Skip to main content
Nucleic Acids Research logoLink to Nucleic Acids Research
. 2003 Jan 1;31(1):187–189. doi: 10.1093/nar/gkg004

HGT-DB: a database of putative horizontally transferred genes in prokaryotic complete genomes

S Garcia-Vallve 1,*, E Guzman 1, M A Montero 1, A Romeu 1
PMCID: PMC165451  PMID: 12519978

Abstract

The Horizontal Gene Transfer DataBase (HGT-DB) is a genomic database that includes statistical parameters such as G+C content, codon and amino-acid usage, as well as information about which genes deviate in these parameters for prokaryotic complete genomes. Under the hypothesis that genes from distantly related species have different nucleotide compositions, these deviated genes may have been acquired by horizontal gene transfer. The current version of the database contains 88 bacterial and archaeal complete genomes, including multiple chromosomes and strains. For each genome, the database provides statistical parameters for all the genes, as well as averages and standard deviations of G+C content, codon usage, relative synonymous codon usage and amino-acid content. It also provides information about correspondence analyses of the codon usage, plus lists of extraneous group of genes in terms of G+C content and lists of putatively acquired genes. With this information, researchers can explore the G+C content and codon usage of a gene when they find incongruities in sequence-based phylogenetic trees. A search engine that allows searches for gene names or keywords for a specific organism is also available. HGT-DB is freely accessible at http://www.fut.es/~debb/HGT.

INTRODUCTION

Horizontal Gene Transfer (HGT), the transfer of genes between different species, is recognized as one of the major forces in prokaryotic genome evolution (1). Acquired genes may provide novel metabolic capabilities and catalyze the diversification of microbial lineages. HGT events can be detected from patterns of best matches to different species and the distribution of genes, or by identifying regions of the genome with unusual compositions or incongruities between phylogenetic trees (2,3). Each of these methods has its advantages and disadvantages (2). The prediction of horizontally transferred genes using atypical nucleotide composition is based on the genome hypothesis (4) that assumes that codon usage and G+C content are distinct global features of each prokaryotic genome. With this method, a significant number of prokaryotic genes have been proposed as having been acquired by HGT (5,6). However, it cannot predict all acquired genes unambiguously (7) because genes may have adjusted to the base composition and codon usage of the host genome (this is called the amelioration process) or because an unusual composition may be due to factors other than HGT (6). Despite these limitations, atypical G+C content and patterns of codon usage are especially useful for detecting the putative origin of the transferred genes (810).

To confirm whether a gene or group of genes has been acquired by HGT, it can be useful to combine multiple lines of evidence (2). If researchers have access to the compositional parameters for each gene from complete genomes, they will be able to explore for themselves the G+C content and codon usage of genes when they find incongruences among sequence-based phylogenetic trees or when they detect putatively transferred genes with other methods. We have, therefore, created the Horizontal Gene Transfer DataBase (HGT-DB) to facilitate compositional analyses and provide additional evidence for discussing the possible foreign origin of the genes of a genome and detecting whether acquired genes have been ameliorated. For each prokaryotic complete genome, the HGT-DB provides averages and standard deviations of G+C content, codon usage, relative synonymous codon usage and amino-acid content, as well as lists of putative horizontally transferred genes, correspondence analyses of the codon usage and lists of extraneous groups of genes in terms of G+C content. For each gene, the database lists several statistical parameters, including total and positional G+C content, and determines whether the gene deviates from the mean values of its own genome. The HGT-DB has so far been used to study strain-specific genes of Helicobacter pylori (11,12) and to exclude putative horizontally transferred genes in genomic or proteomic analyses (13).

SOURCES OF GENOMIC DATA AND METHODS

Sequence files of prokaryotic complete genomes are retrieved from the NCBI ftp server. Total and positional G+C content, codon usage, relative synonymous codon usage and amino-acid content are calculated for each gene. For each genome, except for genes under 300 bp, which can have extraneous compositional values, the averages and standard deviations of the above parameters are calculated. The methods we used to consider whether a gene is extraneous in terms of G+C content or codon usage and a candidate to be acquired by HGT are described in Garcia-Vallve et al. (6). Briefly, genes are considered as extraneous in terms of G+C content or codon usage if they deviate by more than 1.5 standard deviations from the mean values. Genes are considered to be putative horizontally transferred genes if they have extraneous G+C content and codon usage, they are over 300 bp and they do not deviate from the average amino-acid composition. Clusters of genes with a high or low G+C content are also considered to be acquired genes, regardless of their length or codon usage (6). It is important to distinguish highly expressed genes from horizontally transferred genes (6). Highly expressed genes may deviate from the mean values of codon usage because they adapt their codon usage to the more abundant tRNAs. For this reason, ribosomal proteins, a group of highly expressed genes, are filtered and not included in the database predictions. Other groups of highly expressed genes will be included in future versions of the database, but individual analyses to define the group of highly expressed genes for each genome, if there are any, will probably be needed.

Genes proposed as being acquired horizontally are represented in a correspondence analysis in which protein-coding sequences are considered as points in a 59-dimensional space (the stop codons and codons for methionine and tryptophan are not included), and each dimension corresponds to the relative frequency of use of each codon measured with the relative synonymous codon usage (RSCU) values. Correspondence analysis reduces this multidimensional space to a two- or three-dimensional space that can be represented graphically. In these graphs, vertically descended genes are expected to cluster together around the origin, whereas genes predicted as acquisitions are expected to be on the periphery.

ORGANIZATION OF THE DATABASE

The HGT-DB is organized by genome i.e. every prokaryotic genome that has been completely sequenced forms a new entry. Different chromosomes from the same organism, or genomes from the same species but different strains, are found in different entries. The current version of the database contains 88 genomes that are sorted alphabetically and classified taxonomically. Table 1 shows the archaeal and bacterial genomes included in the current version of the database, as well as the number of extraneous genes in terms of G+C content and codon usage. The main page for each genome contains links to additional sections and the mean values and standard deviations of total and positional G+C content, codon usage, relative synonymous codon usage and amino-acid content. The other sections available for each genome are: a correspondence analysis of the codon usage, a list of extraneous regions in terms of G+C content and a list of the proposed horizontally acquired genes. The database also provides access to a tab-delimited file with all the statistical calculations for each gene of a genome. The fields available for each gene in these files are: information about its position (coordinates, strand and length), gene name, function, the Cluster of Orthologous Group, COG, (14) it belongs to, total and positional G+C content, the Mahalanobis distance to the average codon usage (6), amino-acid content deviations, if any, and a prediction of whether the gene belongs to a region with a high or low G+C content or whether it has been acquired by HGT. This information can be also accessed via a search engine that allows searches for gene names or keywords for a specific organism. When searching for a gene name, one can also view the upstream and downstream genes.

Table 1. Species, total number of Open Reading Frames (ORF) and number (N) and percentage (%) of extraneous genes in terms of G+C content and codon usage from archaeal and bacterial complete genomes included in the database.

Genome ORF N %
Archaea   Archaea Archaea
Aeropyrum pernix K1 1840 270 15.7
Archaeoglobus fulgidus 2420 160 7.7
Halobacterium sp. NRC-1 2075 149 8.4
Methanobacterium thermoautotrophicum deltaH 1873 178 10.9
Methanococcus jannaschii 1729 72 4.8
Methanopyrus kandleri AV19 1687 179 11.5
Methanosarcina acetivorans 4540 602 15.1
Methanosarcina mazei 3371 378 12.6
Pyrobaculum aerophilum 2605 308 14.5
Pyrococcus abyssi 1769 121 7.3
Pyrococcus furiosis 2065 134 7.4
Pyrococcus horikoshii 1801 123 7.3
Sulfolobus solfataricus 2977 147 5.4
Sulfolobus tokodaii 2826 132 5.2
Thermoplasma acidophilum 1482 145 10.8
Thermoplasma volcanium 1499 104 7.8
Bacteria      
Agrobacterium tumefaciens str. C58 (Cereon) circular chromosome 2721 194 7.6
A. tumefaciens str. C58 (Cereon) linear chr. 1833 114 6.5
A. tumefaciens str. C58 (U. Wash.) circular chr. 2785 142 5.7
A. tumefaciens str. C58 (U. Wash.) linear chr. 1876 114 6.5
Aquifex aeolicus 1529 70 4.8
Bacillus halodurans C-125 4066 304 8.6
Bacillus subtilis 4112 552 15.0
Borrelia burgdorferi 851 10 1.4
Brucella melitensis chr. I 2059 118 6.5
Brucella melitensis chr. II 1139 59 5.7
Buchnera aphidicola Sg 544 6 1.3
Buchnera sp. APS 564 0 0.0
Campylobacter jejuni 1634 78 5.4
Caulobacter crescentus 3737 135 3.9
Chlorobium tepidum TLS 2252 267 14.5
Chlamydophila pneumoniae J138 1069 49 5.2
Chlamydophila pneumoniae CWL029 1054 58 6.0
Chlamydophila pneumoniae AR39 1112 55 5.9
Chlamydia trachomatis 895 36 4.3
Chlamydia muridarum 909 12 1.5
Clostridium acetobutylicum ATCC824 3672 146 4.4
Clostridium perfringens 2660 75 3.2
Corynebacterium glutamicum 3040 207 7.5
Deinococcus radiodurans chr. 1 2629 86 3.5
Deinococcus radiodurans chr. 2 368 23 6.4
Escherichia coli K12 4279 359 9.2
Escherichia coli O157 5361 625 13.3
Escherichia coli O157:H7:EDL933 5324 593 12.6
Fusobacterium nucleatum ATCC25586 2067 40 2.2
Haemophilus influenzae Rd 1714 87 5.7
Helicobacter pylori 26695 1576 87 6.3
Helicobacter pylori J99 1491 68 4.9
Lactococcus lactis 2267 90 4.5
Listeria innocua 2968 164 6.2
Listeria monocytogenes EGD-e 2846 184 7.1
Mesorhizobium loti 6746 604 9.9
Mycobacterium leprae TN 1605 73 5.1
Mycobacterium tuberculosis H37Rv 3927 176 4.8
Mycobacterium tuberculosis CDC1551 4187 197 5.4
Mycoplasma genitalium G37 484 51 11.9
Mycoplasma pneumoniae M129 689 39 6.2
Mycoplasma pulmonis UAB CTIP 782 28 4.0
Neisseria meningitidis MC58 2079 221 12.5
Neisseria meningitidis Z2491 2065 206 11.7
Nostoc sp. PCC 7120 5366 203 4.4
Pasteurella multocida PM70 2015 117 6.1
Pseudomonas aeruginosa PA01 5567 307 5.9
Ralstonia solanacearum 3440 356 11.2
Rickettsia conorii Malish 7 1374 54 5.6
Rickettsia prowazekii MadridE 835 28 3.6
Salmonella entereica serovar typhi 4395 551 13.9
Salmonella enterica serovar typhimurium LT2 4451 446 11.0
Sinorhizobium meliloti 1021 3341 179 5.8
Staphylococcus aureus Mu50 2714 119 5.1
Staphylococcus aureus MW2 2632 131 5.8
Staphylococcus aureus N315 2594 105 4.6
Streptococcus pneumonia R6 2043 249 14.1
Streptococcus pneumonia TIGR4 2094 258 15.1
Streptococcus pyogenes SF320 1697 136 9.1
Streptococcus pyogenes MGAS8232 1845 157 10.0
Streptomyces coelicolor A3(2) 7512 541 7.8
Synechocystis PCC6803 3167 211 7.3
Thermoanaerobacter tengcongensis 2588 343 14.9
Thermotoga maritima 1858 194 11.6
Treponema pallidum subsp. pallidum 1036 78 8.7
Ureaplasma urealyticum 614 12 2.3
Vibrio cholerae chr. 1 2742 234 10.0
Vibrio cholerae chr. 2 1093 204 22.2
Xanthomonas campestris 4181 285 7.4
Xanthomonas citri 4312 284 7.1
Xylella fastidiosa 2766 458 21.4
Yersinia pestis CO92 3885 316 9.0

The percentages are referred to the genes analyzed, that exclude genes smaller than 300 bp and genes for ribosomal proteins

Forces other than HGT are also responsible for the heterogeneity in the codon usage of all the genes of a genome. The HGT-DB, therefore, has a section containing the correspondence analysis of the relative synonymous codon usage for each genome. This section contains a table with the percentage variability of the six axes that account for the greatest variation in codon usage, a graphical representation of the coordinates of each gene in the first and second axes (the genes proposed as being acquired by HGT and putative highly expressed genes are shown in different colors) and a table with the correlation values between the position of genes in the first or second axis, and the G+C content and several indices of codon bias. These indices are: the effective number of codons (Nc) (15), the intrinsic codon deviation index (ICDI) (16), the translational efficiency index (P2) (17) and the scaled X2 index (18).

DATABASE ACCESS

HGT-DB is freely accessible at http://www.fut.es/~debb/HGT/. The database will be updated several times each year. Changes and new additions to the database can be viewed in the ‘news and previous release’ section.

Acknowledgments

ACKNOWLEDGEMENTS

We thank Kevin Costello of the Language Service of the Rovira i Virgili University for his help with writing the manuscript, and TINET for hosting the database.

REFERENCES

  • 1.Koonin E.V., Makarova,K.S. and Aravind,L. (2001) Horizontal gene transfer in prokaryotes: Quantification and Classification. Annu. Rev. Microbiol., 55, 709–742. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Eisen J.A. (2000) Horizontal gene transfer among microbial genomes: new insights from complete genome analysis. Curr. Opin. Genet. Dev., 10, 606–611. [DOI] [PubMed] [Google Scholar]
  • 3.Ragan M.A. (2001) Detection of lateral gene transfer among microbial genomes. Curr. Opin. Genet. Dev., 11, 620–626. [DOI] [PubMed] [Google Scholar]
  • 4.Grantham R., Gautier,C., Gouy,M., Mercier,R. and Pave,A. (1980) Codon catalog usage and the genome hypothesis. Nucleic Acids Res., 8, r49–r62. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Ochman H., Lawrence,J.G. and Groisman,E.A. (2000) Lateral gene transfer and the nature of bacterial innovation. Nature, 405, 299–304. [DOI] [PubMed] [Google Scholar]
  • 6.Garcia-Vallve S., Romeu,A. and Palau,J. (2000) Horizontal gene transfer in bacterial and archaeal complete genomes. Genome Res., 10, 1719–1725. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Lawrence J.G. and Ochman,H. (2002) Reconciling the many faces of lateral gene transfer. Trends Microbiol., 10, 1–4. [DOI] [PubMed] [Google Scholar]
  • 8.Garcia-Vallve S., Palau,J. and Romeu,A. (1999) Horizontal gene transfer in glycosyl hydrolases inferred from codon usage in Escherichia coli and Bacillus subtilis. Mol. Biol. Evol., 16, 1125–1134. [DOI] [PubMed] [Google Scholar]
  • 9.Garcia-Vallve S., Romeu,A. and Palau,J. (2000) Horizontal gene transfer of glycosyl hydrolases of the rumen fungi. Mol. Biol. Evol., 17, 352–361. [DOI] [PubMed] [Google Scholar]
  • 10.Garcia-Vallve S., Simó,F.X., Montero,M.A., Arola,L. and Romeu,A. (2002) Simultaneous horizontal gene transfer of a gene coding for ribosomal protein L27 and operational genes in Arthrobacter sp. J. Mol. Evol., in press. [DOI] [PubMed] [Google Scholar]
  • 11.Israel D.A., Salama,N., Krishna,U., Rieger,U.M., Atherton,J.C., Falkow,S. and Peek,R.M.,Jr (2001) Helicobacter pylori genetic diversity within the gastric niche of a single human host. Proc. Natl Acad. Sci. USA, 98, 14625–14630. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Garcia-Vallve S., Janssen,P.J. and Ouzounis,C.A. (2002) Genetic variation between Helicobacter pylori strains: gene acquisition or loss? Trends Microbiol., 10, 445–447. [DOI] [PubMed] [Google Scholar]
  • 13.Akashi H. and Gojobori,T. (2002) Metabolic efficiency and amino acid composition in the proteomes of Escherichia coli and Bacillus subtilis. Proc. Natl Acad. Sci. USA, 99, 3695–3700. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Tatusov R.L., Natale,D.A., Garkavtsev,I.V., Tatusova,T.A., Shankavaram,U.T., Rao,B.S., Kiryutin,B., Galperin,M.Y., Fedorova,N.D. and Koonin,E.V. (2001) The COG database: new developments in phylogenetic classification of proteins from complete genomes. Nucleic Acids Res., 29, 22–38. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Wright F. (1990) The effective number of codons used in a gene. Gene, 87, 23–29. [DOI] [PubMed] [Google Scholar]
  • 16.Freire-Picos M.A., Gonzalez-Siso,M.I., Rodriguez-Belmonte,E., Rodriguez-Torres,A.M., Ramil,E. and Cerdan,M.E. (1994) Codon usage in Kluyveromyces lactis and in yeast cytochrome c-encoding genes. Gene, 139, 43–49. [DOI] [PubMed] [Google Scholar]
  • 17.Gouy M. and Gautier,C. (1982) Codon usage in bacteria-correlation with gene expressivity. Nucleic Acids Res., 10, 7055–7074. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Shields D.C. and Sharp,P.M. (1987) Synonymous codon usage in Bacillus subtilis reflects both translational selection and mutational biases. Nucleic Acids Res., 15, 8023–8040. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Nucleic Acids Research are provided here courtesy of Oxford University Press

RESOURCES