Comparative genome analysis of Lactobacillus reuteri and Lactobacillus fermentum reveal a genomic island for reuterin and cobalamin production

Hidetoshi Morita, Hidehiro Toh, Shinji Fukuda, Hiroshi Horikawa, Kenshiro Oshima, Takehito Suzuki, Masaru Murakami, Shin Hisamatsu, Yukio Kato, Tatsuya Takizawa, Hideo Fukuoka, Tetsuhiko Yoshimura, Kikuji Itoh, Daniel J O'Sullivan, Larry L McKay, Hiroshi Ohno, Jun Kikuchi, Toshio Masaoka, Masahira Hattori, Hidetoshi Morita, Hidehiro Toh, Shinji Fukuda, Hiroshi Horikawa, Kenshiro Oshima, Takehito Suzuki, Masaru Murakami, Shin Hisamatsu, Yukio Kato, Tatsuya Takizawa, Hideo Fukuoka, Tetsuhiko Yoshimura, Kikuji Itoh, Daniel J O'Sullivan, Larry L McKay, Hiroshi Ohno, Jun Kikuchi, Toshio Masaoka, Masahira Hattori

Abstract

Lactobacillus reuteri is a heterofermentative lactic acid bacterium that naturally inhabits the gut of humans and other animals. The probiotic effects of L. reuteri have been proposed to be largely associated with the production of the broad-spectrum antimicrobial compound reuterin during anaerobic metabolism of glycerol. We determined the complete genome sequences of the reuterin-producing L. reuteri JCM 1112(T) and its closely related species Lactobacillus fermentum IFO 3956. Both are in the same phylogenetic group within the genus Lactobacillus. Comparative genome analysis revealed that L. reuteri JCM 1112(T) has a unique cluster of 58 genes for the biosynthesis of reuterin and cobalamin (vitamin B(12)). The 58-gene cluster has a lower GC content and is apparently inserted into the conserved region, suggesting that the cluster represents a genomic island acquired from an anomalous source. Two-dimensional nuclear magnetic resonance (2D-NMR) with (13)C(3)-glycerol demonstrated that L. reuteri JCM 1112(T) could convert glycerol to reuterin in vivo, substantiating the potential of L. reuteri JCM 1112(T) to produce reuterin in the intestine. Given that glycerol is shown to be naturally present in feces, the acquired ability to produce reuterin and cobalamin is an adaptive evolutionary response that likely contributes to the probiotic properties of L. reuteri.

Figures

Figure 1
Figure 1
Genome-based phylogenetic analysis of L. reuteri JCM 1112T and L. fermentum IFO 3956. (A) Synteny between the L. reuteri JCM 1112T and L. fermentum IFO 3956 chromosomes. Each dot represents an orthologous gene that was defined by bidirectional best hits based on BLASTP comparisons. Genome numbering was initiated at dnaA in both chromosomes. The shaded region indicates the location of the pdu-cbi-cob-hem cluster in the L. reuteri JCM 1112T genome. (B) Phylogenetic relationships between the genomes of sequenced Lactobacillus species and other lactic acid bacteria, including Lactococcus lactis, inferred from 34 concatenated ribosomal protein amino acid sequences. The scale bar represents an evolutionary distance. Sequences were aligned with ClustalW with a bootstrap trial of 1000 and bootstrap values (%) are indicated at the nodes. An unrooted tree was generated using NJplot.
Figure 2
Figure 2
Proposed glycerol and glucose metabolic pathways in L. reuteri JCM 1112T. The pathways are color coded as follows: blue, glycerol metabolism; green, biosynthesis of reuterin; red, glucose metabolism. Dashed lines indicate an unidentified enzyme in L. reuteri JCM 1112T. The bottom part of the figure shows the structure of the pdu-cbi-cob-hem gene cluster in L. reuteri JCM 1112T. Genes are depicted with arrows indicating the transcription direction with the following colors: yellow, pdu including gupCDE genes; pink, cbi genes; orange, cob genes; blue, hem genes; red, pocR; green, eut; sky blue, transposase gene; and white, other genes. Several lines connect corresponding genes between the pathway and the cluster (enzymes, red; transporters, blue).
Figure 3
Figure 3
(A) A comparison of the genomic location that contains the pdu-cbi-cob-hem gene cluster of L. reuteri JCM 1112T (center) with the corresponding location of L. fermentum IFO 3956 (upper) and L. plantarum WCFS1 (lower). Genes in the pdu-cbi-cob-hem gene cluster are depicted by arrows indicating the transcription direction with the same color codes as described in the Fig. 2. Genes conserved between the three genomes are colored gray and light blue bars indicate orthologous regions. The GC content at the third codon position of the ORFs in L. reuteri JCM 1112T is indicated under each ORF. Red lines represent the GC content at the third codon position of the ORFs in the pdu-cbi-cob-hem cluster (LAR_1583-1640). The blue horizontal line indicates the average GC content (32%) at the third codon position of the remaining ORFs in the L. reuteri JCM 1112T genome excluding the pdu-cbi-cob-hem gene cluster. (B) The pdu-cbi-cob gene cluster arrangement in L. reuteri JCM 1112T, S. typhimurium LT2, L. monocytogenes EGD-e, Y. enterocolitica subsp. enterocolitica 8081, and S. sanguinis SK36 are shown using the same color coding as described in (A).
Figure 4
Figure 4
In vivo detection of 3-HPA-hydrate derived from 13C3-glycerol in the cecal contents of mice colonized by wild-type L. reuteri JCM 1112T (A) and its mutant derivative, LRΔgupCDE (B), as detected by 2D 1H, 13C-HSQC NMR. δ1H (ppm) and δ13C (ppm) of 3-HPA-hydrate 2 (Hh-2) and 3-HPA-hydrate 3 (Hh-3) were observed at 2.43 and 42.6, and 3.79 and 61.6 at pH 7.0, respectively. Although the spot derived from 3-HPA-hydrate 1 (Hh-1) in 1H chemical shift was detected, the spot was not shown in (A), because the range of 1H chemical shift showed between 4.0 to 1.5 ppm. See Supplementary Figure S5 for details.

References

    1. Casas I. A., Dobrogosz W. J. Lactobacillus reuteri: an overview of a new probiotic for humans and animals. Microecol. Therap. 1997;26:221–231.
    1. Casas I. A., Dobrogosz W. J. Validation of the probiotic concept: Lactobacillus reuteri confers broad-spectrum protection against disease in humans and animals. Microb. Ecol. Health Dis. 2000;12:247–285.
    1. Shornikova A. V., Casas I. A., Isolauri E., Mykkänen H., Vesikari T. Lactobacillus reuteri as a therapeutic agent in acute diarrhea in young children. J. Pediatr. Gastroenterol. Nutr. 1997;24:399–404.
    1. Valeur N., Engel P., Carbajal N., Connolly E., Ladefoged K. Colonization and immunomodulation by Lactobacillus reuteri ATCC 55730 in the human gastrointestinal tract. Appl. Environ. Microbiol. 2004;70:1176–1181.
    1. Chang Y. H., Kim J. K., Kim H. J., Kim W. Y., Kim Y. B., Park Y. H. Selection of a potential probiotic Lactobacillus strain and subsequent in vivo studies. Antonie van Leeuwenhoek. 2001;80:193–199.
    1. Rosenfeldt V., Michaelsen K. F., Jakobsen M., et al. Effect of probiotic Lactobacillus strains on acute diarrhea in a cohort of nonhospitalized children attending day-care centers. Pediatr. Infect. Dis. J. 2002;21:417–419.
    1. Weizman Z., Asli G., Alsheikh A. Effect of a probiotic infant formula on infections in child care centers: comparison of two probiotic agents. Pediatrics. 2005;115:5–9.
    1. Niv E., Naftali T., Hallak R., Vaisman N. The efficacy of Lactobacillus reuteri ATCC 55730 in the treatment of patients with irritable bowel syndrome: a double blind, placebo-controlled, randomized study. Clin. Nutr. 2005;24:925–931.
    1. Talarico T. L., Casas I. A., Chung T. C., Dobrogosz W. J. Production and isolation of reuterin: a growth inhibitor produced by Lactobacillus reuteri. Antimicrob. Agents Chemother. 1988;32:1854–1858.
    1. Chung T. C., Axelsson L., Lindgren S. E., Dobrogosz W. J. In vitro studies on reuterin synthesis by Lactobacillus reuteri. Microb. Ecol. Health Dis. 1989;2:137–144.
    1. Talarico T. L., Dobrogosz W. J. Chemical characterization of an antimicrobial substance produced by Lactobacillus reuteri. Antimicrob. Agents Chemother. 1989;33:674–679.
    1. Lüthi-Peng Q., Schärer S., Puhan Z. Production and stability of 3-hydroxypropionaldehyde in Lactobacillus reuteri. Appl. Microbiol. Biotechnol. 2002;60:73–80.
    1. Vollenweider S., Lacroix C. 3-Hydroxypropionaldehyde: applications and perspectives of biotechnological production. Appl. Microbiol. Biotechnol. 2004;64:16–27.
    1. Ennahar S., Cai Y., Fujita Y. Phylogenetic diversity of lactic acid bacteria associated with paddy rice silage as determined by 16S ribosomal DNA analysis. Appl. Environ. Microbiol. 2003;69:444–451.
    1. Klein G., Pack A., Bonaparte C., Reuter G. Taxonomy and physiology of probiotic lactic acid bacteria. Int. J. Food Microbiol. 1998;41:103–125.
    1. Strompfová V., Marcináková M., Simonová M., Bogovic-Matijasić B., Lauková A. Application of potential probiotic Lactobacillus fermentum AD1 strain in healthy dogs. Anaerobe. 2006;12:75–79.
    1. Peran L., Camuesco D., Comalada M., et al. Lactobacillus fermentum, a probiotic capable to release glutathione, prevents colonic inflammation in the TNBS model of rat colitis. Int. J. Colorectal Dis. 2006;21:737–746.
    1. Geier M. S., Butler R. N., Giffard P. M., Howarth G. S. Lactobacillus fermentum BR11, a potential new probiotic, alleviates symptoms of colitis induced bydextran sulfate sodium (DSS) in rats. Int. J. Food Microbiol. 2007;114:267–274.
    1. Kleerebezem M., Boekhorst J., van Kranenburg R., et al. Complete genome sequence of Lactobacillus plantarum WCFS1. Proc. Natl. Acad. Sci. USA. 2003;100:1990–1995.
    1. Pridmore R. D., Berger B., Desiere F., et al. The genome sequence of the probiotic intestinal bacterium Lactobacillus johnsonii NCC 533. Proc. Natl. Acad. Sci. USA. 2004;101:2512–2517.
    1. Altermann E., Russell W. M., Azcarate-Peril M. A., et al. Complete genome sequence of the probiotic lactic acid bacterium Lactobacillus acidophilus NCFM. Proc. Natl. Acad. Sci. USA. 2005;102:3906–3912.
    1. Chaillou S., Champomier-Vergès M. C., Cornet M., et al. Complete genome sequence of the meat-borne lactic acid bacterium Lactobacillus sakei 23K. Nat. Biotechnol. 2005;23:1527–1533.
    1. Claesson M. J., Li Y., Leahy S., et al. Multireplicon genome architecture of Lactobacillus salivarius. Proc. Natl. Acad. Sci. USA. 2006;103:6718–6723.
    1. van de Guchte M., Penaud S., Grimaldi C., et al. The complete genome sequence of Lactobacillus bulgaricus reveals extensive and ongoing reductive evolution. Proc. Natl. Acad. Sci. USA. 2006;103:9274–9279.
    1. Makarova K., Slesarev A., Wolf Y., et al. Comparative genomics of the lactic acid bacteria. Proc. Natl. Acad. Sci. USA. 2006;103:15611–15616.
    1. Callanan M., Kaleta P., O'Callaghan J., et al. Genome sequence of Lactobacillus helveticus, an organism distinguished by selective gene loss and insertion sequence element expansion. J. Bacteriol. 2008;190:727–735.
    1. Gordon D., Desmarais C., Green P. Automated finishing with autofinish. Genome Res. 2001;11:614–625.
    1. Delcher A. L., Harmon D., Kasif S., White O., Salzberg S. L. Improved microbial gene identification with GLIMMER. Nucleic Acids Res. 1999;27:4636–4641.
    1. Lowe T., Eddy S. R. tRNAscan-SE: a program for improved detection of transfer RNA genes in genomic sequence. Nucleic Acids Res. 1997;25:955–964.
    1. Tatusov R. L., Natale D. A., Garkavtsev I. V., et al. The COG database: new developments in phylogenetic classification of proteins from complete genomes. Nucleic Acids Res. 2001;29:22–28.
    1. Bateman A., Coin L., Durbin R., et al. The Pfam protein families database. Nucleic Acids Res. 2004;32:138–141.
    1. Rice P., Longden I., Bleasby A. EMBOSS: the European molecular biology open software suite. Trends Genet. 2000;16:276–277.
    1. Tamura M., Hirayama K., Itoh K. Comparison of colonic bacterial enzymes in gnotobiotic mice monoassociated with different intestinal bacteria. Microb. Ecol. Health Dis. 1996;9:287–294.
    1. Sakaguchi T., Sakaguchi S., Nakamura I., Kudo Y., Ichiman Y., Yoshida K. Positive reaction in mouse ligated intestinal loop assay with nonenterotoxigenic and nonhemolytic strains of Staphylococcus aureus. J. Clin. Microbiol. 1988;26:600–601.
    1. Kikuchi J., Hirayama T. Practical aspects of stable isotope labeling of higher plants for a hetero-nuclear multi-dimensional NMR-based metabolomics. Methods Mol. Biol. 2006;358:273–286.
    1. Vuister G. W., Bax A. Resolution enhancement and spectral editing of uniformly C-13-enriched proteins by homonuclear broad-band C-13 decoupling. J. Magn. Reson. 1992;98:428–435.
    1. Parrella F., Sanchez-Ferrando A., Virgili A. Improved sensitivity in gradient-based 1D and 2D multiplicity-edited HSQC experiments. J. Magn. Reson. 1997;126:274–277.
    1. Schmieder P., Kurz M., Kessler H. Determination of heteronuclear long-range couplings to heteronuclei in natural abundance by two- and three-dimensional NMR spectroscopy. J. Biomol. NMR. 1991;1:403–420.
    1. Kandler O., Stetter K. O., Kohl R. Lactobacillus reuteri sp. nov., a new species of heterofermentative lactobacilli. Zentbl. Bakteriol. Mikrobiol. Hyg. I Abt. Orig., 1980;C1:264–269.
    1. Molenaar D., Bringel F., Schuren F. H., et al. Exploring Lactobacillus plantarum genome diversity by using microarrays. J. Bacteriol. 2005;187:6119–6127.
    1. Daniel R., Bobik T. A., Gottschalk G. Biochemistry of coenzyme B12-dependent glycerol and diol dehydratases and organization of the encoding genes. FEMS Microbiol. Rev. 1998;22:553–566.
    1. Talarico T. L., Dobrogosz W. J. Purification and characterization of glycerol dehydratase from Lactobacillus reuteri. Appl. Environ. Microbiol. 1990;56:1195–1197.
    1. Bobik T. A., Xu Y., Jeter R. M., et al. Propanediol utilization genes (pdu) of Salmonella typhimurium: three genes for the propanediol dehydratase. J. Bacteriol. 1997;179:6633–6639.
    1. Sauvageot N., Muller C., Hartke A., Auffray Y., Laplace J. M. Characterisation of the diol dehydratase pdu operon of Lactobacillus collinoides. FEMS Microbiol. Lett. 2002;209:69–74.
    1. Gorga A., Claisse O., Lonvaud-Funel A. Organisation of the genes encoding glycerol dehydratase of Lactobacillus collinoides, Lactobacillus hilgardii and Lactobacillus diolivorans. Sci. Aliments. 2002;22:151–160.
    1. Lawrence J. G. Gene organization: selection, selfishness, and serendipity. Annu. Rev. Microbiol. 2003;57:419–440.
    1. Bobik T. A., Ailion M., Roth J. R. A single regulatory gene integrates control of vitamin B12 synthesis and propanediol degradation. J. Bacteriol. 1992;174:2253–2266.
    1. McClelland M., Sanderson K. E., Spieth J., et al. Complete genome sequence of Salmonella enterica serovar Typhimurium LT2. Nature. 2001;413:852–856.
    1. Glaser P., Frangeul L., Buchrieser C., et al. Comparative genomics of Listeria species. Science. 2001;294:849–852.
    1. Yang F., Yang J., Zhang X., et al. Genome dynamics and diversity of Shigella species, the etiologic agents of bacillary dysentery. Nucleic Acids Res. 2005;33:6445–6458.
    1. Thomson N. R., Howard S., Wren B. W., et al. The complete genome sequence and comparative genome analysis of the high pathogenicity Yersinia enterocolitica strain 8081. PLoS Genet. 2006;2
    1. Xu P., Alves J. M., Kitten T., et al. Genome of the opportunistic pathogen Streptococcus sanguinis. J. Bacteriol. 2007;189:3166–3175.
    1. Sharp P. M., Li W. H. The codon adaptation index: a measure of directional synonymous codon usage bias, and its potential applications. Nucleic Acids Res. 1987;15:1281–1295.
    1. Barbirato F., Larguier A., Conte T., Astruc S., Bories A. Sensitivity to pH, product inhibition and inhibition by NAD+ of 1,3-propanediol dehydrogenase purified from Enterobacter agglomerans CNCM 1210. Arch. Microbiol. 1997;168:160–163.
    1. Taranto M. P., Vera J. L., Hugenholtz J., de Valdez G. F., Sesma F. Lactobacillus reuteri CRL1098 produces cobalamin. J. Bacteriol. 2003;185:5643–5647.
    1. Sauvageot N., Pichereau V., Louarme L., Hartke A., Auffray Y., Laplace J. M. Purification, characterization and subunits identification of the diol dehydratase of Lactobacillus collinoides. Eur. J. Biochem. 2002;269:5731–5737.
    1. Porwollik S., Wong R.M., McClelland M. Evolutionary genomics of Salmonella: gene acquisitions revealed by microarray analysis. Proc. Natl. Acad. Sci. USA. 2002;99:8956–8961.

Source: PubMed

3
Suscribir