Skip to main content
Genome Biology logoLink to Genome Biology
. 2007 Dec 13;8(12):R265. doi: 10.1186/gb-2007-8-12-r265

Quantification of global transcription patterns in prokaryotes using spotted microarrays

Ben Sidders 1, Mike Withers 1, Sharon L Kendall 1, Joanna Bacon 3, Simon J Waddell 4, Jason Hinds 4, Paul Golby 5, Farahnaz Movahedzadeh 1,6, Robert A Cox 7, Rosangela Frita 1, Annemieke MC ten Bokum 8, Lorenz Wernisch 2, Neil G Stoker 1,
PMCID: PMC2246267  PMID: 18078514

Short abstract

An analysis is described, applicable to any spotted microarray dataset that is produced using genomic DNA as a reference for quantifying prokaryotic levels of mRNA on a genome-wide scale.

Abstract

We describe an analysis, applicable to any spotted microarray dataset produced using genomic DNA as a reference, that quantifies prokaryotic levels of mRNA on a genome-wide scale. Applying this to Mycobacterium tuberculosis, we validate the technique, show a correlation between level of expression and biological importance, define the complement of invariant genes and analyze absolute levels of expression by functional class to develop ways of understanding an organism's biology without comparison to another growth condition.

Background

The biological landscape has been transformed by the sequencing of genomes, and more recently by global gene expression analyses using microarrays [1,2]. Microarrays contain DNA probes representing all coding sequences in a genome, which are either synthesized in situ or are spotted onto a modified glass surface [3]. Comparison of mRNA from two conditions by competitive hybridization to these probes is used to identify differentially expressed genes [1]. In the case of spotted microarrays, these are performed either with labeled cDNA prepared from separate mRNA preparations co-hybridized to the same array, or as is increasingly the case, by employing genomic DNA (gDNA) as a standard reference [4]. In the latter case, each cDNA preparation is hybridized separately alongside a gDNA reference and differential expression is determined using a ratio of ratios. The use of gDNA corrects for most spatial and spot-dependent biases inherent with microarrays, and also allows direct comparison between multiple datasets [4]. These are sometimes called type 2 experiments, with RNA:RNA hybridizations being type 1 [5]. Traditionally, microarray experiments focus almost exclusively on changes in gene expression, and in the case of a type 1 experiment this is the only possible interpretation.

Focusing on changes in expression has helped to direct us toward genes that warrant further investigation; however, it has been shown in recent meta-analyses that up-regulated genes may bear little correlation to other measures of biological importance [6-8]. One reason for this lack of correlation is that, in a traditional microarray experiment, absolute levels of mRNA are not considered; thus, no difference is reported between a gene where expression increases from 20 to 100 copies and one where it increases from 20,000 to 100,000 copies, yet the biological inference may be very different. Furthermore, all genes whose level of expression does not alter significantly between conditions are completely ignored and we do not know if they are constitutively off or on (and if so, at what level). Differential expression analysis thus provides us with an incomplete view of the transcriptome, whereas the determination of global mRNA levels could, in part, address this.

Global mRNA abundance analysis is particularly applicable in prokaryotes, where, in contrast to the situation in eukaryotes, transcription and translation are tightly coupled [9,10]. In prokaryotes, therefore, absolute mRNA levels might be expected to accurately predict levels of protein. In support of this, it has been shown in both Escherichia coli and Mycobacterium smegmatis that the most readily detectable (and hence most abundant) proteins correspond to genes with high transcript levels [11,12]. Also, in experiments where transcriptomic and proteomic data were compared, for the majority of genes, changes at the transcriptional level were mirrored at the protein level [13,14]. Furthermore, a comprehensive study of mRNA and protein levels in a sulfur-reducing bacterium identified a modest global correlation between the two but found that the majority of the variation could be attributed to errors in the protein analytical techniques, indicating the actual correlation could be much stronger [15].

Surprisingly, the study of absolute levels of mRNA on a global scale has largely been ignored, despite attempts that have been made to extract meaningful quantitative information from microarrays. These include spiking various control samples of known concentration into the hybridization mixture [16,17], and using synthesized oligos complementary to every spot on an array at a known concentration as a normaliser [18]. Another recent report described the use of the Affymetrix gene chip platform to provide a quantitative view of gene expression levels in prokaryotes [19]. These approaches are often impractical or, especially with commercial systems, prohibitively expensive. Type 2 experiments performed on spotted arrays on the other hand, which use gDNA as a reference, are already routinely being performed, require a minimal cost increase and could allow us to study the relative abundance of each mRNA species [17] in parallel with traditional fold expression analyses.

Here we have focused on the determination of genome-wide mRNA levels of M. tuberculosis using type 2 microarrays for which we had a large number of datasets available. We have developed and validated the approach, characterized the genes whose level of expression is the highest in the transcriptome in vitro and those whose level of expression remains consistently high across a variety of environmental conditions. In addition, we have coupled genome-wide levels of mRNA abundance with a functional classification system in order to develop ways of understanding an organism's biology without comparison to another growth condition.

Results and discussion

Calculating relative mRNA abundance

Genome-wide transcriptional analyses have until now focused almost exclusively on differential expression. Methods that have been developed to quantify absolute mRNA abundances have largely been ignored or proved impractical. However, the use of gDNA as a reference channel in traditional microarray experiments is increasingly common [4], and as this is equivalent to an equimolar concentration of all transcripts we have investigated using this as a normaliser that would allow us to generate a measure of genome wide mRNA abundances.

Initially we calculated relative mRNA abundances for M. tuberculosis growing aerobically in chemostat cultures as we had access to the RNA used for hybridization to the arrays, allowing us to experimentally validate our analysis. RNA and microarray procedures were carried out as described in the Materials and methods and [20]. To obtain a measure of mRNA abundance, we first removed the local background fluorescence from each probe spot on the microarray. The fluorescence intensity from the RNA channel was then normalized to that of the gDNA channel, after which the technical and biological replicates were averaged.

We then found it necessary to correct for a probe length effect present in the data. In a traditional microarray (type 1) experiment, comparisons are made between two cDNA populations hybridized to one spot. Although in most cases it is necessary to control for factors such as the spatial dependent effects of hybridization, and normalizations such as loess are routinely implemented for this purpose [21], it is not necessary to control for individual spot variations as these would be negated when calculating fold expression ratios. In our case, where we are attempting to draw comparison between signals generated from different spots on the array, we are faced with additional factors that could introduce bias to our results - those that differ between spots on the array, for example, the probe GC content and length. Using the signal reported from a gDNA hybridization, we investigated these factors. We found that the length of the probes, which ranged in length from 60 to 1,000 bp, affected the hybridization whereby longer probes report higher signal intensities (Figure 1a). We suspect this may be because larger probes are able to bind more DNA and bind it more strongly than shorter probes. We corrected for this effect using a model of linear regression (Figure 1b; and see Materials and methods).

Figure 1.

Figure 1

Probe length normalization and quantified gene expression levels in M. tuberculosis. (a) We found that longer probes correlated with increased fluorescent intensities, which then biased the ppm values we obtained. (b) We are able to remove this bias using a model of linear regression. The three distinct groupings visible in the figure are an artifact of the probe lengths targeted during their synthesis by PCR. (c) The level of expression for each gene in the genome, as determined by our analysis from chemostat grown wild-type M. tuberculosis H37Rv, is shown ordered as they appear in the chromosome. (d) The log frequency distribution of mRNA abundances from (c). A clear skew to the right, containing a subset of very highly expressed genes, is typical of the distributions we have found.

Finally, as the sum of all fluorescence intensities is equal to the sum of all labeled mRNA, the measures of mRNA abundance were converted from unintuitive ratios to a proportional value; parts per million (ppm). Table 1 shows the mRNA abundance values for the 50 most highly expressed genes of M. tuberculosis. Figure 1c shows the mRNA levels, in ppm, for each gene in the M. tuberculosis chromosome and their log distribution is shown in Figure 1d. It is clear that the mRNA abundances, even once log transformed, are not normally distributed, which reflects the observations of others [22].

Table 1.

The 50 most highly expressed genes in vitro

Rv Name ppm Function [40]
1 Rv0009 ppiA 13,634 Protein translation and modification
2 Rv2527 Rv2527 10,020 Conserved hypotheticals
3 Rv3418c groES 8,651 Chaperones-heat shock*
4 Rv3615c Rv3615c 8,531 Conserved hypotheticals*
5 Rv0440 groEL2 8,300 Chaperones-heat shock*
6 Rv3258c Rv3258c 5,430 Unknown*
7 Rv3616c Rv3616c 5,370 Conserved hypotheticals*
8 Rv3477 PE31 5,138 PE subfamily
9 Rv2244 acpM 4,935 Synthesis of fatty and mycolic acids*
10 Rv0060 Rv0060 4,616 Unknown*
11 Rv3874 Rv3874 4,490 Conserved hypotheticals
12 Rv3875 esat6 4,115 SP, L, P and A
13 Rv3648c cspA 4,101 Adaptations and atypical conditions*
14 Rv3053c nrdH 3,934 2'-Deoxyribonucleotide metabolism
15 Rv3614c Rv3614c 3,748 Conserved hypotheticals*
16 Rv1078 pra 3,582 Conserved hypotheticals
17 Rv2780 ald 3,577 Amino acids and amines
18 Rv1388 mIHF 3,499 Nucleoproteins*
19 Rv0003 recF 3,371 DNA R, R R and R
20 Rv3786c Rv3786c 3,164 Unknown
21 Rv1641 infC 2,993 Protein translation and modification*
22 Rv3269 Rv3269 2,778 Chaperones-heat shock
23 Rv1398c Rv1398c 2,777 Conserved hypotheticals
24 Rv3407 Rv3407 2,732 Conserved hypotheticals
25 Rv2245 kasA 2,693 Synthesis of fatty and mycolic acids*
26 Rv0685 tuf 2,688 Protein translation and modification*
27 Rv2145c wag31 2,652 SP, L, P and A*
28 Rv1306 atpF 2,620 ATP-proton motive force*
29 Rv0005 gyrB 2,466 DNA R, R R and R*
30 Rv1305 atpE 2,455 ATP-proton motive force*
31 Rv0016c pbpA 2,423 Murein sacculus and peptidoglycan
32 Rv1622c cydB 2,418 Electron transport*
33 Rv3219 whiB1 2,373 Repressors-activators
34 Rv3583c Rv3583c 2,345 Repressors-activators
35 Rv1980c mpt64 2,254 SP, L, P and A
36 Rv1072 Rv1072 2,205 Other membrane proteins
37 Rv2457c clpX 2,200 Proteins, peptides and glycopeptides*
38 Rv1958c Rv1958c 2,195 Unknown
39 Rv3763 lpqH 2,180 Lipoproteins (lppA-lpr0)
40 Rv1872c lldD2 2,149 Aerobic respiration
41 Rv3461c rpmJ 2,133 Ribosomal protein synthesis*
42 Rv1361c PPE19 2,127 PPE family
43 Rv0097 Rv0097 2,123 Conserved hypotheticals
44 Rv1971 Rv1971 2,034 Virulence
45 Rv3051c nrdE 1,984 2'-Deoxyribonucleotide metabolism*
46 Rv2346c Rv2346c 1,937 Conserved hypotheticals
47 Rv3679 Rv3679 1,931 Anions*
48 Rv1298 rpmE 1,883 Ribosomal protein synthesis*
49 Rv0108c Rv0108c 1,837 Unknown
50 Rv2193 ctaE 1,793 Aerobic respiration*

*Essential genes (TraSH) [26,27]. Surface polysaccharides, lipopolysaccharides, proteins and antigens. DNA replication, repair, recombination and restriction-modification.

Validation of mRNA abundance data

In order to validate our estimates of abundance, we performed quantitative reverse transcriptase PCR (RTq-PCR) on a large selection of genes that we had predicted to span the mRNA abundance spectrum (n = 24). The RTq-PCR was carried out on a sample of the same RNA used for the microarray hybridizations. The measures of mRNA abundance as predicted from the microarray analysis show a good correlation (Spearman's rank = 0.86, p < 0.0001) with the absolute copy number as determined by RTq-PCR data (Figure 2).

Figure 2.

Figure 2

Microarray analysis validation. There is a strong correlation (0.86, Spearman's rank, p < 0.0001) between mRNA levels as predicted by our microarray analysis and mRNA copy number as determined by RTq-PCR.

Further validation of the method was provided by its high reproducibility when applied to data sets from independent laboratories using the same microarray designs. Correlations were determined for a variety of mRNA abundance data from both M. tuberculosis and the highly similar Mycobacterium bovis [23]. Chemostat-grown M. tuberculosis showed a correlation of 0.8 (p < 0.0001) with the homologous genes in chemostat-grown M. bovis. The same was true of batch-cultured M. tuberculosis and M. bovis grown in different institutions. However, there was a lower correlation of 0.5 (p < 0.0001) between chemostat and batch cultured M. tuberculosis from different laboratories, suggesting that the method of culture significantly affects the transcriptome.

mRNA abundance, protein abundance and gene importance

To explore the possibility that global measures of mRNA abundance are an important indicator of prokaryotic biology, we compared our mRNA abundance data with proteome and gene essentiality data. Demonstrating a correlation between mRNA and protein levels is difficult without the availability of genome-wide measures of protein abundance. We looked instead at the list of M. tuberculosis proteins identified to date from two-dimensional PAGE analysis and stored in an online database [24]. As the mRNA abundances are not normally distributed, we determined the frequency of identified proteins in each quartile of the abundance distribution. Of the 283 unique proteins identified in M. tuberculosis cell lysates and supernatants, the majority (187 proteins, 66%) are expressed in the most abundant quartile (Figure 3a). Others have suggested that proteomic experiments have an intrinsic bias toward abundant proteins [12,25] and this would support our hypothesis that the most abundant transcripts produce high levels of protein in bacteria. In addition, our analysis makes no allowance for differential rates of translation initiation or mRNA/protein degradation, so this finding reflects how tightly coupled the systems of transcription and translation are in prokaryotes [9,10].

Figure 3.

Figure 3

Correlations between mRNA and biological importance. (a) Proteins identified by two-dimensional PAGE/MS [24] correlates with the most highly expressed genes (Chi-squared test for trend in proportions = 251.9, df = 1, p value < 0.0001). (b) Similarly, there is a significant relationship between expression level and essentiality as determined by TraSH [7,26,27] (Chi-squared test for trend in proportions = 161.2, df = 1, p value < 0.0001).

As there is little correlation between reports of biological importance and gene induction [6-8] we instead looked at the correlation with mRNA abundance. We compared the genome-wide values of mRNA abundance with the genes identified as being essential in M. tuberculosis by genome-wide transposon mutant library (TraSH) screens [7,26,27]. For our RTq-PCR validated data from M. tuberculosis growing under aerobic chemostat conditions we found that there is a significant relationship between expression level and essentiality on a global scale (Chi-squared test for trend in proportions = 161.2, df = 1, p value < 0.0001; Figure 3b). This is in contrast to the lack of correlation with fold-induction upon infection [6,7] and illustrates the potential importance of determining mRNA abundances on a global scale. The correlation may reflect our previous finding that the more highly expressed a gene, the more protein is produced. Although there are obvious examples where proteins with essential functions, such as the cell division apparatus or many enzymes, need not be (or indeed cannot be) highly expressed [28], prokaryotic cells would waste considerable energy synthesizing large amounts of proteins that do not have essential functions.

The most highly expressed genes of M. tuberculosis

The genome-wide distribution of mRNA levels from M. tuberculosis cultured aerobically in the chemostat is typical of the distributions we have found (Figure 1d). Despite being log transformed, the distribution shows a skew to the right, suggesting the presence of a highly expressed subpopulation. As we and others have shown, the coupling of transcription and translation in prokaryotes means that there is an enormous material investment in transcribing a gene at a high level. We therefore analyzed the most abundant mRNAs in some detail, focusing on the 95th percentile of genes, that is, the 5% most abundant transcripts, n = 198 (Additional data file 1). Of these 198 genes, 89 (45%) were reported as essential in the TraSH experiments, which is significantly higher than the 23% of all genes that are essential (Χ2 p < 0.001). Of the 89 essential genes, 76 (38%) were essential in vitro, 11 (5%) in vivo and 2 (1%) are essential for survival in macrophages [7,26,27].

The most highly expressed gene in vitro was ppiA (Rv0009), a probable peptidyl-prolyl cis-trans isomerase involved in protein folding, which makes up 13,634 ppm (which is equivalent to 1.3%) of the total mRNA population. As would be expected, many of the very abundant transcripts belong to the protein synthesis machinery, including thirty ribosomal proteins, six translation initiation factors and various components of RNA polymerase.

Several of the very abundant genes have previously been characterized as highly expressed and extensively documented as important virulence determinants of M. tuberculosis. In particular, some members of the esx gene family have been shown to be critical in infection, although dispensable in vitro. The paradigms for this family are esat6 (Rv3875) and cfp10 (Rv3874), whose products form a secreted complex that interacts with host cells [29]. Furthermore, they are potent immunogens with potential roles as both subunit vaccines and diagnostic agents [30,31]. The esx family in M. tuberculosis contains 23 esat6-like genes, with 11 esat/cfp gene pairs [32]. Including esat6 and cfp10, we identified 12 members (52%) of this family as being amongst the most highly transcribed of all genes. One such pair of genes, esxV (Rv3619c) and esxW (Rv3620c), are adjacent to, but not transcribed with, the SNM (secretion in mycobacteria) operon containing Rv3616c (espA), Rv3615c (snm9) and Rv3614c (snm10), which we also find are very highly expressed. Two groups have recently shown that the SNM system functions to export both ESAT-6 and CFP-10 [33,34].

We have also observed five highly expressed transcripts belonging to the PE/PPE family; a set of approximately 100 genes encoding proteins with proline and glutamate rich motifs that are found exclusively within the mycobacteria [35]. Some members of this family are located adjacent to esx genes, suggesting a functional association, and it is now known that a PE/PPE pair form a stable dimer, reminiscent of ESAT-6/CFP-10 themselves [36]. Of the five PE/PPE genes in our very abundant transcript list, two are located adjacent to highly expressed esx family gene pairs: PPE18 (Rv1196) with esxKL (Rv1197/Rv1198), and PE19 (Rv1791) with esxMN (Rv1792/Rv1793). The significance of linked high expression levels between esx and PE/PPE genes suggests a co-functionality critical to M. tuberculosis biology.

Genes of unknown function

Thirty-three percent of the coding sequences in M. tuberculosis were classified as having no known function in a re-annotation of the genome [37]. A similar proportion (60 of 198, 30%) of the transcripts we have classified as very abundant in M. tuberculosis are annotated as unknown, hypothetical or conserved hypothetical proteins. The organism consumes considerable energy in their expression so it is likely they have important functions, and indeed 12 of these unknown genes are essential. Using both BLASTP and functional predictions generated with the hidden Markov model profile tool SHARKhunt [38], we were unable to ascribe any further functions to these genes with the exception of Rv0097, which has close homology to a taurine dioxygenase of the Streptomyces, Ralstonia and Chromobacterium species. It has been reported [39] that the Rv0097 homologue in M. bovis is located in an operon essential for the synthesis of the virulence associated lipids phthiocerol dimycocerosate esters (PDIMs) and could function as an α-ketoglutarate-dependent dioxygenase, the super family to which taurine dioxygenases belong.

The invariome

Much effort goes into looking for genes whose expression is modulated in different environments. Although it is likely that most, if not all, genes are regulated to some extent, using the analysis described here it is possible to search for genes whose expression does not change significantly, which we term 'invariant'. These may represent genes whose functions are so important that they cannot be switched off. To identify these invariant genes, we compared the mRNA abundance in a total of six data sets including various growth conditions, such as chemostat, batch culture, low oxygen and growth in macrophages. We focused on genes that were within the 85th percentile and found that 133 genes were consistently highly expressed across all of the conditions tested (Table 2).

Table 2.

The 133 genes of the 'abundant invariome'

Rv Name Avg ppm Stdev Essential
1 Rv3874 lhp 5,414 3,950 -
2 Rv3418c groES 5,189 2,593 In vitro
3 Rv0440 groEL2 4,438 2,385 In vitro
4 Rv3615c Rv3615c 3,887 2,539 In vivo
5 Rv0009 ppiA 3,460 4,587 -
6 Rv3616c Rv3616c 2,619 1,457 In vivo
7 Rv3477 PE31 2,537 1,553 -
8 Rv2244 acpM 2,475 1,304 In vitro
9 Rv3875 esat6 2,472 1,229 -
10 Rv1398c Rv1398c 2,449 1,311 -
11 Rv3648c cspA 2,372 1,149 In vitro
12 Rv2245 kasA 2,236 481 In vitro
13 Rv3614c Rv3614c 2,232 847 In vivo
14 Rv1307 atpH 2,151 1,195 In vitro
15 Rv0667 rpoB 2,105 563 In vitro
16 Rv1388 mihF 2,103 1,013 In vitro
17 Rv0685 tuf 2,100 792 In vitro
18 Rv3583c Rv3583c 2,096 2,027 -
19 Rv3053c nrdH 1,930 1,339 -
20 Rv1133c metE 1,915 323 In vitro
21 Rv1072 Rv1072 1,897 560 -
22 Rv1872c lldD2 1,817 574 -
23 Rv3461c rpmJ 1,814 795 In vitro
24 Rv2457c clpX 1,790 343 In vitro
25 Rv0700 rpsJ 1,693 1,148 In vitro
26 Rv1078 pra 1,643 971 -
27 Rv1298 rpmE 1,529 543 In vitro
28 Rv2840c Rv2840c 1,495 566 -
29 Rv1630 rpsA 1,491 439 In vitro
30 Rv0046c ino1 1,488 620 -
31 Rv1886c fbpB 1,464 1,168 -
32 Rv2196 qcrB 1,455 472 In vitro
33 Rv3443c rplM 1,447 300 In vitro
34 Rv0701 rplC 1,421 880 In vitro
35 Rv0682 rpsL 1,419 576 In vitro
36 Rv3219 whiB1 1,384 795 -
37 Rv0702 rplD 1,364 619 In vitro
38 Rv0289 Rv0289 1,351 845 In vitro
39 Rv2200c ctaC 1,332 1,406 In vitro
40 Rv1980c mpt64 1,316 629 -
41 Rv1306 atpF 1,246 695 In vitro
42 Rv2193 ctaE 1,217 334 In vitro
43 Rv1310 atpD 1,184 412 In vitro
44 Rv1174c Rv1174c 1,165 424 -
45 Rv1308 atpA 1,148 349 In vitro
46 Rv3051c nrdE 1,123 578 In vitro
47 Rv1305 atpE 1,086 696 In vitro
48 Rv0683 rpsG 1,080 541 In vitro
49 Rv1297 rho 1,074 281 In vitro
50 Rv2461c clpP1 1,028 346 -
51 Rv0655 mkl 1,024 385 In vivo
52 Rv3052c nrdI 1,018 551 -
53 Rv3801c fadD32 1,015 209 In vitro
54 Rv0005 gyrB 1,011 684 In vitro
55 Rv0704 rplB 1,011 528 In vitro
56 Rv3412 Rv3412 1,002 141 -
57 Rv0250c Rv0250c 995 439 -
58 Rv2460c clpP2 991 393 In vitro
59 Rv2204c Rv2204c 963 277 -
60 Rv3478 PPE60 957 360 -
61 Rv0703 rplW 956 556 In vitro
62 Rv2094c tatA 949 241 -
63 Rv1303 Rv1303 939 637 In vitro
64 Rv3456c rplQ 937 297 -
65 Rv0719 rplF 925 406 In vitro
66 Rv0684 fusA1 923 367 In vitro
67 Rv2347c Rv2347c 920 447 -
68 Rv0715 rplX 906 342 In vitro
69 Rv1197 Rv1197 906 440 -
70 Rv1479 moxR1 898 162 In vitro
71 Rv0718 rpsH 886 243 In vitro
72 Rv3460c rpsM 882 373 -
73 Rv2194 qcrC 870 194 In vitro
74 Rv2195 qcrA 868 189 In vitro
75 Rv0860 fadB 863 420 -
76 Rv1309 atpG 861 401 In vitro
77 Rv0243 fadA2 859 231 -
78 Rv3248c sahH 850 243 In vitro
79 Rv0020c TB39.8 850 224 -
80 Rv3584 lpqE 845 246 -
81 Rv1793 Rv1793 836 309 -
82 Rv3620c Rv3620c 827 375 -
83 Rv1410c Rv1410c 815 188 In vivo
84 Rv3459c rpsK 809 206 In vitro
85 Rv0483 lprQ 805 341 -
86 Rv3043c ctaD 803 221 In vitro
87 Rv3029c fixA 801 232 In vitro
88 Rv2868c gcpE 799 384 -
89 Rv1304 atpB 796 203 In vivo
90 Rv1642 rpmI 784 388 -
91 Rv1794 Rv1794 781 254 -
92 Rv0288 cfp7 781 286 -
93 Rv3810 pirG 778 131 In vivo
94 Rv1543 Rv1543 770 222 -
95 Rv3680 Rv3680 765 294 -
96 Rv3457c rpoA 760 293 In vitro
97 Rv3045 adhC 756 272 -
98 Rv1792 Rv1792 753 326 -
99 Rv2969c Rv2969c 738 141 In vitro
100 Rv1177 fdxC 736 303 In vitro
101 Rv3867 Rv3867 735 200 -
102 Rv1038c Rv1038c 724 391 -
103 Rv2890c rpsB 715 120 In vitro
104 Rv3224 Rv3224 709 303 -
105 Rv3458c rpsD 707 208 In vitro
106 Rv2785c rpsO 690 316 -
107 Rv2986c hupB 687 255 In vitro
108 Rv0174 mce1F 683 211 -
109 Rv3211 rhlE 676 251 -
110 Rv1436 gap 673 234 In vitro
111 Rv0351 grpE 672 282 In vitro
112 Rv2764c thyA 667 239 -
113 Rv1311 atpC 660 160 In vitro
114 Rv0432 sodC 657 177 -
115 Rv1791 PE19 655 232 -
116 Rv0932c pstS2 652 249 -
117 Rv2971 Rv2971 645 188 In vitro
118 Rv1300 hemK 644 220 In vitro
119 Rv2703 sigA 643 114 In vitro
120 Rv2203 Rv2203 633 196 -
121 Rv0423c thiC 614 149 In vitro
122 Rv2587c secD 602 267 -
123 Rv1887 Rv1887 601 148 -
124 Rv0313 Rv0313 588 136 -
125 Rv0502 Rv0502 559 147 -
126 Rv3841 bfrB 542 107 -
127 Rv2115c Rv2115c 529 106 -
128 Rv3587c Rv3587c 526 68 -
129 Rv2110c prcB 516 151 In vitro
130 Rv1987 Rv1987 495 136 -
131 Rv2454c Rv2454c 489 55 -
132 Rv0125 pepA 483 80 -
133 Rv1324 Rv1324 474 152 -

Using our measure of absolute mRNA levels, we have been able to identify the genes whose level of expression remains consistently high across a variety of growth conditions. These genes remain amongst the top 15% most highly expressed genes across all of the conditions tested. We have termed the genes whose level of expression does not vary greatly as invariant, and, therefore, the subset of genes included in this table is dubbed the 'abundant invariome'.

Notable members of this abundant invariome include several esx-related genes (esat6, cfp10 and the SNM secretion system), the paired PE31 and PPE60, groEL2 (the 65 kDa antigen) and the ATP synthase operon (atpA-atpH). Compared to a global 23% of genes that are essential, 53% (70 of 133, Χ2 p < 0.001) of the genes in the abundant invariome are essential for either in vitro growth or survival in mice or macrophages [7,26,27]. Twenty-two of our abundant invariome genes have no known function, two of which (Rv0289 and Rv1303) are essential. These and other members of the abundant invariome are primary candidates for future functional studies that may elucidate key mycobacterial biology.

As well as an obvious hypothesis generation role, such invariant gene analyses might have uses, for example, in identifying strong antigens, constitutive promoters and stable housekeeping genes expressed in all environments. Furthermore, as significant proportions of the abundant invariome are essential, this analysis also has the potential to identify drug targets or candidate virulence factors in prokaryotes even if no other biological information is known.

Highly transcribed functional categories

The advent of systems biology is encouraging the development of techniques that reveal more holistic information about biological systems. We have therefore combined our measures of M. tuberculosis mRNA abundance with a non-overlapping classification system based on known or predicted functions [40,41]. Not only does this accord with the aims of systems biology, it would also remove the need to routinely compare expression profiles to that of artificial laboratory conditions and could, therefore, be more biologically meaningful.

Using this analysis to study the transcriptome of M. tuberculosis in a disease relevant [42] low oxygen state (chemostat grown at a dissolved oxygen tension (DOT) of 0.2% [20]) reveals that, at the broadest scope of the classification system, 29% of the mRNA in the transcriptome codes for proteins involved in small molecule metabolism, 19% for macromolecule metabolism, 7% for 'other' functions (virulence, and so on) and 7% for cellular processes. In addition, 38% of the in vitro low oxygen transcriptome codes for proteins of unknown function, illustrating how little of mycobacterial biology has been characterized to date.

To determine which functional classes, at all levels of the classification system, were significantly over- or under-represented in the transcriptome, we chose, for comparison, three different robust and nonparametric approaches: robust linear modeling, bootstrap-t using the Q statistic of Davison and Hinkley [43], and a bootstrap-t using trimmed means and winsorized variances [44]. We removed all classes with fewer than four data points to be able to obtain variance estimates after trimming. As an example of how this might be used, we again focused on the low oxygen transcriptome. In this example, many of the functional classes that we have shown to be significantly over-represented within the transcriptome by all three tests reflect the growth rate in the chemostat (maintained at a 24 hour doubling time [20]), including the protein translation machinery, the chaperones, the RNA and DNA synthesis mechanisms, and the ribosomal proteins (Additional data file 2). Also abundant are classes involved with energy metabolism, including ATP synthesis and the TCA cycle, as well as macromolecule synthesis, including the fatty and mycolic acid anabolic pathways.

Using either of the bootstrap-t methods appears to be too stringent to reveal classes that we would immediately recognize as reflecting the adaptation to low oxygen. However, the classes deemed significant by the robust linear modeling method include the glyoxylate shunt enzymes, which are essential in vivo for the anaplerosis of acetyl-CoA when growing on fatty acids [45,46], and the oxidative electron transport system known to operate under reduced oxygen tensions in mycobacteria [47].

Our functional analyses are only preliminary; we are limited by the lack of a comprehensive bacterial gene ontology. However, we suggest that this is a biologically relevant approach that could be expanded and used to identify the key cellular and metabolic processes required by an organism in a particular growth condition. It will link well with other systems biology analyses to produce useful insights into bacterial physiological states and, for example, could be used to determine the processes, rather than the components, required for infection and latency in M. tuberculosis.

Conclusion

We have developed a method of microarray analysis that quantifies levels of mRNA on a genome-wide scale. Our method of analysis can be applied to any spotted microarray data set produced using gDNA as a reference channel. Applying this analysis to the prokaryote M. tuberculosis, we have identified the most highly expressed genes and note correlations with gene essentiality as well as with a basic measure of protein abundance. We have also been able to define the subset of genes that are invariantly highly expressed and find that more than half are essential for growth in vitro or survival in vivo. In addition, we are also able to produce a functionally organized holistic view of the transcriptome. Alongside traditional changes in expression, mRNA abundance analysis can, therefore, greatly enhance the utility of microarray data and has numerous additional uses that will aid genetic research into prokaryotic organisms.

Materials and methods

Microarrays

Six microarray datasets have been used in this study (Table 3). The microarrays used for hybridization were the BμG@S TB version 1 arrays (Array Express accession: A-BUGS-1) [48], for data sets 1 to 3, containing 3,924 spotted PCR products and TB version 2 arrays (A-BUGS-23) [48], for data sets 4 to 6, containing 4,410 spotted PCR products, including additional open reading frames from M. bovis strain AF2122/97. Fully annotated microarray data have been deposited in BμG@Sbase (accession number E-BUGS-60) [49] and also ArrayExpress (accession number E-BUGS-60).

Table 3.

Microarray datasets used in this study

Description Origin Reference Data storage
1 Wild-type Mtb H37Rv aerobic chemostat CAMR, UK [20] BμG@Sbase: E-BUGS-60
2 Wild-type Mtb H37Rv low oxygen chemostat - 0.2% DOT CAMR, UK [20] BμG@Sbase: E-BUGS-60
3 Wild-type Mtb H37Rv aerobic rolling batch culture RVC, UK Unpublished BμG@Sbase: E-BUGS-60
4 Wild-type Mbovis AF2122/97 aerobic chemostat VLA, UK Unpublished BμG@Sbase: E-BUGS-60
5 Wild-type Mbovis AF2122/97 aerobic rolling batch culture VLA, UK Unpublished BμG@Sbase: E-BUGS-60
6 Wild-type Mtb H37Rv harvested from macrophages SGUL, UK Unpublished BμG@Sbase: E-BUGS-60

Bacterial culture

Batch cultures of M. tuberculosis and M. bovis were grown in 100 ml Middlebrook 7H9 (Becton, Dickinson and Co. Franklin Lakes, NJ, USA) supplemented with 10% Middlebrook OADC (Becton, Dickinson and Co.) in 1 liter capacity flasks that were continuously rolled at 37°C. Chemostat cultures, RNA extraction and microarray hybridizations were performed as detailed in Bacon et al. 2004 [20]. Briefly, 500 ml cultures were grown to steady state conditions in 1 liter chemostat fermentation vessels maintained at 37°C, with a pH of 6.9 and a generation time of 24 hours. Aerobic cultures were kept at a DOT of 10% whilst low oxygen cultures were maintained at 0.2% DOT.

RNA and genomic DNA extraction

Aliquots of the bacterial cultures (10 ml) were sampled directly into four volumes of a guanadinium thiocyanate based stop solution. After centrifugation and resuspension in either Trizol (Invitrogen, Carlsbad, California, USA) or additional stop solution, cells were lysed using a Ribolyser (Hybaid, Teddington, England) or Precellys 24 (Bertin Technologies, Montigny-le-Bretonneux, France) and RNA was extracted using a chloroform precipitation followed by purification with the RNeasy Mini Kit (Qiagen, Hilden, Germany). RNA was then treated with deoxyribonuclease (DNase 1 amplification grade, Life Technologies Inc/Invitrogen, Carlsbad, California, USA). Genomic DNA was isolated from pellets of stationary phase mycobacterial cultures following previously described procedures [50].

RNA/DNA labeling and hybridization

For the majority of the datasets RNA was extracted from four independent biological replicates and each was labeled in triplicate using 8 μg total RNA as a template for reverse transcriptase (Superscript II RNAse H, Life Technologies Inc.) in the presence of random primers (Invitrogen, Carlsbad, California, USA) and Cy5 labeled dCTP. Genomic DNA (1 μg) was used as a template for DNA polymerase (Klenow, Invitrogen, Carlsbad, California, USA) in the presence of random primers and Cy3 labeled dCTP.

Purified Cy3/Cy5 labeled DNA were combined and added to the array underneath a cover slip before being sealed in a hybridization chamber and submerged in a water bath at 65°C for 16-20 hours. Scanning was performed with a dual laser scanner (Affymetrix 428, MWG Biotech, Ebersberg, Germany) at a gain below saturation of the most intense spots. Both spot and local background levels were quantified from the resulting images using ImaGene 4.0 (MWG Biotech).

mRNA abundance calculation

The Perl computing language and the R statistical environment [51] were used to perform all data and statistical analysis. The YASMA [52] microarray analysis package for R was used to input and structure the raw data.

Initially, all control spots on the array were removed from the dataset, including all representing ribosomal RNA. The local background noise, as determined by the image quantification software, was subtracted from each spot. No data values were excluded from this study as we reasoned weak signals (after background subtraction) were reflective of low abundance transcripts.

For each spot i on the array the fluorescent intensity from the cDNA (RNA) channel was normalized by simple division to the fluorescent intensity of the gDNA channel:

Normalized intensity (Ri) = cDNAi/gDNAi

The correlation between hybridization replicates within each dataset was confirmed to ensure there were no extreme outliers. Technical and biological replicates were then averaged to provide a single normalized intensity value for each gene on the array.

To account for the observed probe length bias (see Results and discussion), signal intensity was normalized to probe length using a model of linear regression of log intensity on probe length (Figure 1a,b):

Probe normalized intensity (loge Rni) = loge Ri - (intercept + slope × probe lengthi)

The corrected Rni values were converted back to a raw scale and for ease of understanding are depicted as a proportional value, expressed in ppm, based on the assumption that the sum of all intensity values represents the sum of the transcript (mRNA) population within the sample:

ppm = (Rni/∑i-ith Rn) × 106

RTq-PCR

To assess if the measures of transcript abundance from the array analysis truly reflect that of the RNA sample we carried out RTq-PCR on 24 genes predicted to span the spectrum of mRNA abundances as determined by the mRNA abundance analysis. The RNA samples used were those extracted and used for the microarray hybridizations in data set 1 [20]. Total RNA (400 ng) was reverse transcribed using Superscript Reverse Transcriptase III (Invitrogen) according to the manufacturer's instructions. cDNA (5 ng) was subsequently used for RTq-PCR using the DyNAmo SYBR green qPCR kit (Finnzymes, Espoo, Finland), according to the manufacturer's instructions, in a DNA Engine Opticon 2 thermal cycler (MJ Research, Waltham, MA, USA). For each reaction a set of gDNA standards of known copy number were used to produce a standard curve from which a copy number could accurately be determined. The data were analyzed using Opticon Monitor v2.0.

Functional category analysis

The genes of M. tuberculosis were grouped based on a Riley-like classification system obtained from the Sanger Institute [40]. The classification system is non-overlapping and hierarchical, and thus has six highest level functional categories: small-molecule metabolism, macromolecule metabolism, cell processes, other, conserved hypothetical and unknowns, each of which splits into more specific subcategories. In total, 3,925 genes are classified in this system.

We were looking to compare the level of expression in each functional class with that in other classes to discover which classes might be over- or under-represented within the transcriptome. We used ANOVA based approaches to detect functional classes that show significant over- or under-representation compared to the rest. We estimated location parameters for the log ppm values for each class and assessed the statistical significance of the contrast of a location parameter for a particular class and the average of location parameters for all the other classes. As seen in Figure 1d, the mRNA abundances are not normally distributed even after log-transformation, so we could not assume that the distribution of abundances within each class would be. Moreover, the variances changed considerably between functional classes. We therefore chose three different robust and nonparametric approaches to estimate the location parameters and to establish significance of contrasts: robust linear modeling; a bootstrap-t using the Q statistic of Davison and Hinkley [43]; and a bootstrap-t using trimmed means and winsorized variances [44]. We removed all classes with fewer than four data points to be able to obtain variance estimates after trimming.

For the robust linear model we used the rlm function from the R package MASS [53], with Huber's Psi function and default settings. In the bootstrap-t with Q values as pivot we trimmed points with their robustly estimated residues in the top 20% and bottom 20% quantiles before performing the bootstrap analysis to protect the estimations against outliers. For the second bootstrap-t we used trimmed means with 20% of the lower and 20% of the upper quantiles removed, corresponding to 20% winsorized variances for the pivot. Resampling was done within each functional class, since we take them as fixed effects. We collected 10,000 bootstrap samples to allow for multiple testing correction. The method of Benjamini and Yekutieli [54] was used to calculate a false discovery rate allowing for dependencies between functional classes.

The results from all tests are provided as supporting information (Additional data file 2). Classes were considered significantly different from others if the adjusted p value was less than 0.05.

Abbreviations

DOT, dissolved oxygen tension; gDNA, genomic DNA; ppm, parts per million; RTq-PCR, quantitative reverse transcriptase PCR.

Authors' contributions

BS, LW and NS designed the research; BS, MW and LW performed the research; SLK, RAC, RF, AMCtB, SW, JH, PG, FM, and JB contributed reagents or data to this study; BS, LW and NS wrote the paper.

Additional data files

The following additional data are available with the online version of this paper. Additional data file 1 is a table listing the 198 genes of the 95th percentile; the very abundant transcripts in M. tuberculosis. Additional data file 2 is a table that includes the quantified level of each functional category and details of those deemed significantly more or less abundant in the low oxygen transcriptome, including data from the three approaches to assess significance.

Supplementary Material

Additional data file 1

The 198 genes of the 95th percentile; the very abundant transcripts in M. tuberculosis.

Click here for file (60KB, xls)
Additional data file 2

The quantified level of each functional category and details of those deemed significantly more or less abundant in the low oxygen transcriptome, including data from the three approaches to assess significance.

Click here for file (39.5KB, xls)

Acknowledgments

Acknowledgements

BS is funded by a Royal Veterinary College scholarship. MW is funded by the EU MM-TB STREP grant no. 012187. SLK and AtB are supported by Wellcome Trust grants 073237 and 081455. We acknowledge BμG@S (the Bacterial Microarray Group at St George's, University of London) and The Wellcome Trust for funding the multi-collaborative microbial pathogen microarray facility under its Functional Genomics Resources Initiative.

Contributor Information

Ben Sidders, Email: bsidders@rvc.ac.uk.

Mike Withers, Email: mwithers@rvc.ac.uk.

Sharon L Kendall, Email: skendall@rvc.ac.uk.

Joanna Bacon, Email: Joanna.Bacon@hpa.org.uk.

Simon J Waddell, Email: swaddell@sgul.ac.uk.

Jason Hinds, Email: j.hinds@sgul.ac.uk.

Paul Golby, Email: p.golby@vla.defra.gsi.gov.uk.

Farahnaz Movahedzadeh, Email: movahed@uic.edu.

Robert A Cox, Email: s.cox@moose.co.uk.

Rosangela Frita, Email: rfrita@rvc.ac.uk.

Annemieke MC ten Bokum, Email: Annemieke.tenBokum@lshtm.ac.uk.

Lorenz Wernisch, Email: l.wernisch@mail.cryst.bbk.ac.uk.

Neil G Stoker, Email: nstoker@rvc.ac.uk.

References

  1. Schena M, Shalon D, Davis RW, Brown PO. Quantitative monitoring of gene expression patterns with a complementary DNA microarray. Science. 1995;270:467–470. doi: 10.1126/science.270.5235.467. [DOI] [PubMed] [Google Scholar]
  2. Conway T, Schoolnik GK. Microarray expression profiling: capturing a genome-wide portrait of the transcriptome. Mol Microbiol. 2003;47:879–889. doi: 10.1046/j.1365-2958.2003.03338.x. [DOI] [PubMed] [Google Scholar]
  3. Heller MJ. DNA microarray technology: devices, systems, and applications. Annu Rev Biomed Eng. 2002;4:129–153. doi: 10.1146/annurev.bioeng.4.020702.153438. [DOI] [PubMed] [Google Scholar]
  4. Talaat AM, Howard ST, Hale W, 4th, Lyons R, Garner H, Johnston SA. Genomic DNA standards for gene expression profiling in Mycobacterium tuberculosis. Nucleic Acids Res. 2002;30:e104. doi: 10.1093/nar/gnf103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Sidders B, Stoker NG. Transcriptome analysis: towards a comprehensive understanding of global transcription activity. In: Hacker J, Dobrindt U, editor. Pathogenomics: Genome Analysis of Pathogenic Microbes. Weinham, Germany: Wiley-VCH; 2006. pp. 21–41. [Google Scholar]
  6. Kendall SL, Rison SC, Movahedzadeh F, Frita R, Stoker NG. What do microarrays really tell us about M. tuberculosis? Trends Microbiol. 2004;12:537–544. doi: 10.1016/j.tim.2004.10.005. [DOI] [PubMed] [Google Scholar]
  7. Rengarajan J, Bloom BR, Rubin EJ. Genome-wide requirements for Mycobacterium tuberculosis adaptation and survival in macrophages. Proc Natl Acad Sci USA. 2005;102:8327–8332. doi: 10.1073/pnas.0503272102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Miklos GL, Maleszka R. Microarray reality checks in the context of a complex disease. Nat Biotechnol. 2004;22:615–621. doi: 10.1038/nbt965. [DOI] [PubMed] [Google Scholar]
  9. Miller OL, Jr, Hamkalo BA, Thomas CA., Jr Visualization of bacterial genes in action. Science. 1970;169:392–395. doi: 10.1126/science.169.3943.392. [DOI] [PubMed] [Google Scholar]
  10. Cox RA. A scheme for the analysis of microarray measurement based on a quantitative theoretical framework for bacterial cell growth: application to studies of Mycobacterium tuberculosis. Microbiology. 2007;153:3337–3349. doi: 10.1099/mic.0.2007/005868-0. [DOI] [PubMed] [Google Scholar]
  11. Corbin RW, Paliy O, Yang F, Shabanowitz J, Platt M, Lyons CE, Jr, Root K, McAuliffe J, Jordan MI, Kustu S, et al. Toward a protein profile of Escherichia coli: comparison to its transcription profile. Proc Natl Acad Sci USA. 2003;100:9232–9237. doi: 10.1073/pnas.1533294100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Wang R, Prince JT, Marcotte EM. Mass spectrometry of the M. smegmatis proteome: Protein expression levels correlate with function, operons, and codon bias. Genome Res. 2005;15:1118–1126. doi: 10.1101/gr.3994105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Eymann C, Homuth G, Scharf C, Hecker M. Bacillus subtilis functional genomics: global characterization of the stringent response by proteome and transcriptome analysis. J Bacteriol. 2002;184:2500–2520. doi: 10.1128/JB.184.9.2500-2520.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Yoshida K-i, Kobayashi K, Miwa Y, Kang C-M, Matsunaga M, Yamaguchi H, Tojo S, Yamamoto M, Nishi R, Ogasawara N, et al. Combined transcriptome and proteome analysis as a powerful approach to study genes under glucose repression in Bacillus subtilis. Nucleic Acids Res. 2001;29:683–692. doi: 10.1093/nar/29.3.683. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Nie L, Wu G, Zhang W. Correlation between mRNA and protein abundance in Desulfovibrio vulgaris: a multiple regression to identify sources of variations. Biochem Biophys Res Commun. 2006;339:603–610. doi: 10.1016/j.bbrc.2005.11.055. [DOI] [PubMed] [Google Scholar]
  16. Caldwell R, Sapolsky R, Weyler W, Maile RR, Causey SC, Ferrari E. Correlation between Bacillus subtilis scoC phenotype and gene expression determined using microarrays for transcriptome analysis. J Bacteriol. 2001;183:7329–7340. doi: 10.1128/JB.183.24.7329-7340.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Bernstein JA, Khodursky AB, Lin PH, Lin-Chao S, Cohen SN. Global analysis of mRNA decay and abundance in Escherichia coli at single-gene resolution using two-color fluorescent DNA microarrays. Proc Natl Acad Sci USA. 2002;99:9697–9702. doi: 10.1073/pnas.112318199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Dudley AM, Aach J, Steffen MA, Church GM. Measuring absolute expression with microarrays with a calibrated reference sample and an extended signal intensity range. Proc Natl Acad Sci USA. 2002;99:7554–7559. doi: 10.1073/pnas.112683499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Fu L, Fu-Liu C. The gene expression data of Mycobacterium tuberculosis based on Affymetrix gene chips provide insight into regulatory and hypothetical genes. BMC Microbiology. 2007;7:37. doi: 10.1186/1471-2180-7-37. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Bacon J, James BW, Wernisch L, Williams A, Morley KA, Hatch GJ, Mangan JA, Hinds J, Stoker NG, Butcher PD, et al. The influence of reduced oxygen availability on pathogenicity and gene expression in Mycobacterium tuberculosis. Tuberculosis. 2004;84:205–217. doi: 10.1016/j.tube.2003.12.011. [DOI] [PubMed] [Google Scholar]
  21. Smyth GK, Speed T. Normalization of cDNA microarray data. Methods. 2003;31:265–273. doi: 10.1016/S1046-2023(03)00155-5. [DOI] [PubMed] [Google Scholar]
  22. Danchin A, Sekowska A. Expression profiling in reference bacteria: dreams and reality. Genome Biol. 2000;1:REVIEWS1024. doi: 10.1186/gb-2000-1-4-reviews1024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Garnier T, Eiglmeier K, Camus J-C, Medina N, Mansoor H, Pryor M, Duthoy S, Grondin S, Lacroix C, Monsempe C, et al. The complete genome sequence of Mycobacterium bovis. Proc Natl Acad Sci USA. 2003;100:7877–7882. doi: 10.1073/pnas.1130426100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Mollenkopf H-J, Jungblut PR, Raupach B, Mattow J, Lamer S, Zimny-Arndt U, Schaible UE, Kaufmann SHE. A dynamic two-dimensional polyacrylamide gel electrophoresis database: The mycobacterial proteome via Internet. Electrophoresis. 1999;20:2172–2180. doi: 10.1002/(SICI)1522-2683(19990801)20:11&#x0003c;2172::AID-ELPS2172&#x0003e;3.0.CO;2-M. [DOI] [PubMed] [Google Scholar]
  25. Liu H, Sadygov RG, Yates JR. A model for random sampling and estimation of relative protein abundance in shotgun proteomics. Anal Chem. 2004;76:4193–4201. doi: 10.1021/ac0498563. [DOI] [PubMed] [Google Scholar]
  26. Sassetti CM, Boyd DH, Rubin EJ. Genes required for mycobacterial growth defined by high density mutagenesis. Mol Microbiol. 2003;48:77–84. doi: 10.1046/j.1365-2958.2003.03425.x. [DOI] [PubMed] [Google Scholar]
  27. Sassetti CM, Rubin EJ. Genetic requirements for mycobacterial survival during infection. Proc Natl Acad Sci USA. 2003;100:12989–12994. doi: 10.1073/pnas.2134250100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Stoker N, Broome-Smith J, Edelman A, Spratt B. Organization and subcloning of the dacA-rodA-pbpA cluster of cell shape genes in Escherichia coli. J Bacteriol. 1983;155:847–853. doi: 10.1128/jb.155.2.847-853.1983. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Renshaw PS, Panagiotidou P, Whelan A, Gordon SV, Hewinson RG, Williamson RA, Carr MD. Conclusive evidence that the major T-cell antigens of the Mycobacterium tuberculosis complex ESAT-6 and CFP-10 form a tight, 1:1 complex and characterization of the structural properties of ESAT-6, CFP-10, and the ESAT-6-CFP-10 complex. Implications for pathogenesis and virulence. J Biol Chem. 2002;277:21598–21603. doi: 10.1074/jbc.M201625200. [DOI] [PubMed] [Google Scholar]
  30. Sorensen A, Nagai S, Houen G, Andersen P, Andersen A. Purification and characterization of a low-molecular-mass T-cell antigen secreted by Mycobacterium tuberculosis. Infect Immun. 1995;63:1710–1717. doi: 10.1128/iai.63.5.1710-1717.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Pym AS, Brodin P, Majlessi L, Brosch R, Demangel C, Williams A, Griffiths KE, Marchal G, Leclerc C, Cole ST. Recombinant BCG exporting ESAT-6 confers enhanced protection against tuberculosis. Nat Med. 2003;9:533–539. doi: 10.1038/nm859. [DOI] [PubMed] [Google Scholar]
  32. Brodin P, Rosenkrands I, Andersen P, Cole ST, Brosch R. ESAT-6 proteins: protective antigens and virulence factors? Trends Microbiol. 2004;12:500–508. doi: 10.1016/j.tim.2004.09.007. [DOI] [PubMed] [Google Scholar]
  33. Fortune SM, Jaeger A, Sarracino DA, Chase MR, Sassetti CM, Sherman DR, Bloom BR, Rubin EJ. Mutually dependent secretion of proteins required for mycobacterial virulence. Proc Natl Acad Sci USA. 2005;102:10676–10681. doi: 10.1073/pnas.0504922102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Macgurn JA, Raghavan S, Stanley SA, Cox JS. A non-RD1 gene cluster is required for Snm secretion in Mycobacterium tuberculosis. Mol Microbiol. 2005;57:1653–1663. doi: 10.1111/j.1365-2958.2005.04800.x. [DOI] [PubMed] [Google Scholar]
  35. Brennan MJ, Delogu G. The PE multigene family: a 'molecular mantra' for mycobacteria. Trends Microbiol. 2002;10:246–249. doi: 10.1016/S0966-842X(02)02335-1. [DOI] [PubMed] [Google Scholar]
  36. Strong M, Sawaya MR, Wang S, Phillips M, Cascio D, Eisenberg D. Toward the structural genomics of complexes: Crystal structure of a PE/PPE protein complex from Mycobacterium tuberculosis. Proc Natl Acad Sci USA. 2006;103:8060–8065. doi: 10.1073/pnas.0602606103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Camus J-C, Pryor MJ, Medigue C, Cole ST. Re-annotation of the genome sequence of Mycobacterium tuberculosis H37Rv. Microbiology. 2002;148:2967–2973. doi: 10.1099/00221287-148-10-2967. [DOI] [PubMed] [Google Scholar]
  38. Pinney JW, Shirley MW, McConkey GA, Westhead DR. metaSHARK: software for automated metabolic network prediction from DNA sequence and its application to the genomes of Plasmodium falciparum and Eimeria tenella. Nucleic Acids Res. 2005;33:1399–1409. doi: 10.1093/nar/gki285. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Hotter GS, Wards BJ, Mouat P, Besra GS, Gomes J, Singh M, Bassett S, Kawakami P, Wheeler PR, de Lisle GW, et al. Transposon mutagenesis of Mb0100 at the ppe1-nrp locus in Mycobacterium bovis disrupts phthiocerol dimycocerosate (PDIM) and glycosylphenol-PDIM biosynthesis, producing an avirulent strain with vaccine properties at least equal to those of M. bovis BCG. J Bacteriol. 2005;187:2267–2277. doi: 10.1128/JB.187.7.2267-2277.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Sanger Institute M. tuberculosis H37Rv Functional Classification http://www.sanger.ac.uk/Projects/M_tuberculosis/Gene_list/
  41. Cole ST, Brosch R, Parkhill J, Garnier T, Churcher C, Harris D, Gordon SV, Eiglmeier K, Gas S, Barry CE, III, et al. Deciphering the biology of Mycobacterium tuberculosis from the complete genome sequence. Nature. 1998;393:537–544. doi: 10.1038/31159. [DOI] [PubMed] [Google Scholar]
  42. Dannenberg AM., Jr Roles of cytotoxic delayed-type hypersensitivity and macrophage-activating cell-mediated immunity in the pathogenesis of tuberculosis. Immunobiology. 1994;191:461–473. doi: 10.1016/S0171-2985(11)80452-3. [DOI] [PubMed] [Google Scholar]
  43. Davison A, Hinkley D. Bootstrap Methods and their Application. Cambridge: Cambridge University Press; 1997. [Google Scholar]
  44. Lunneborg C. Data Analysis by Resampling: Concepts and Applications. Pacific Grove, California: Duxbury Press; 2000. [Google Scholar]
  45. McKinney JD, zu Bentrup KH, Munoz-Elias EJ, Miczak A, Chen B, Chan W-T, Swenson D, Sacchettini JC, Jacobs WR, Russell DG. Persistence of Mycobacterium tuberculosis in macrophages and mice requires the glyoxylate shunt enzyme isocitrate lyase. Nature. 2000;406:735–738. doi: 10.1038/35021074. [DOI] [PubMed] [Google Scholar]
  46. Munoz-Elias EJ, McKinney JD. Mycobacterium tuberculosis isocitrate lyases 1 and 2 are jointly required for in vivo growth and virulence. Nat Med. 2005;11:638–644. doi: 10.1038/nm1252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Kana BD, Weinstein EA, Avarbock D, Dawes SS, Rubin H, Mizrahi V. Characterization of the cydAB-Encoded Cytochrome bd oxidase from Mycobacterium smegmatis. J Bacteriol. 2001;183:7076–7086. doi: 10.1128/JB.183.24.7076-7086.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. The Multi-Collaborative Microbial Pathogen Microarray Facility: BuG@S http://bugs.sghms.ac.uk/ [DOI] [PMC free article] [PubMed]
  49. Microarray Data Storage http://bugs.sgul.ac.uk/E-BUGS-60
  50. Belisle JT, Sonnenberg MG. Isolation of genomic DNA from mycobacteria. Methods Mol Biol. 1998;101:31–44. doi: 10.1385/0-89603-471-2:31. [DOI] [PubMed] [Google Scholar]
  51. The Comprehensive R Archive Network http://cran.r-project.org/
  52. Wernisch L, Kendall SL, Soneji S, Wietzorrek A, Parish T, Hinds J, Butcher PD, Stoker NG. Analysis of whole-genome microarray replicates using mixed models. Bioinformatics. 2003;19:53–61. doi: 10.1093/bioinformatics/19.1.53. [DOI] [PubMed] [Google Scholar]
  53. Venables W, Ripley B. Modern Applied Statistics with S. New York: Springer; 2002. [Google Scholar]
  54. Benjamini Y, Yekutieli D. The control of the false discovery rate in multiple testing under dependency. Ann Stat. 2001;29:1165–1188. doi: 10.1214/aos/1013699998. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Additional data file 1

The 198 genes of the 95th percentile; the very abundant transcripts in M. tuberculosis.

Click here for file (60KB, xls)
Additional data file 2

The quantified level of each functional category and details of those deemed significantly more or less abundant in the low oxygen transcriptome, including data from the three approaches to assess significance.

Click here for file (39.5KB, xls)

Articles from Genome Biology are provided here courtesy of BMC

RESOURCES