Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Mutations in SELENBP1, encoding a novel human methanethiol oxidase, cause extraoral halitosis

Abstract

Selenium-binding protein 1 (SELENBP1) has been associated with several cancers, although its exact role is unknown. We show that SELENBP1 is a methanethiol oxidase (MTO), related to the MTO in methylotrophic bacteria, that converts methanethiol to H2O2, formaldehyde, and H2S, an activity not previously known to exist in humans. We identified mutations in SELENBP1 in five patients with cabbage-like breath odor. The malodor was attributable to high levels of methanethiol and dimethylsulfide, the main odorous compounds in their breath. Elevated urinary excretion of dimethylsulfoxide was associated with MTO deficiency. Patient fibroblasts had low SELENBP1 protein levels and were deficient in MTO enzymatic activity; these effects were reversed by lentivirus-mediated expression of wild-type SELENBP1. Selenbp1-knockout mice showed biochemical characteristics similar to those in humans. Our data reveal a potentially frequent inborn error of metabolism that results from MTO deficiency and leads to a malodor syndrome.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Sulfur metabolism.
Fig. 2: Families with extraoral halitosis.
Fig. 3: Analysis of missense mutations in SELENBP1.
Fig. 4: Analysis of SELENBP1 expression and MTO activity in human cell lines.
Fig. 5: MTO activity and DMS levels in Selenbp1-KO mice.

Similar content being viewed by others

References

  1. Tangerman, A. & Winkel, E. G. The portable gas chromatograph OralChroma: a method of choice to detect oral and extraoral halitosis. J. Breath Res. 2, 017010 (2008).

    CAS  PubMed  Google Scholar 

  2. Tangerman, A. & Winkel, E. G. Intra- and extraoral halitosis: finding of a new form of extraoral blood-borne halitosis caused by dimethyl sulphide. J. Clin. Periodontol. 34, 748–755 (2007).

    PubMed  Google Scholar 

  3. Tangerman, A. & Winkel, E. G. Extraoral halitosis: an overview. J. Breath Res. 4, 017003 (2010).

    CAS  PubMed  Google Scholar 

  4. Harvey-Woodworth, C. N. Dimethylsulphidemia: the significance of dimethyl sulphide in extraoral, blood borne halitosis. Br. Dent. J. 214, E20 (2013).

    CAS  PubMed  Google Scholar 

  5. Szabo, C. et al. Regulation of mitochondrial bioenergetic function by hydrogen sulfide. Part I. Biochemical and physiological mechanisms. Br. J. Pharmacol. 171, 2099–2122 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. He, X. & Slupsky, C. M. Metabolic fingerprint of dimethyl sulfone (DMSO2) in microbial-mammalian co-metabolism. J. Proteome Res. 13, 5281–5292 (2014).

    CAS  PubMed  Google Scholar 

  7. Blom, H. J. & Tangerman, A. Methanethiol metabolism in whole blood. J. Lab. Clin. Med. 111, 606–610 (1988).

    CAS  PubMed  Google Scholar 

  8. Walker, V., Mills, G. A., Fortune, P. M. & Wheeler, R. Neonatal encephalopathy with a pungent body odour. Arch. Dis. Child. Fetal Neonatal Ed. 77, F65–F66 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Engelke, U. F. et al. Dimethyl sulfone in human cerebrospinal fluid and blood plasma confirmed by one-dimensional 1H and two-dimensional 1H-13C NMR. NMR Biomed. 18, 331–336 (2005).

    CAS  PubMed  Google Scholar 

  10. Yamagishi, K. et al. Generation of gaseous sulfur-containing compounds in tumour tissue and suppression of gas diffusion as an antitumour treatment. Gut 61, 554–561 (2012).

    CAS  PubMed  Google Scholar 

  11. Machado, R. F. et al. Detection of lung cancer by sensor array analyses of exhaled breath. Am. J. Respir. Crit. Care Med. 171, 1286–1291 (2005).

    PubMed  PubMed Central  Google Scholar 

  12. Li, Z. et al. Real-time ultrasensitive VUV-PIMS detection of representative endogenous volatile markers in cancers. Cancer Biomark. 16, 477–487 (2016).

    CAS  PubMed  Google Scholar 

  13. Buszewski, B., Ulanowska, A., Kowalkowski, T. & Cieśliński, K. Investigation of lung cancer biomarkers by hyphenated separation techniques and chemometrics. Clin. Chem. Lab. Med. 50, 573–581 (2011).

    PubMed  Google Scholar 

  14. Pol, A., Op den Camp, H. J., Mees, S. G., Kersten, M. A. & van der Drift, C. Isolation of a dimethylsulfide-utilizing Hyphomicrobium species and its application in biofiltration of polluted air. Biodegradation 5, 105–112 (1994).

    CAS  PubMed  Google Scholar 

  15. Eyice, Ö. et al. Bacterial SBP56 identified as a Cu-dependent methanethiol oxidase widely distributed in the biosphere. ISME J. https://doi.org/10.1038/ismej.2017.148 (2017).

    PubMed  PubMed Central  Google Scholar 

  16. Raucci, R. et al. Structural and functional studies of the human selenium binding protein-1 and its involvement in hepatocellular carcinoma. Biochim. Biophys. Acta 1814, 513–522 (2011).

    CAS  PubMed  Google Scholar 

  17. Bansal, M. P. et al. DNA sequencing of a mouse liver protein that binds selenium: implications for selenium’s mechanism of action in cancer prevention. Carcinogenesis 11, 2071–2073 (1990).

    CAS  PubMed  Google Scholar 

  18. Bansal, M. P., Oborn, C. J., Danielson, K. G. & Medina, D. Evidence for two selenium-binding proteins distinct from glutathione peroxidase in mouse liver. Carcinogenesis 10, 541–546 (1989).

    CAS  PubMed  Google Scholar 

  19. Jerome-Morais, A. et al. Inverse association between glutathione peroxidase activity and both selenium-binding protein 1 levels and Gleason score in human prostate tissue. Prostate 72, 1006–1012 (2012).

    CAS  PubMed  Google Scholar 

  20. Ansong, E. et al. Evidence that selenium binding protein 1 is a tumor suppressor in prostate cancer. PLoS One 10, e0127295 (2015).

    PubMed  PubMed Central  Google Scholar 

  21. Chen, G. et al. Reduced selenium-binding protein 1 expression is associated with poor outcome in lung adenocarcinomas. J. Pathol. 202, 321–329 (2004).

    CAS  PubMed  Google Scholar 

  22. Willemsen, M. A., Engelke, U. F., van der Graaf, M. & Wevers, R. A. Methylsulfonylmethane (MSM) ingestion causes a significant resonance in proton magnetic resonance spectra of brain and cerebrospinal fluid. Neuropediatrics 37, 312–314 (2006).

    CAS  PubMed  Google Scholar 

  23. Suylen, G. M. H., Large, P. J., van Dijken, J. P. & Kuenen, J. G. Methyl mercaptan oxidase, a key enzyme in the metabolism of methylated sulphur compounds by Hyphomicrobium EG. J. Gen. Microbiol. 133, 2989–2997 (1987).

    CAS  Google Scholar 

  24. Gould, W. D. & Kanagawa, T. Purification and properties of methyl mercaptan oxidase from Thiobacillus thioparus Tk-M. J. Gen. Microbiol. 138, 217–221 (1992).

    CAS  Google Scholar 

  25. Pohl, N. M. et al. Transcriptional regulation and biological functions of selenium-binding protein 1 in colorectal cancer in vitro and in nude mouse xenografts. PLoS One 4, e7774 (2009).

    PubMed  PubMed Central  Google Scholar 

  26. Jeong, J. Y., Wang, Y. & Sytkowski, A. J. Human selenium binding protein-1 (hSP56) interacts with VDU1 in a selenium-dependent manner. Biochem. Biophys. Res. Commun. 379, 583–588 (2009).

    CAS  PubMed  Google Scholar 

  27. Miyaguchi, K. Localization of selenium-binding protein at the tips of rapidly extending protrusions. Histochem. Cell Biol. 121, 371–376 (2004).

    CAS  PubMed  Google Scholar 

  28. Tangerman, A., Meuwese-Arends, M. T. & Jansen, J. B. Cause and composition of foetor hepaticus. Lancet 343, 483 (1994).

    CAS  PubMed  Google Scholar 

  29. Besouw, M., Tangerman, A., Cornelissen, E., Rioux, P. & Levtchenko, E. Halitosis in cystinosis patients after administration of immediate-release cysteamine bitartrate compared to delayed-release cysteamine bitartrate. Mol. Genet. Metab. 107, 234–236 (2012).

    CAS  PubMed  Google Scholar 

  30. Mudd, S. H. et al. Isolated persistent hypermethioninemia. Am. J. Hum. Genet. 57, 882–892 (1995).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Glatt, S. J. et al. Comparative gene expression analysis of blood and brain provides concurrent validation of SELENBP1 up-regulation in schizophrenia. Proc. Natl. Acad. Sci. USA 102, 15533–15538 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Kanazawa, T. et al. The utility of SELENBP1 gene expression as a biomarker for major psychotic disorders: replication in schizophrenia and extension to bipolar disorder with psychosis. Am. J. Med. Genet. B. Neuropsychiatr. Genet. 147B, 686–689 (2008).

    PubMed  Google Scholar 

  33. Prabakaran, S. et al. 2-D DIGE analysis of liver and red blood cells provides further evidence for oxidative stress in schizophrenia. J. Proteome Res. 6, 141–149 (2007).

    CAS  PubMed  Google Scholar 

  34. Furne, J., Springfield, J., Koenig, T., DeMaster, E. & Levitt, M. D. Oxidation of hydrogen sulfide and methanethiol to thiosulfate by rat tissues: a specialized function of the colonic mucosa. Biochem. Pharmacol. 62, 255–259 (2001).

    CAS  PubMed  Google Scholar 

  35. Chang, P. W. et al. Isolation, characterization, and chromosomal mapping of a novel cDNA clone encoding human selenium binding protein. J. Cell. Biochem. 64, 217–224 (1997).

    CAS  PubMed  Google Scholar 

  36. Galvao, J. et al. Unexpected low-dose toxicity of the universal solvent DMSO. FASEB J. 28, 1317–1330 (2014).

    CAS  PubMed  Google Scholar 

  37. Finkel, T. Oxidant signals and oxidative stress. Curr. Opin. Cell Biol. 15, 247–254 (2003).

    CAS  PubMed  Google Scholar 

  38. Barr, L. A. & Calvert, J. W. Discoveries of hydrogen sulfide as a novel cardiovascular therapeutic. Circ. J. 78, 2111–2118 (2014).

    CAS  PubMed  Google Scholar 

  39. Wallace, J. L. & Wang, R. Hydrogen sulfide-based therapeutics: exploiting a unique but ubiquitous gasotransmitter. Nat. Rev. Drug Discov. 14, 329–345 (2015).

    CAS  PubMed  Google Scholar 

  40. Bos, E. M., van Goor, H., Joles, J. A., Whiteman, M. & Leuvenink, H. G. Hydrogen sulfide: physiological properties and therapeutic potential in ischaemia. Br. J. Pharmacol. 172, 1479–1493 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Banerjee, R. Catalytic promiscuity and heme-dependent redox regulation of H2S synthesis. Curr. Opin. Chem. Biol. 37, 115–121 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Yang, H., Nevo, E. & Tashian, R. E. Unexpected expression of carbonic anhydrase I and selenium-binding protein as the only major non-heme proteins in erythrocytes of the subterranean mole rat (Spalax ehrenbergi). FEBS Lett. 430, 343–347 (1998).

    CAS  PubMed  Google Scholar 

  43. Ringrose, J. H. et al. Highly efficient depletion strategy for the two most abundant erythrocyte soluble proteins improves proteome coverage dramatically. J. Proteome Res. 7, 3060–3063 (2008).

    CAS  PubMed  Google Scholar 

  44. Wagner, C. A. Hydrogen sulfide: a new gaseous signal molecule and blood pressure regulator. J. Nephrol. 22, 173–176 (2009).

    CAS  PubMed  Google Scholar 

  45. Yang, J. et al. Erythrocytic hydrogen sulfide production is increased in children with vasovagal syncope. J. Pediatr. 166, 965–969 (2015).

    CAS  PubMed  Google Scholar 

  46. Yang, W. & Diamond, A. M. Selenium-binding protein 1 as a tumor suppressor and a prognostic indicator of clinical outcome. Biomark. Res. 1, 15 (2013).

    PubMed  PubMed Central  Google Scholar 

  47. Mochalski, P. et al. Release and uptake of volatile organic compounds by human hepatocellular carcinoma cells (HepG2) in vitro. Cancer Cell Int. 13, 72 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Fang, W. et al. Functional and physical interaction between the selenium-binding protein 1 (SBP1) and the glutathione peroxidase 1 selenoprotein. Carcinogenesis 31, 1360–1366 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Baliga, M. S. et al. Selenoprotein deficiency enhances radiation-induced micronuclei formation. Mol. Nutr. Food Res. 52, 1300–1304 (2008).

    CAS  PubMed  Google Scholar 

  50. Lubos, E., Loscalzo, J. & Handy, D. E. Glutathione peroxidase-1 in health and disease: from molecular mechanisms to therapeutic opportunities. Antioxid. Redox Signal. 15, 1957–1997 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Jezierski, T., Walczak, M., Ligor, T., Rudnicka, J. & Buszewski, B. Study of the art: canine olfaction used for cancer detection on the basis of breath odour. Perspectives and limitations. J. Breath Res. 9, 027001 (2015).

    PubMed  Google Scholar 

  52. Tamura, K., Stecher, G., Peterson, D., Filipski, A. & Kumar, S. MEGA6: Molecular Evolutionary Genetics Analysis version 6.0. Mol. Biol. Evol. 30, 2725–2729 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Tangerman, A., Meuwese-Arends, M. T. & van Tongeren, J. H. A new sensitive assay for measuring volatile sulphur compounds in human breath by Tenax trapping and gas chromatography and its application in liver cirrhosis. Clin. Chim. Acta 130, 103–110 (1983).

    CAS  PubMed  Google Scholar 

  54. Wevers, R. A., Engelke, U. & Heerschap, A. High-resolution 1H-NMR spectroscopy of blood plasma for metabolic studies. Clin. Chem. 40, 1245–1250 (1994).

    CAS  PubMed  Google Scholar 

  55. Wevers, R. A. et al. Standardized method for high-resolution 1H-NMR of cerebrospinal fluid. Clin. Chem. 41, 744–751 (1995).

    CAS  PubMed  Google Scholar 

  56. Moolenaar, S. H. et al. Prolidase deficiency diagnosed by 1H NMR spectroscopy of urine. J. Inherit. Metab. Dis. 24, 843–850 (2001).

    CAS  PubMed  Google Scholar 

  57. Derikx, P. J., Op Den Camp, H. J., van der Drift, C., van Griensven, L. J. & Vogels, G. D. Odorous sulfur compounds emitted during production of compost used as a substrate in mushroom cultivation. Appl. Environ. Microbiol. 56, 176–180 (1990).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Tangerman, A. Determination of volatile sulphur compounds in air at the parts per trillion level by Tenax trapping and gas chromatography. J. Chromatogr. 366, 205–216 (1986).

    CAS  PubMed  Google Scholar 

  59. Kircher, M. et al. A general framework for estimating the relative pathogenicity of human genetic variants. Nat. Genet. 46, 310–315 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Wortmann, S. B. et al. CLPB mutations cause 3-methylglutaconic aciduria, progressive brain atrophy, intellectual disability, congenital neutropenia, cataracts, movement disorder. Am. J. Hum. Genet. 96, 245–257 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Vriend, G. WHAT IF: a molecular modeling and drug design program. J. Mol. Graph. 8, 52–56 (1990).

    CAS  PubMed  Google Scholar 

  62. Krieger, E., Koraimann, G. & Vriend, G. Increasing the precision of comparative models with YASARA NOVA: a self-parameterizing force field. Proteins 47, 393–402 (2002).

    CAS  PubMed  Google Scholar 

  63. Renkema, G. H. et al. SDHA mutations causing a multisystem mitochondrial disease: novel mutations and genetic overlap with hereditary tumors. Eur. J. Hum. Genet. 23, 202–209 (2015).

    CAS  PubMed  Google Scholar 

  64. Douabul, A. A. & Priley, J. P. Solubility of gases in distilled water and seawater. 5. Hydrogen sulfide. Deep-Sea Res. 26, 259–268 (1979).

    CAS  Google Scholar 

  65. Tan, B. et al. New method for quantification of gasotransmitter hydrogen sulfide in biological matrices by LC-MS/MS. Sci. Rep. 7, 46278 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors gratefully thank the patients and families for their help in this study. The contributions of W. Lehnert, I. Goldschmidt, and C. von Schnakenburg to early investigations on family A are gratefully acknowledged. M. Antoine and T. van Alen are kindly acknowledged for their technical expertise, and Ö. Eyice-Broadbent is acknowledged for work on identifying the bacterial MTO. G. Linthorst, F. Wijburg, and H. Blom are acknowledged for help in obtaining good-quality patient samples. The authors thank P. Klaren for statistical consultation. The work was supported by the following grants: a UK Biotechnology and Biological Sciences Research Council grant to H.S. (reference BB/H003851/1), an ERC grant to H.J.M.O.d.C. (ERC 669371-Volcano), and a grant from the E.C. Noyons Foundation to R.A.W.

Author information

Authors and Affiliations

Authors

Contributions

H.J.M.O.d.C. and R.A.W. conceived the study and coordinated and supervised the different teams; H.J.M.O.d.C., A.P., and G H.R. designed the assays of the sulfur metabolites and MTO enzyme activity; A.P., U.F.E., and A.T. measured sulfur metabolites and/or MTO enzyme activity; E.G.W. performed the Oral Chroma technique for halitosis detection; U.F.E. was responsible for the body-fluid NMR measurements; A.P.M.d.B. provided the tissue expression data of SELENBP1; K.C.L. and R.S.A. made the KO mouse line; L.v.d.H. and M.O.E. performed Sanger sequencing of SELENBP1; H. Omran and H. Olbrich performed linkage analysis and whole-exome sequencing in family A; C.G. performed the frequency calculations for the SELENBP1 defect; R.J.R. and G H.R. performed the cell culture and lentiviral complementation studies; J.O.S., K.O.S., J.S.S., and A.T. performed clinical and biochemical phenotyping of the patients and family members; H.S. and H.J.M.O.d.C. provided the gene sequence of methanethiol oxidase from Hyphomicrobium strain VS; H.V. made the SELENBP1 three-dimensional model and mapped the affected amino acid residues; A.T. collected initial measurements of sulfur-containing metabolites in the blood and breath and carried out several studies on individuals with halitosis; A.P., G.H.R. and R.A.W. prepared the manuscript, to which various coauthors contributed. All coauthors edited and reviewed the final manuscript.

Corresponding author

Correspondence to Ron A. Wevers.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Additional information

Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Integrated Supplementary Information

Supplementary Figure 1 Breath analysis in the dental clinic

Breath samples of the extra-oral halitosis patient CII-2 (dotted line), a control person (dashed line), and an intra-oral halitosis patient (solid line) analysed in a portable gas chromatograph (OralChromaTM). Vertical lines indicate the retention times of H2S, MT and DMS. The vertical line indicated with * is a non-sulfur volatile background peak as described in: Hanada, M. et al. Portable oral malodor analyzer using highly sensitive In2O3 gas sensor combined with a simple gas chromatography system. Analytica Chimica Acta 475, 27-35 (2003).

Supplementary Figure 2 Alignment of the Hyphomicrobium strain VS MT-oxidase and the human SELENBP1

The full-length alignment was performed using the Pairwise Alignment tool of the Protein Information Resource (http://pir.georgetown.edu/pirwww/search/). The Smith-Waterman score was 398. The proteins show 26.0% identity (highlighted in yellow and by *) and 54.2% similarity (highlighted in green) in 461 aa overlap. The signal peptide of the bacterial enzymes is underlined. In red putative TTQ residues (W) and copper ligands (H).  

Supplementary Figure 3 Newly established assay for MTO activity

a MT oxidation results in H2S production. The amounts of MT (circles) and H2S (squares) were followed in time. Two standard assay mixtures (in 250 ml serum bottles) containing MT with- (closed symbols) or without (open symbols) Zn (0.2 mM) were started by the addition of 2 µl human control erythrocyte extract as a MTO source at the time point indicated by the arrowhead. Without Zn in the medium the molar amount of MT that has been converted is seemingly higher that the amount of H2S formed. Zn-ions will capture formed H2S. After MT was depleted the reaction mixtures were acidified thus releasing the H2S again (indicated with an arrow). In the incubation mixture with Zn the amount of sulfide formed is equimolar to the amount of MT that has been converted. In the absence of Zn-ions the formed H2S can be further enzymatically metabolized towards thiosulfate (Ref) thus explaining the lower molar recovery of H2S in the absence of Zn-ions. b Kinetic analysis of the MTO assay. Two identical standard assays using 10 µl human control erythrocyte extract were started at t = 0 under anaerobic conditions. The closed arrowhead indicates the addition of 20 ml of oxygen to incubation indicated with . The open arrowhead indicates the addition of oxygen to the other incubation () twenty minutes later. c MTO activity in colon cancer cell lines HT29 (▲) and SW480 (■). Oxygen dependence of the reaction in 3 ml exetainers is shown with cell line HT29 under anoxic condition (Δ), arrow indicates the addition of 0.25 ml of oxygen. Background disappearance of MT (). Each data point is the average of duplicate incubations that showed less than 5% difference. d Michaelis-Menten plot of MTO activity measurements done with control erythrocyte extract using different substrate concentrations. The apparent Km is indicated. e-g Kinetic analysis of MTO activity in human erythrocyte extract e, HT29 colon cancer cell extracts f and wild type mouse liver extracts (g, duplicate measurements as open and closed symbols) under low starting concentrations of MT. In each panel the line is fitted on the data using Michaelis-Menten kinetics. This resulted in apparent Km values of 5, 4.7 and 6 nM, respectively. Ref: Szabo, C. et al. Regulation of mitochondrial bioenergetic function by hydrogen sulfide. Part I. Biochemical and physiological mechanisms. Br J Pharmacol 171, 2099-122 (2014).

Supplementary Figure 4 Intracellular localization of SELENBP1

Stably transduced patient cells were stained for indirect fluorescence analysis with anti-V5 antibodies that detect the C-terminal tag of SELENBP1 (red). Nuclei were stained with Hoechst dye (blue).

Supplementary Figure 5 THAP4-mutation analysis in pedigree A

Sanger sequencing of the THAP4 mutation in family A shows that the c.1400C>A mutation occurs in homozygous form in the male individual AII-3 with the malodour and the neurological features while in heterozygous form in the female individual AII-2 with the malodour syndrome without neurological features. None of the other 3 sibs in this family has this THAP4 loss of function mutation in homozygous form. Black symbols in the pedigree indicate the malodour in individuals AII-2 and AII-3.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–5, Supplementary Table 2 and Supplementary Note.

Life Sciences Reporting Summary

Supplementary Table 1

Variants from ExAc and in-house databases for CFTR, PAH and SELENBP1. CADD_PHRED scores and variant frequencies were used to calculate the presumed frequency for SELENBP1 mutations, as described in the Supplementary Note

Supplementary Data 1

Full length blots

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pol, A., Renkema, G.H., Tangerman, A. et al. Mutations in SELENBP1, encoding a novel human methanethiol oxidase, cause extraoral halitosis. Nat Genet 50, 120–129 (2018). https://doi.org/10.1038/s41588-017-0006-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-017-0006-7

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing