Skip to main content
Journal of Bacteriology logoLink to Journal of Bacteriology
. 2002 Aug;184(15):4288–4295. doi: 10.1128/JB.184.15.4288-4295.2002

Transcriptome and Proteome Analysis of Bacillus subtilis Gene Expression Modulated by Amino Acid Availability

Ulrike Mäder 1,*, Georg Homuth 1, Christian Scharf 1, Knut Büttner 1, Rüdiger Bode 1, Michael Hecker 1
PMCID: PMC135197  PMID: 12107147

Abstract

A comprehensive study of Bacillus subtilis gene expression patterns in response to amino acid availability was performed by means of proteomics and transcriptomics. The methods of two-dimensional protein gel electrophoresis and DNA macroarray technology were combined to analyze cells exponentially grown in minimal medium with and without 0.2% Casamino Acids (CAA). This approach revealed about 120 genes predominantly involved in amino acid biosynthesis, sporulation, and competence, which were downregulated in CAA-containing medium. Determination of sporulation frequencies confirmed the physiological relevance of the expression data.


The soil bacterium Bacillus subtilis is capable of synthesizing all proteinogenic amino acids. Besides their function as building blocks for cellular proteins, amino acids represent precursors in the biosynthesis of nucleotides and other cellular components such as cell wall polymers. Amino acid biosynthetic pathways are regulated on the level of enzyme activity as well as on the level of enzyme synthesis to ensure cellular adaptation to various requirements for amino acids under different growth conditions.

The expression of many amino acid biosynthetic genes in B. subtilis is controlled by transcription antitermination mechanisms: the ilv-leu operon (15, 27), the cysES operon (14), and the proBA operon (3) belong to the T-box family, which includes most of the aminoacyl-tRNA synthetase genes (17, 31). These genes are regulated by tRNA-mediated antitermination in response to starvation for a particular amino acid (16). The genes of the S-box regulon are controlled by a transcription antitermination system in response to methionine availability (18). The S-box-specific leader region elements were identified in 11 transcriptional units in the B. subtilis genome, whereby the majority of the 26 gene products fulfills functions in sulfate assimilation and methionine biosynthesis (18, 26, 29). Regulation of tryptophan biosynthetic genes by transcription attenuation and translation control mechanisms is mediated by the RNA binding protein TRAP as well as by a T-box-dependent regulatory mechanism (reviewed in reference 1). Expression of the lysC gene encoding the lysine feedback-controlled aspartokinase II (30) is regulated by lysine availability via an antitermination system, too (20). In B. subtilis, the arginine biosynthetic operons are repressed by the AhrC regulatory protein, which is activated in the presence arginine (6, 37).

In this study we report on the gene expression profile of B. subtilis exponentially grown in minimal medium with and without 0.2% Casamino Acids (CAA), thereby providing an insight into the response of B. subtilis to different amino acid availabilities. The genes which were differentially expressed under the two growth conditions included those for amino acid biosynthesis, sporulation, and competence development.

Characterization of the proteome under conditions of different amino acid availability.

The B. subtilis 168 strain was cultivated aerobically at 37°C in a minimal medium (pH 7.5) containing 50 mM Tris, 8 mM MgSO4, 13 mM KCl, 18 mM NaCl, 0.6 mM KH2PO4, 2 mM CaCl2, 0.001 mM FeSO4, 0.01 mM MnSO4, 10 mM glutamine, 0.2% (wt/vol) glucose, and 0.8 mM tryptophan. Cells were grown in the presence or absence of 0.2% CAA (vitamin free; Difco, Detroit, Mich.) and harvested in the exponential growth phase after reaching an optical density at 500 nm of 0.5. Compared to the cultures in minimal medium (G = 45 min), shorter generation times (G = 25 min) were observed for the cultures in CAA-containing minimal medium (see Fig. 2). Preparation of protein extracts and two-dimensional protein gel electrophoresis were performed as previously described (2). About 65 protein spots present on the control gel in the pH range of 4 to 7 decreased in intensity or were completely absent when the medium was supplemented with CAA (Fig. 1A). In addition, narrow pH gradient gels were utilized in the pH range of 4.5 to 5.5, which allows for a better resolution of the most overcrowded region of the pH 4 to 7 gels (Fig. 1B). Altogether, 58 protein spots representing 50 different proteins downregulated by CAA (Table 1 laser desorption ionization-time of flight mass spectrometry as previously described (2).

FIG. 2.

FIG. 2.

Growth curve (closed symbols) and sporulation frequencies (open symbols) of B. subtilis 168 cultivated in minimal medium (circles) and in minimal medium supplemented with 0.2% CAA (diamonds). The number of spores per milliliter of culture was determined as the number of heat-resistant (80°C for 30 min) CFU on Luria-Bertani plates, and the number of viable cells was determined as the total number of CFU (before heat treatment). Sporulation frequencies were defined as the percentage of heat-resistant CFU. Four independent experiments,which gave comparable results, were carried out.

FIG. 1.

FIG. 1.

Dual-channel image analysis of two-dimensional protein patterns of B. subtilis 168 exponentially grown in minimal medium and minimal medium supplemented with 0.2% CAA. Cytosolic protein extracts were separated by two-dimensional gel electrophoresis in the pH gradient of 4 to 7 (A) and in the pH gradient of 4.5 to 5.5 (B). Dual-channel images of the silver-stained gels were created by computer-aided transformation of the gel images by using the software DECODON Delta2D (DECODON GmbH, Greifswald, Germany). Red spots represent proteins whose synthesis was decreased in the presence of CAA.

TABLE 1.

Genes with significantly higher expression during exponential growth of B. subtilis in minimal medium without CAA as revealed by transcriptome and proteome analysesa

Gene and category Function Induction (expt 1, expt 2) Transcriptional organization
Amino acid biosynthesis
    argC N-acetylglutamate gamma-semialdehyde dehydrogenase 6.0, 3.8 argC-argJ-argB-argD-carA-carB-argF
    argJ Ornithine acetyltransferase; amino acid acetyltransferase 20.5, 11.6 argC-argJ-argB-argD-carA-carB-argF
    argB N-acetylglutamate 5-phosphotransferase 29.3, 34.2 argC-argJ-argB-argD-carA-carB-argF
    argD N-acetylornithine aminotransferase 4.7, 8.6 argC-argJ-argB-argD-carA-carB-argF
    carA Carbamoyl-phosphate transferase (subunit A) 30.0, 16.2 argC-argJ-argB-argD-carA-carB-argF
    carB Carbamoyl-phosphate transferase (subunit B) 4.6, 5.1 argC-argJ-argB-argD-carA-carB-argF
    argF Ornithine carbamyoltransferase 5.7, 9.6 argC-argJ-argB-argD-carA-carB-argF
    argG Argininosuccinate synthase 17.3, 23.9 argG-argH-ytzD
    argH Argininosuccinate lyase 21.4, 27.5 argG-argH-ytzD
    ytzD Unknown function —, — argG-argH-ytzD
    cysH Phosphoadenosine phosphosulfate sulfotransferase 3.3, 3.8 cysH-cysP-sat-cysC-ylnD-ylnE-ylnF
    cysP Sulfate permease 5.5, 3.3 cysH-cysP-sat-cysC-ylnD-ylnE-ylnF
    sat Sulfate adenylyltransferase 4.7, 4.6 cysH-cysP-sat-cysC-ylnD-ylnE-ylnF
    cysC Adenylylsulfate kinase 2.7, 4.7 cysH-cysP-sat-cysC-ylnD-ylnE-ylnF
    ylnD Similar to uroporphyrine-III C-methyltransferase 5.7, 5.2 cysH-cysP-sat-cysC-ylnD-ylnE-ylnF
    ylnE Unknown function 4.3, 3.1 cysH-cysP-sat-cysC-ylnD-ylnE-ylnF
    ylnF Similar to uroporphyrine-III C-methyltransferase 4.0, 3.6 cysH-cysP-sat-cysC-ylnD-ylnE-ylnF
    cysK Cysteine synthase A 2.0, 2.2 cysK
    hisZ Histidyl-tRNA synthetase —, — hisZ-hisG-hisD-hisB-hisH-hisA-hisF-hisI
    hisG ATP phosphoribosyltransferase —, 1.1 hisZ-hisG-hisD-hisB-hisH-hisA-hisF-hisI
    hisD Histidinol dehydrogenase 4.5, 7.2 hisZ-hisG-hisD-hisB-hisH-hisA-hisF-hisI
    hisB Imidazoleglycerol-phosphate dehydratase 5.7, 5.6 hisZ-hisG-hisD-hisB-hisH-hisA-hisF-hisI
    hisH Amidotransferase 5.2, 5.1 hisZ-hisG-hisD-hisB-hisH-hisA-hisF-hisI
    hisA Phosphoribosylformimino-5-aminoimidazole carboxamideribotide isomerase 5.5, 7.3 hisZ-hisG-hisD-hisB-hisH-hisA-hisF-hisI
    hisF Cyclase-like protein (synthesis of imidazole glycerol phosphate) 4.2, 6.6 hisZ-hisG-hisD-hisB-hisH-hisA-hisF-hisI
    hisI Phosphoribosyl-ATP pyrophosphohydrolase; phosphribosyl-AMP cyclohydrolase 8.5, 10.4 hisZ-hisG-hisD-hisB-hisH-hisA-hisF-hisI
    hom Homoserine dehydrogenase 2.2, 2.1 hom-thrC-thrB
    thrC Threonine synthase 2.0, 2.2 hom-thrC-thrB
    thrB Homoserine kinase 3.2, 3.5 hom-thrC-thrB
    ilyA Threonine dehydratase 3.9, 3.6 ilvA-ypmP
    ypmP Unknown function —, — ilvA-ypmP
    ilvB Acetolactate synthase (large subunit) 7.1, 7.9 ilvB-ilvH-ilvC-leuA-leuB-leuC-leuD
    ilvH Acetolactate synthase (small subunit) 11.0, 7.0 ilvB-ilvH-ilvC-leuA-leuB-leuC-leuD
    ilvC Ketol-acid reductoisomerase 9.9, 10.1 ilvB-ilvH-ilvC-leuA-leuB-leuC-leuD
    leuA 2-Isopropylmalate synthase 9.5, 9.7 ilvB-ilvH-ilvC-leuA-leuB-leuC-leuD
    leuB 3-Isopropylmalate dehydrogenase 9.6, 10.6 ilvB-ilvH-ilvC-leuA-leuB-leuC-leuD
    leuC 3-Isopropylmalate dehydratase (large subunit) 10.0, 10.9 ilvB-ilvH-ilvC-leuA-leuB-leuC-leuD
    leuD 3-Isopropylmalate dehydratase (small subunit) 7.0, 10.5 ilvB-ilvH-ilvC-leuA-leuB-leuC-leuD
    ilvD Dihydroxyacid dehydratase 4.6, 4.4 ilvD
    lvsC Aspartokinase II (alpha and beta subunit) 8.5, 13.4 lysC
    metE Cobalamin-independent methionine synthase 56.8, 59.9 metC
    ybgE Similar to branched chain amino acid aminotransferase 3.0, 4.1 ybgE
    yitJ Similar to methionine synthase 4.7, — yitJ
    yjcI Similar to cystathionine gamma-synthase 9.9, 6.9 yjcI-yjcJ
    yjcJ Similar to cystathionine beta-lyase 14.0, 9.9 yjcI-yjcJ
    yoaD Similar to phosphoglycerate dehydrogenase 3.1, — yoaD-yoaC-yoaB
    yoaC Similar to xylulokinase 2.5, 3.2 yoaD-yoaC-yoaB
    yoaB Unknown function 14.5, 9.2 yoaD-yoaC-yoaB
    yvgR Similar to sulfite reductase 2.1, 2.9 yvgR-yvgQ
    yvgQ Similar to sulfite reductase 5.0, 8.6 yvgR-yvgQ
    yxjG Similar to methionine synthase 10.5, 7.6 yxjG
    yxjH Similar to methionine synthase —, — yxjH
Competence
    comER Nonessential gene for competence 3.2, 3.4 comER
    comGA Required for exogenous DNA binding 2.0, 5.3 comGA-comGB-comGC-comGD-comGE-comGF-comGG
    comGB Required for exogenous DNA binding 2.2, 4.2 comGA-comGB-comGC-comGD-comGE-comGF-comGG
    comGC Required for exogenous DNA binding 3.2, 7.1 comGA-comGB-comGC-comGD-comGE-comGF-comGG
    comGD Required for exogenous DNA binding 3.0, 5.6 comGA-comGB-comGC-comGD-comGE-comGF-comGG
    comGE Required for exogenous DNA binding —, — comGA-comGB-comGC-comGD-comGE-comGF-comGG
    comGF Required for exogenous DNA binding 4.2, 4.2 comGA-comGB-comGC-comGD-comGE-comGF-comGG
    comGG Required for exogenous DNA binding 2.4, 4.3 comGA-comGB-comGC-comGD-comGE-comGF-comGG
    comK Competence transcription factor 3.1, 4.8 comK
Transition state functions and sporulation
    appD Oligopeptide ABC transporter (ATP-binding protein) —, — appD-appF-appA-appB-appC
    appF Oligopeptide ABC transporter (ATP-binding protein) 0.7, 0.9 appD-appF-appA-appB-appC
    appA Oligopeptide ABC transporter (peptide-binding protein) 21.9, 22.5 appD-appF-appA-appB-appC
    appB Oligopeptide ABC transporter (permease) —, 5.7 appD-appF-appA-appB-appC
    appC Oligopeptide ABC transporter (permease) —, — appD-appF-appA-appB-appC
    cotE Spore coat protein 13.5, 5.8 cotE
    cotV Spore coat protein 6.1, — cotV-cotW-cotX
    cotW Spore coat protein 3.5, 3.5 cotV-cotW-cotX
    cotX Spore coat protein 7.5, — cotV-cotW-cotX
    dacF Penicillin binding protein; required for spore cortex synthesis 4.6, 2.4 dacF-spoIIAA-spoIIAB-sigF
    spoIIAA Anti-anti-sigma factor (antagonist of SpoIIAB) 0.9, 0.9 dacF-spoIIAA-spoIIAB-sigF
    spoIIAB Anti-sigma factor (antagonist of sigma F); serine kinase —, — dacF-spoIIAA-spoIIAB-sigF
    sigF RNA polymerase sporulation-specific sigma factor 5.9, 5.3 dacF-spoIIAA-spoIIAB-sigF
    gerM Germination (cortex hydrolysis) and sporulation (putative role in peptidoglycan synthesis) 3.3, 3.8 gerM
    prkA Serine protein kinase 2.5, 2.5 prkA
    qcrA Menaquinol:cytochrome c oxidoreductase (iron-sulfur subunit) 4.1, 4.0 qcrA-qcrB-qcrC
    qcrB Menaquinol:cytochrome c oxidoreductase (cytochrome b subunit) —, — qcrA-qcrB-qcrC
    qcrC Menaquinol:cytochrome c oxidoreductase (cytochrome c subunit) 3.1, 5.3 qcrA-qcrB-qcrC
    rapA Aspartyl phosphate phosphatase 3.2, 3.4 rapA-phrA
    phrA Inhibitor of the activity of phosphatase RapA 4.1, 1.8 rapA-phrA
    rsfA Probable transcriptional regulator of sigma F-dependent genes 4.5, 4.2 rsfA
    sp0A Two-component response regulator 2.9, 2.5 spo0A
    spoIIB Required for endospore development 6.4, 3.7 spoIIB
    spoIIIAG Mutants block sporulation after engulfment 5.5, 4.3 spoIIIAG-spoIIIAH
    spoIIIAH Mutants block sporulation after engulfment 7.8, 4.3 spoIIIAG-spoIIIAH
    usd Required for translation of spoIIID —, — usd-spoIIID
    spoIIID Transcriptional regulator of sigma E- and sigma K-dependent genes 10.2, 4.7 usd-spoIIID
    spoIVA Required for proper spore cortex formation and coat assembly 2.4, 2.5 spoIVA
    spoVID Required for assembly of the spore coat 3.2, 3.5 spoVID-ysxE
    ysxE Unknown function —, — spoVID-ysxE
    sspE Small acid-soluble spore protein 4.3, 2.5 sspE
    yjbX Unknown function; glutamic acid-rich protein 7.2, 6.2 yjbX
    ylaK Similar to phosphate starvation inducible protein 6.5, 4.7 ylaK
    ylbO Unknown function 6.8, 3.0 ylbO
    yqxM Unknown function 0.9, 1.1 yqxM-sipW-tasA
    sipW Type I signal peptidase 0.9, 1.3 yqxM-sipW-tasA
    tasA Spore-associated antimicrobial protein required for spore coat assembly 2.0, 3.1 yqxM-sipW-tasA
    ytfI Unknown function —, — ytfI-ytfJ
    ytfJ Unknown function 3.7, 3.8 ytfI-ytfJ
    yuiC Unknown function 3.1, 3.0 yuiC
    ywcI Unknown function 10.1, 7.3 ywcI-sacT
    sacT Transcriptional antiterminator (regulation of the sacP operon) 9.9, 6.4 ywcI-sacT
    yvyD Similar to a sigma 54 modulating factor 2.4, 3.4 yvyD
Other functions
    dat Probable d-alanine aminotransferase 1.5, 1.7 dat
    gcvT Probable aminomethyltransferase 1.9, 1.9 gcvT-gcvPA-gcvPB
    gcvPA Probable glycine decarboxylase (subunit 1) 2.8, 2.6 gcvT-gcvPA-gcvPB
    gcvPB Probable glycine decarboxylase (subunit 2) 2.4, 2.6 gcvT-gcvPA-gcvPB
    mpr Extracellular metalloprotease 6.6, 3.2 mpr-ybfJ
    ybfJ Unknown function —, — mpr-ybfJ
    yfhK Similar to cell division inhibitor 3.2, 6.5 yfhK
    yhgB Unknown function —, — yhgB-yhfA-yhaA
    yhfA Unknown function —, — yhgB-yhfA-yhaA
    yhaA Similar to aminoacylase 4.2, 4.5 yhgB-yhfA-yhaA
    ykrT Unknown function 1.6, 1.7 ykrT-ykrS
    ykrS Similar to initiation factor eIF-2B (alpha subunit) 2.5, 3.5 ykrT-ykrS
    ykrW Similar to ribulose-bisphosphate carboxylase 2.7, 3.3 ykrW-ykrX-ykrY-ykrZ
    ykrX Unknown function 3.9, 4.9 ykrW-ykrX-ykrY-ykrZ
    ykrY Unknown function 6.8, 4.8 ykrW-ykrX-ykrY-ykrZ
    ykrZ Unknown function 3.1, 2.1 ykrW-ykrX-ykrY-ykrZ
    ykuN Similar to flavodoxin —, — ykuN-ykuO-ykuP-ykuQ
    ykuO Unknown function 2.8, 2.9 ykuN-ykuO-ykuP-ykuQ
    ykuP Similar to flavodoxin 4.7, 3.5 ykuN-ykuO-ykuP-ykuQ
    ykuQ Similar to tetrahydrodipicolinate succinylase 0.9, 1.3 ykuN-ykuO-ykuP-ykuQ
    ykwC Similar to 3-hydroxyisobutyrate dehydrogenase 1.7, 1.9 ykwC
    yodF Similar to proline permease 4.5, 4.4 yodF
    yojA Similar to gluconate permease —, — yojA-yojB-yojC
    yojB Unknown function 3.8, 3.9 yojA-yojB-yojC
    yojC Unknown function —, — yojA-yojB-yojC
    yqiX Similar to amino acid ABC transporter (binding protein) 7.3, 10.3 yqiX-yqiY-yqiZ
    yqiY Similar to amino acid ABC transporter (permease) 5.3, 7.1 yqiX-yqiY-yqiZ
    yqiZ Similar to amino acid ABC transporter (ATP binding) 7.1, 16.0 yqiX-yqiY-yqiZ
    yuaF Unknown function —, — yuaF-yuaG-yuaI
    yuaG Similar to epidermal surface antigen 3.4, 3.1 yuaF-yuaG-yuaI
    yuaI Unknown function —, — yuaF-yuaG-yuaI
    yvaC Unknown function —, — yvaC-yvaB
    yvaB Similar to NAD(P)H dehydrogenase (quinone) 2.3, 1.7 yvaC-yvaB
    ywfH Similar to 3-oxoacyl-acyl-carrier protein reductase 3.8, 3.2 ywfH
Unknown functions
    ybdO Unknown function 3.2, 7.2 ybdO
    yfhB Unknown function 3.7, 3.0 yfhB
    yfmA Unknown function —, — yfmA-yflT
    yflT Unknown function 4.6, 40.9 yfmA-yflT
    yjcE Unknown function 3.8, 3.6 yjcE-yjcD
    yjcD Similar to ATP-dependent DNA helicase 0.7, 1.4 yjcE-yjcD
    yjcH Unknown function 1.7, 3.1 yjcH-yjcG-yjcF
    yjcG Unknown function 3.7, 3.6 yjcH-yjcG-yjcF
    yjcF Unknown function 2.4, 1.8 yjcH-yjcG-yjcF
    ykvR Unknown function 3.2, 3.3 ykvR
    ylaJ Unknown function 3.1, 3.0 ylaJ
    ylqB Unknown function 3.3, 3.8 ylqB
    yqgZ Unknown function 4.6, 3.4 yqgZ
    yvaW Unknown function —, — yvaW-yvaX-yvaY
    yvaX Unknown function —, — yvaW-yvaX-yvaY
    yvaY Unknown function 10.6, 12.2 yvaW-yvaX-yvaY
a

Significantly regulated genes are given in bold face. Significant regulation was defined as at least threefold changes in the mRNA levels in both macroarray experiments. Genes were also regarded as significantly regulated when at least twofold changes in the mRNA levels were confirmed by the proteome analysis or the operon structure. Underlined gene names indicate higher expression in minimal medium without CAA as revealed by the proteome analysis. The calculated expression level ratios are shown for both independent macroarray experiments in the column “induction,” whereby dashes indicate that specific signals for these genes were below the significance threshold. The putative functions of the y-gene-encoded proteins were obtained from the SubtiList database.

Of these 50 proteins, 35 represented amino acid biosynthetic enzymes involved in the synthesis of lysine, methionine, threonine, arginine, cysteine, histidine, leucine, isoleucine, and valine. Furthermore, the sporulation proteins SpoIVA, SpoOA, and TasA, the serine protein kinase PrkA, the proteins Dat and GcvT involved in metabolism of amino acids, and 9 proteins with still-unknown functions (Y-proteins) were identified to be downregulated by addition of CAA.

Analysis of the transcriptome under conditions of differentamino acid availability.

B. subtilis strain 168 was cultivated in the described minimal medium with and without 0.2% CAA. Total RNA was isolated from exponentially growing cells (optical density at 500 nm of 0.5) and was checked by Northern blot analysis (data not shown). Cell harvesting, preparation of RNA, and macroarray analysis with Panorama B. subtilis gene arrays and specific cDNA labeling primers (Sigma-Genosys, The Woodlands, Tex.) were performed as described by Eymann et al. (10). Two macroarray experiments were carried out by using independently isolated RNA preparations and different array batches. Quantification of hybridization signals, background subtraction, and calculation of normalized intensity values of the individual spots were performed with the ArrayVision software (version 5.1; Imaging Research, St. Catherines, Ontario, Canada) as described by Eymann et al. (10). Expression level ratios of three or more in both independent experiments were considered significant. Final evaluation of the macroarray data included the consideration of putative operons derived from the genome sequence, using the SubtiList database (http://genolist.pasteur.fr/SubtiList/) as well as previously known transcriptional units.

Scatter plots comparing the normalized intensity values revealed that mRNA levels of the majority of genes did not differ significantly between both growth conditions, whereas about 100 genes were expressed at a level more than threefold higher in minimal medium without CAA (data not shown). According to the criteria specified in the footnote to Table 1, altogether 114 genes showed significantly elevated expression levels. Of these genes, most encode proteins with functions in amino acid biosynthesis (42 genes), transition state processes and sporulation (32 genes), and competence (8 genes). The patterns of CAA-regulated genes found by the proteomic and transcriptomic approaches were similar, whereby about 50% of the differentially expressed genes could be detected in the proteome analysis. Interestingly, only three genes (guaC, purK, and yxjA) involved in nucleotide metabolism and transport were expressed at a significantly higher level during growth in CAA-containing medium.

At the mRNA level, 16 transcriptional units involved in amino acid biosynthesis were identified to be significantly downregulated by CAA (Table 1). Previous studies suggested regulation by amino acid availability for the operons argCJBD-carAB-argF, argGH (37), hom-thrCB (40), and ilvBHC-leuABCD (15), the lysC gene (20), and also the S-box-regulated transcriptional units cysHP-sat-cysC-ylnDEF, yjcIJ, yoaDCB, ykrWXYZ, ykrTS, metE, yitJ, and yxjG. In this study, of the 11 transcriptional units potentially belonging to the S box regulon (18) the 8 mentioned above were expressed at a significantly lower level in CAA-containing medium. Of the approximately 50 B. subtilis genes involved in amino acid transport, only the yqiXYZ operon encoding an amino acid ABC transport system (33) and the monocistronic-transcribed yodF gene encoding a putative proline permease showed significantly different expression levels in response to amino acid availability. Like the argCJBD-carAB-argF and argGH operons, the yqiXYZ operon is preceded by an AhrC recognition site (25). As a further result of the transcriptome study, addition of 0.2% CAA did not affect expression of glycine, serine, proline, tyrosine, and phenylalanine biosynthetic genes. Glutamine and tryptophan biosynthesis was not expected to be regulated under the conditions compared in this study because of the presence of these amino acids in both cultivation media.

A second group of genes expressed at a significantly higher level during growth in the absence of CAA encodes proteins with functions in competence development. This group includes the comK gene encoding the competence transcription factor that activates the genes involved in DNA binding and uptake (reviewed in reference 8). Furthermore, nearly the complete comG operon (4) shared the same expression pattern. The srfA operon also involved in competence regulation exhibited an approximately twofold increased mRNA level. These results are in agreement with previous observations that competence is repressed by the addition of CAA during exponential growth of B. subtilis in minimal medium (35).

As shown in Table 1, many B. subtilis genes that encode products with functions in transient-phase adaptation and sporulation exhibited significantly higher mRNA levels in minimal medium without CAA. Among these genes were the Spo0A dependently regulated operons appDFABC (21, 22), qcrABC (41), and ywcI-sacT (11). The phosphorylated response regulator Spo0A activates transcription of many sporulation genes and negatively regulates genes preventing sporulation, such as abrB (19). The abrB gene encodes a repressor of early-stationary-phase and sporulation genes, including the sigH gene. Thus, by repressing abrB transcription Spo0A-P stimulates synthesis of σH and thereby enhances its own transcription. Besides spo0A itself, the σH-dependently transcribed genes tasA (34, 38), spoVG, and yvyD were upregulated in the absence of CAA. Most of the sporulation genes significantly upregulated in the absence of CAA belong to the σE regulon (U. Völker, personal communication), which comprises early-mother-cell-specific genes.

Sporulation frequency in response to amino acid availability.

Sporulation genes were shown to be regulated in response to amino acid availability during exponential growth of B. subtilis. To verify the physiological relevance of the proteome and transcriptome data, sporulation frequencies were determined at several time points during growth of B. subtilis in the described minimal medium with and without CAA. The cultures grown under the two different conditions were inoculated with the same preculture. As shown in Fig. 2, the sporulation frequency of the minimal medium culture without CAA increased continuously up to 14 h after inoculation, whereas addition of CAA almost completely prevented the appearance of spores at least up to 14 h after inoculation. These data confirmed that a significantly lower amount of the population enters the sporulation process during exponential growth in the presence of CAA.

Concluding remarks.

Several adaptive processes are involved in the response of B. subtilis to growth-limiting levels of nutrients. As revealed by this study, expression of sporulation and competence genes is affected by amino acid availability during exponential growth of B. subtilis. Due to the fact that a relatively small portion of the culture enters sporulation in the exponential growth phase, only strongly expressed sporulation genes caused significant signals in the transcriptome analysis.

In agreement with these results, Cosby and Zuber (5) described a negative effect of amino acids on expression of early-stage sporulation genes. They reported that addition of CAA to minimal medium affects sigH expression as well as σH-dependent transcription. The alternative σ factor σH is required for the transcription of early-stationary-phase and sporulation genes and represents the first σ factor in a gene expression cascade resulting in spore formation (23). As reported by Eymann et al. (9), a few σH-dependent genes (spo0A, spoVG, yvyD, and ytxGHI) are induced in response to amino acid starvation in a RelA-dependent manner. It was shown earlier that a relA mutant sporulates less effectively than the wild type after a shift down from CAA-containing medium to medium without CAA (24). The ribosome-bound ppGpp synthetase RelA is activated by uncharged tRNAs or by glucose starvation. Increased ppGpp levels mediate the stringent control which allows adaptation of cell growth to the present nutrient conditions. Cells growing in minimal medium might be partially starved for amino acids, which possibly elevates the basal ppGpp level. In this study, higher expression of the σH-dependent genes yvyD, spo0A, spoVG, and tasA in minimal medium without CAA was shown during exponential growth. It is interesting that almost the same σH-dependent genes which are induced in a RelA-dependent manner in amino-acid-starved cells were expressed at significantly higher levels during growth in minimal medium without CAA. In B. subtilis the CodY regulator mediates amino acid repression of several genes involved in nitrogen metabolism (7, 12, 13, 36, 39) as well as competence development (35), motility (28), and sporulation. Recently, Ratnayake-Lecamwasam et al. (32) reported that CodY represents a GTP-sensing protein and functions as a repressor under conditions of high GTP levels. They suggested that the stringent response might be involved in the inactivation of the CodY regulator by decreasing the cellular GTP pool. As spo0A is repressed by CodY (32), a decrease in the GTP level might result in enhanced sigH transcription. In the present study, expression of the genes spo0A and comK as well as the srfA operon was shown to be downregulated in CAA-containing medium, whereas most of the other known CodY-dependent genes did not share the same regulatory pattern, indicating the multiple regulation of CodY controlled genes.

Acknowledgments

This work was supported from grants of the DFG, the BMBF, and the Fonds der Chemischen Industrie to M.H.

We are indebted to U. Völker for valuable information concerning the sporulation sigma factor regulons.

REFERENCES

  • 1.Babitzke, P., and P. Gollnick. 2001. Posttranscription initiation control of tryptophan metabolism in Bacillus subtilis by the trp RNA-binding attenuation protein (TRAP), anti-TRAP, and RNA structure. J. Bacteriol. 183:5795-5802. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Büttner, K., J. Bernhardt, C. Scharf, R. Schmid, U. Mäder, C. Eymann, H. Antelmann, U. Völker, A. Völker, and M. Hecker. 2001. A comprehensive two-dimensional map of cytosolic proteins of Bacillus subtilis. Electrophoresis 22:2908-2935. [DOI] [PubMed] [Google Scholar]
  • 3.Chopin, A., V. Biaudet, and S. D. Ehrlich. 1998. Analysis of the Bacillus subtilis genome sequence reveals nine new T-box leaders. Mol. Microbiol. 29:662-664. [DOI] [PubMed] [Google Scholar]
  • 4.Chung, Y. S., and D. Dubnau. 1998. All seven comG open reading frames are required for DNA binding during transformation of competent Bacillus subtilis. J. Bacteriol. 180:41-45. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Cosby, W. M., and P. Zuber. 1997. Regulation of Bacillus subtilis σH (spo0H) and AbrB in response to changes in external pH. J. Bacteriol. 179:6778-6787. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Czaplewski, L. G., A. K. North, M. C. Smith, S. Baumberg, and P. G. Stockley. 1992. Purification and initial characterization of AhrC: the regulator of arginine metabolism genes in Bacillus subtilis. Mol. Microbiol. 6:267-275. [DOI] [PubMed] [Google Scholar]
  • 7.Debarbouille, M., R. Gardan, M. Arnaud, and G. Rapoport. 1999. Role of bkdR, a transcriptional activator of the sigL-dependent isoleucine and valine degradation pathway in Bacillus subtilis. J. Bacteriol. 181:2059-2066. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Dubnau, D., and K. Turgay. 2000. Regulation of competence in Bacillus subtilis and its relation to stress response, p. 249-260. In G. Storz and R. Hengge-Aronis (ed.), Bacterial stress responses. ASM Press, Washington, D.C.
  • 9.Eymann, C., G. Mittenhuber, and M. Hecker. 2001. The stringent response, σH-dependent gene expression and sporulation in Bacillus subtilis. Mol. Gen. Genet. 264:913-923. [DOI] [PubMed] [Google Scholar]
  • 10.Eymann, C., G. Homuth, C. Scharf, and M. Hecker. 2002. Bacillus subtilis functional genomics: characterization of the stringent response by proteome and transcriptome analysis. J. Bacteriol. 184:2500-2520. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Fawcett, P., P. Eichenberger, R. Losick, and P. Youngman. 2000. The transcriptional profile of early to middle sporulation in Bacillus subtilis. Proc. Natl. Acad. Sci. USA 97:8063-8068. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Ferson, A. E., L. V. Wray, Jr., and S. H. Fisher. 1996. Expression of the Bacillus subtilis gabP gene is regulated independently in response to nitrogen and amino acid availability. Mol. Microbiol. 22:693-701. [DOI] [PubMed] [Google Scholar]
  • 13.Fisher, S. H., K. Rohrer, and A. E. Ferson. 1996. Role of CodY in regulation of the Bacillus subtilis hut operon. J. Bacteriol. 178:3779-3784. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Gagnon, Y., R., R. Breton, H. Putzer, M. Pelchat, M. Grunberg-Manago, and J. Lapointe. 1994. Clustering and cotranscription of the Bacillus subtilis genes encoding the aminoacyl-tRNA synthetases specific for glutamate and for cysteine and the first enzyme for cysteine biosynthesis. J. Biol. Chem. 269:2473-2482. [PubMed] [Google Scholar]
  • 15.Grandoni, J. A., S. A. Zahler, and J. M. Calvo. 1992. Transcriptional regulation of the ilv-leu operon of Bacillus subtilis. J. Bacteriol. 174:3212-3219. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Grundy, F. J., and T. M. Henkin. 1993. tRNA as a positive regulator of transcription antitermination in B. subtilis. Cell 74:475-482. [DOI] [PubMed] [Google Scholar]
  • 17.Grundy, F. J., and T. M. Henkin. 1994. Conservation of a transcription antitermination mechanism in aminoacyl-tRNA synthetase and amino acid biosynthesis genes in gram-positive bacteria. J. Mol. Biol. 235:798-804. [DOI] [PubMed] [Google Scholar]
  • 18.Grundy, F. J., and T. M. Henkin. 1998. The S box regulon: a new global transcription termination control system for methionine and cysteine biosynthesis genes in gram-positive bacteria. Mol. Microbiol. 30:737-749. [DOI] [PubMed] [Google Scholar]
  • 19.Hoch, J. 1995. Control of cellular development in sporulating bacteria by the phosphorelay two-component signal transduction system, p. 129-144. In J. Hoch and T. Sihavy (ed.), Two-component signal transduction. ASM Press, Washington, D.C.
  • 20.Kochhar, S., and H. Paulus. 1996. Lysine-induced premature transcription termination in the lysC operon of Bacillus subtilis. Microbiology 142:1635-1639. [DOI] [PubMed] [Google Scholar]
  • 21.Koide, A., and J. A. Hoch. 1994. Identification of a second oligopeptide transport system in Bacillus subtilis and determination of its role in sporulation. Mol. Microbiol. 13:417-426. [DOI] [PubMed] [Google Scholar]
  • 22.Koide, A., M. Perego, and J. A. Hoch. 1999. ScoC regulates peptide transport and sporulation initiation in Bacillus subtilis. J. Bacteriol. 181:4114-4117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Kroos, L., and Y. T. Yu. 2000. Regulation of sigma factor activity during Bacillus subtilis development. Curr. Opin. Microbiol. 3:553-560. [DOI] [PubMed] [Google Scholar]
  • 24.Lopez, J. M., A. Dromerick, and E. Freese. 1981. Response of guanosine 5′-triphosphate concentration to nutritional changes and its significance for Bacillus subtilis sporulation. J. Bacteriol. 146:605-613. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Makarova, K. S., A. A. Mironov, and M. S. Gelfand. 2001. Conservation of the binding site for the arginine repressor in all bacterial lineages. Genome Biol. 2:0013.1-0013.8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Mansilla, M. C., D. Albanesi, and D. de Mendoza. 2000. Transcriptional control of the sulfur-regulated cysH operon, containing genes involved in l-cysteine biosynthesis in Bacillus subtilis. J. Bacteriol. 182:5885-5892. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Marta, P. T., R. D. Ladner, and J. A. Grandoni. 1996. A CUC triplet confers leucine-dependent regulation of the Bacillus subtilis ilv-leu operon. J. Bacteriol. 178:2150-2153. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Mirel, D. B., W. F. Estacio, M. Mathieu, E. Olmsted, J. Ramirez, and L. M. Marquez-Magana. 2000. Environmental regulation of Bacillus subtilis σD-dependent gene expression. J. Bacteriol. 182:3055-3062. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Murphy, B. A., F. J. Grundy, and T. M. Henkin. 2002. Prediction of gene function in methylthioadenosine recycling from regulatory signals. J. Bacteriol. 184:2314-2318. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Paulus, H. 1993. Biosynthesis of the aspartate family of amino acids, p. 237-267. In A. L. Sonenshein, J. A. Hoch, and R. Losick (ed.), Bacillus subtilis and other gram-positive bacteria: physiology, biochemistry, and Molecular Genetics. ASM Press, Washington, D.C.
  • 31.Pelchat, M., and J. Lapointe. 1999. Aminoacyl-tRNA synthetase genes of Bacillus subtilis: organization and regulation. Biochem. Cell Biol. 77:343-347. [PubMed] [Google Scholar]
  • 32.Ratnayake-Lecamwasam, M., P. Serror, K. W. Wong, and A. L. Sonenshein. 2001. Bacillus subtilis CodY represses early-stationary-phase genes by sensing GTP levels. Genes Dev. 15:1093-1103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Sekowska, A., S. Robin, J. J. Daudin, A. Henaut, and A. Danchin. 2001. Extracting biological information from DNA arrays: an unexpected link between arginine and methionine metabolism in Bacillus subtilis. Genome Biol. 2:0019.1-0019.12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Serrano, M., R. Zilhao, E. Ricca, A. J. Ozin, C. P. Moran, Jr., and A. O. Henriques. 1999. A Bacillus subtilis secreted protein with a role in endospore coat assembly and function. J. Bacteriol. 181:3632-3643. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Serror, P., and A. L. Sonenshein. 1996. CodY is required for nutritional repression of Bacillus subtilis genetic competence. J. Bacteriol. 178:5910-5915. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Slack, F. J., P. Serror, E. Joyce, and A. L. Sonenshein. 1995. A gene required for nutritional repression of the Bacillus subtilis dipeptide permease operon. Mol. Microbiol. 15:689-702. [DOI] [PubMed] [Google Scholar]
  • 37.Smith, M. C., L. Czaplewski, A. K. North, S. Baumberg, and P. G. Stockley. 1989. Nucleotide sequences required for regulation of arginine biosynthesis promoters are conserved between Bacillus subtilis and Escherichia coli. Mol. Microbiol. 3:23-28. [DOI] [PubMed] [Google Scholar]
  • 38.Stöver, A. G., and A. Driks. 1999. Regulation of synthesis of the Bacillus subtilis transition-phase, spore-associated antibacterial protein TasA. J. Bacteriol. 181:5476-5481. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Wray, L. V., Jr., A. E. Ferson, and S. H. Fisher. 1997. Expression of the Bacillus subtilis ureABC operon is controlled by multiple regulatory factors including CodY, GlnR, TnrA, and Spo0H. J. Bacteriol. 179:5494-5501. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Yeggy, J. P., and D. P. Stahly. 1980. Sporulation and regulation of homoserine dehydrogenase in Bacillus subtilis. Can. J. Microbiol. 26:1386-1391. [DOI] [PubMed] [Google Scholar]
  • 41.Yu, J., L. Hederstedt, and P. J. Piggot. 1995. The cytochrome bc complex (menaquinone:cytochrome c reductase) in Bacillus subtilis has a nontraditional subunit organization. J. Bacteriol. 177:6751-6760. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Journal of Bacteriology are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES