, please select Medlars format which will download as a .txt file. Simply select your manager software from the list below and click Download.

View Options

Figures and Media

Figures

Media

Tables

Share

Share

Share the article link

Share with email

Email a colleague

Share on social media

American Society for Microbiology ("ASM") is committed to maintaining your confidence and trust with respect to the information we collect from you on websites owned and operated by ASM ("ASM Web Sites") and other sources. This Privacy Policy sets forth the information we collect about you, how we use this information and the choices you have about how we use such information.
FIND OUT MORE about the privacy policy