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.

  • Letter
  • Published:

CLCN2 chloride channel mutations in familial hyperaldosteronism type II

Abstract

Primary aldosteronism, a common cause of severe hypertension1, features constitutive production of the adrenal steroid aldosterone. We analyzed a multiplex family with familial hyperaldosteronism type II (FH-II)2 and 80 additional probands with unsolved early-onset primary aldosteronism. Eight probands had novel heterozygous variants in CLCN2, including two de novo mutations and four independent occurrences of a mutation encoding an identical p.Arg172Gln substitution; all relatives with early-onset primary aldosteronism carried the CLCN2 variant found in the proband. CLCN2 encodes a voltage-gated chloride channel expressed in adrenal glomerulosa that opens at hyperpolarized membrane potentials. Channel opening depolarizes glomerulosa cells and induces expression of aldosterone synthase, the rate-limiting enzyme for aldosterone biosynthesis. Mutant channels show gain of function, with higher open probabilities at the glomerulosa resting potential. These findings for the first time demonstrate a role of anion channels in glomerulosa membrane potential determination, aldosterone production and hypertension. They establish the cause of a substantial fraction of early-onset primary aldosteronism.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Fig. 1: Kindreds with hypertension and primary aldosteronism with CLCN2 mutations.
Fig. 2: Expression of ClC-2 in human adrenal gland.
Fig. 3: CLCN2 mutations increase excitatory anion efflux by modifying the voltage dependence of channel opening.
Fig. 4: ClC-2 increases aldosterone synthase expression in H295R cells.

Similar content being viewed by others

References

  1. Monticone, S. et al. Prevalence and clinical manifestations of primary aldosteronism encountered in primary care practice. J. Am. Coll. Cardiol. 69, 1811–1820 (2017).

    Article  PubMed  Google Scholar 

  2. Stowasser, M. et al. Familial hyperaldosteronism type II: five families with a new variety of primary aldosteronism. Clin. Exp. Pharmacol. Physiol. 19, 319–322 (1992).

    Article  CAS  PubMed  Google Scholar 

  3. NCD Risk Factor Collaboration (NCD-RisC). Worldwide trends in blood pressure from 1975 to 2015: a pooled analysis of 1479 population-based measurement studies with 19.1 million participants. Lancet 389, 37–55 (2017).

    Article  Google Scholar 

  4. Lim, S. S. et al. A comparative risk assessment of burden of disease and injury attributable to 67 risk factors and risk factor clusters in 21 regions, 1990–2010: a systematic analysis for the Global Burden of Disease Study 2010. Lancet 380, 2224–2260 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  5. Funder, J. W. et al. The management of primary aldosteronism: case detection, diagnosis, and treatment: an Endocrine Society Clinical Practice Guideline. J. Clin. Endocrinol. Metab. 101, 1889–1916 (2016).

    Article  CAS  PubMed  Google Scholar 

  6. Choi, M. et al. K+ channel mutations in adrenal aldosterone-producing adenomas and hereditary hypertension. Science 331, 768–772 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Scholl, U. I. et al. Somatic and germline CACNA1D calcium channel mutations in aldosterone-producing adenomas and primary aldosteronism. Nat. Genet. 45, 1050–1054 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Azizan, E. A. et al. Somatic mutations in ATP1A1 and CACNA1D underlie a common subtype of adrenal hypertension. Nat. Genet. 45, 1055–1060 (2013).

    Article  CAS  PubMed  Google Scholar 

  9. Beuschlein, F. et al. Somatic mutations in ATP1A1 and ATP2B3 lead to aldosterone-producing adenomas and secondary hypertension. Nat. Genet. 45, 440–444 (2013).

    Article  CAS  PubMed  Google Scholar 

  10. Fernandes-Rosa, F. L. et al. Genetic spectrum and clinical correlates of somatic mutations in aldosterone-producing adenoma. Hypertension 64, 354–361 (2014).

    Article  CAS  PubMed  Google Scholar 

  11. Lifton, R. P. et al. A chimaeric 11β-hydroxylase/aldosterone synthase gene causes glucocorticoid-remediable aldosteronism and human hypertension. Nature 355, 262–265 (1992).

    Article  CAS  PubMed  Google Scholar 

  12. Scholl, U. I. et al. Hypertension with or without adrenal hyperplasia due to different inherited mutations in the potassium channel KCNJ5. Proc. Natl. Acad. Sci. USA 109, 2533–2538 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Scholl, U. I. et al. Recurrent gain of function mutation in calcium channel CACNA1H causes early-onset hypertension with primary aldosteronism. eLife 4, e06315 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  14. Korah, H. E. & Scholl, U. I. An update on familial hyperaldosteronism. Horm. Metab. Res. 47, 941–946 (2015).

    Article  CAS  PubMed  Google Scholar 

  15. Spät, A. & Hunyady, L. Control of aldosterone secretion: a model for convergence in cellular signaling pathways. Physiol. Rev. 84, 489–539 (2004).

    Article  PubMed  Google Scholar 

  16. Carss, K. J., Stowasser, M., Gordon, R. D. & O’Shaughnessy, K. M. Further study of chromosome 7p22 to identify the molecular basis of familial hyperaldosteronism type II. J. Hum. Hypertens. 25, 560–564 (2011).

    Article  CAS  PubMed  Google Scholar 

  17. Dong, C. et al. Comparison and integration of deleteriousness prediction methods for nonsynonymous SNVs in whole exome sequencing studies. Hum. Mol. Genet. 24, 2125–2137 (2015).

    Article  CAS  PubMed  Google Scholar 

  18. Genin, E., Tullio-Pelet, A., Begeot, F., Lyonnet, S. & Abel, L. Estimating the age of rare disease mutations: the example of Triple-A syndrome. J. Med. Genet. 41, 445–449 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Dogan, R. I., Getoor, L., Wilbur, W. J. & Mount, S. M. SplicePort—an interactive splice-site analysis tool. Nucleic Acids Res. 35, W285–W291 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  20. Thiemann, A., Gründer, S., Pusch, M. & Jentsch, T. J. A chloride channel widely expressed in epithelial and non-epithelial cells. Nature 356, 57–60 (1992).

    Article  CAS  PubMed  Google Scholar 

  21. Untiet, V. et al. Glutamate transporter–associated anion channels adjust intracellular chloride concentrations during glial maturation. Glia 65, 388–400 (2017).

    Article  PubMed  Google Scholar 

  22. Hu, C., Rusin, C. G., Tan, Z., Guagliardo, N. A. & Barrett, P. Q. Zona glomerulosa cells of the mouse adrenal cortex are intrinsic electrical oscillators. J. Clin. Invest. 122, 2046–2053 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Jeworutzki, E. et al. GlialCAM, a protein defective in a leukodystrophy, serves as a ClC-2 Cl channel auxiliary subunit. Neuron 73, 951–961 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Stölting, G. et al. Regulation of ClC-2 gating by intracellular ATP. Pflugers Arch. 465, 1423–1437 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  25. Niemeyer, M. I., Cid, L. P., Zúñiga, L., Catalán, M. & Sepúlveda, F. V. A conserved pore-lining glutamate as a voltage- and chloride-dependent gate in the ClC-2 chloride channel. J. Physiol. 553, 873–879 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Jentsch, T. J. Discovery of CLC transport proteins: cloning, structure, function and pathophysiology. J. Physiol. 593, 4091–4109 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Rainey, W. E., Bird, I. M. & Mason, J. I. The NCI-H295 cell line: a pluripotent model for human adrenocortical studies. Mol. Cell. Endocrinol. 100, 45–50 (1994).

    Article  CAS  PubMed  Google Scholar 

  28. Bassett, M. H., Suzuki, T., Sasano, H., White, P. C. & Rainey, W. E. The orphan nuclear receptors NURR1 and NGFIB regulate adrenal aldosterone production. Mol. Endocrinol. 18, 279–290 (2004).

    Article  CAS  PubMed  Google Scholar 

  29. Romero, D. G. et al. Regulators of G-protein signaling 4 in adrenal gland: localization, regulation, and role in aldosterone secretion. J. Endocrinol. 194, 429–440 (2007).

    Article  CAS  PubMed  Google Scholar 

  30. Garcia-Olivares, J. et al. Gating of human ClC-2 chloride channels and regulation by carboxy-terminal domains. J. Physiol. 586, 5325–5336 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Tauber, P. et al. Cellular pathophysiology of an adrenal adenoma–associated mutant of the plasma membrane Ca2+-ATPase ATP2B3. Endocrinology 157, 2489–2499 (2016).

    Article  CAS  PubMed  Google Scholar 

  32. Lafferty, A. R. et al. A novel genetic locus for low renin hypertension: familial hyperaldosteronism type II maps to chromosome 7 (7p22). J. Med. Genet. 37, 831–835 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Depienne, C. et al. Brain white matter oedema due to ClC-2 chloride channel deficiency: an observational analytical study. Lancet Neurol. 12, 659–668 (2013).

    Article  CAS  PubMed  Google Scholar 

  34. Blanz, J. et al. Leukoencephalopathy upon disruption of the chloride channel ClC-2. J. Neurosci. 27, 6581–6589 (2007).

    Article  CAS  PubMed  Google Scholar 

  35. Stowasser, M. et al. Clinical, biochemical and genetic approaches to the detection of familial hyperaldosteronism type I. J. Hypertens. 13, 1610–1613 (1995).

    CAS  PubMed  Google Scholar 

  36. Chorvátová, A., Gendron, L., Bilodeau, L., Gallo-Payet, N. & Payet, M. D. A Ras-dependent chloride current activated by adrenocorticotropin in rat adrenal zona glomerulosa cells. Endocrinology 141, 684–692 (2000).

    Article  PubMed  Google Scholar 

  37. Torpy, D. J. et al. Familial hyperaldosteronism type II: description of a large kindred and exclusion of the aldosterone synthase (CYP11B2) gene. J. Clin. Endocrinol. Metab. 83, 3214–3218 (1998).

    CAS  PubMed  Google Scholar 

  38. Krumm, N. et al. Excess of rare, inherited truncating mutations in autism. Nat. Genet. 47, 582–588 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Verkman, A. S. Development and biological applications of chloride-sensitive fluorescent indicators. Am. J. Physiol. 259, C375–C388 (1990).

    Article  CAS  PubMed  Google Scholar 

  40. Kaneko, H., Putzier, I., Frings, S., Kaupp, U. B. & Gensch, T. Chloride accumulation in mammalian olfactory sensory neurons. J. Neurosci. 24, 7931–7938 (2004).

    Article  CAS  PubMed  Google Scholar 

  41. Bevensee, M. O., Apkon, M. & Boron, W. F. Intracellular pH regulation in cultured astrocytes from rat hippocampus. II. Electrogenic Na/HCO3 cotransport. J. Gen. Physiol. 110, 467–483 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Chao, A. C., Dix, J. A., Sellers, M. C. & Verkman, A. S. Fluorescence measurement of chloride transport in monolayer cultured cells. Mechanisms of chloride transport in fibroblasts. Biophys. J. 56, 1071–1081 (1989).

    CAS  PubMed  Google Scholar 

  43. Accardi, A. & Pusch, M. Fast and slow gating relaxations in the muscle chloride channel CLC-1. J. Gen. Physiol. 116, 433–444 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. de Santiago, J. A., Nehrke, K. & Arreola, J. Quantitative analysis of the voltage-dependent gating of mouse parotid ClC-2 chloride channel. J. Gen. Physiol. 126, 591–603 (2005).

    Article  PubMed  PubMed Central  Google Scholar 

  45. Rhee, J. S., Ebihara, S. & Akaike, N. Gramicidin perforated patch-clamp technique reveals glycine-gated outward chloride current in dissociated nucleus solitarii neurons of the rat. J. Neurophysiol. 72, 1103–1108 (1994).

    Article  CAS  PubMed  Google Scholar 

  46. Kim, D. et al. TopHat2: accurate alignment of transcriptomes in the presence of insertions, deletions and gene fusions. Genome Biol. 14, R36 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  47. Trapnell, C. et al. Differential analysis of gene regulation at transcript resolution with RNA-seq. Nat. Biotechnol. 31, 46–53 (2013).

    Article  CAS  PubMed  Google Scholar 

  48. Samocha, K. E. et al. A framework for the interpretation of de novo mutation in human disease. Nat. Genet. 46, 944–950 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Zou, J. et al. Quantifying unobserved protein-coding variants in human populations provides a roadmap for large-scale sequencing projects. Nat. Commun. 7, 13293 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Lek, M. et al. Analysis of protein-coding genetic variation in 60,706 humans. Nature 536, 285–291 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank our patients and their families for their invaluable contributions, the Yale Center for Genome Analysis for next-generation sequencing, the Center for Advanced Imaging (CAi) at Heinrich Heine University for providing a confocal microscope, S. Weidtkamp-Peters and S. Hänsch for technical assistance, J. Zhang for helpful discussions, E. Seidel, N. Erlenhardt and N. Klöcker for providing immunoprecipitation protocols and helpful discussions, C. Gomez-Sanchez (University of Mississippi) for providing plasmids, W. Rainey (University of Michigan) for providing HAC15 cells and M. Haase (Heinrich Heine University Düsseldorf) for providing H295R cells. Computational support and infrastructure were in part provided by the Centre for Information and Media Technology (Düsseldorf). This study was supported in part by the Ministerium für Kultur und Wissenschaft des Landes Nordrhein-Westfalen (Rückkehrprogramm and Junges Kolleg) and the Deutsche Forschungsgemeinschaft (SCHO 1386/2-1) (all to U.I.S.) and the NIH Center for Mendelian Genomics (5U54HG006504), NIH P01DK17433 and the Howard Hughes Medical Institute (all to R.P.L.).

Author information

Authors and Affiliations

Authors

Contributions

U.I.S., G.S., C.F. and R.P.L. designed the study. M.S. and R.G. recruited and characterized family 3. A.M.O., C.G., D.M., R.W.L., D.P.J., G.C., P.G., E.L., C.N.-W. and R.P.L. ascertained and recruited probands with early-onset primary aldosteronism. A.A.V., E.L. and U.I.S. recruited additional members of selected families. C.N.-W., S.X. and A.W. prepared DNA samples. U.I.S., A.A.V., S.C.J., T.Y., M.C. and R.P.L. analyzed exome sequencing results. U.I.S. identified the disease-associated gene. C.N.-W., A.T. and A.A.V. performed and analyzed the results of targeted DNA sequencing. J.S. performed immunohistochemistry, immunoprecipitation and real-time PCR. J.S. and A.T. made constructs and generated stable cell lines. J.S. and J.C. prepared samples for and analyzed the results of RNA sequencing. J.S. and U.I.S. performed and analyzed the results of splicing assays and confocal microscopy. G.S., H.T., V.U. and C.F. performed and analyzed the results of FLIM and electrophysiology. M.K. and P.M. performed and analyzed the results of mass spectrometry. L.C.R. read and revised the manuscript. U.I.S. wrote the initial draft of the manuscript, with contributions and/or revisions from all authors.

Corresponding author

Correspondence to Ute I. Scholl.

Ethics declarations

Competing interests

Heinrich Heine University Düsseldorf has filed a patent application: EP17209972, Diagnosis and Therapy of Primary Aldosteronism.

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 Sanger sequences.

a, Sanger sequences of kindreds with CLCN2 variants. M/+ denotes the indicated novel CLCN2 variant in the heterozygous state, and +/+ denotes homozygous wild-type sequence. Mutant bases are indicated by a red frame, and encoded amino acid sequences are shown above. b, Splicing assay for the p.Lys362del mutation. Exons 9–11 of CLCN2 were cloned, and HEK293T cells were transfected with WT DNA and DNA carrying the variant found in kindred 1492. RNA was isolated and cDNA was transcribed. Shown are Sanger sequences of PCRs covering the exon 10/11 splice site affected by the variant. A new donor site on exon 10 is used, resulting in deletion of the last 3 bp of exon 10 (red rectangle in WT sequence, red line in mutant sequence). These results were consistent with in silico prediction.

Supplementary Figure 2 Immunohistochemistry.

a, Immunohistochemistry as in Fig. 2 was performed on the adrenal gland of a second human subject (one of two technical replicates shown). Peptide control, antibody was preincubated with immunogenic peptide. Scale bar, 100 µm. C, capsule; G, glomerulosa; F, fasciculata. b, Immunofluorescent staining of mouse adrenal gland, with DAB2 as a marker of the zona glomerulosa. Scale bar, 10 µm.

Supplementary Figure 3 Gating analysis, time constants of activation/deactivation and current density–voltage plots of WT and mutant ClC-2 channels.

a, Whole-cell patch–clamp recordings of representative ClC-2MUT and voltage protocol (150 mM Cl outside, 75 mM Cl inside; see the Methods for solutions) for the indicated mutations are shown. b, Time constants for WT and mutant ClC-2 channels (see the Methods for details). Mean values ± 95% confidence intervals are shown (WT, p.Arg172Gln and p.Ser865Arg are reproduced from Fig. 3; WT, n = 11; p.Met22Lys, n = 12; p.Tyr26Asn, n = 13; p.Arg172Gln, n = 13; p.Lys362del, n = 12; p.Ser865Arg, n = 11). c, Plot of steady-state current densities. Values are shown as means ± 95% confidence intervals; WT, n = 11; p.Met22Lys, n = 12; p.Tyr26Asn, n = 13; p.Arg172Gln, n = 13; p.Lys362del, n = 12; p.Ser865Arg, n = 11. Median values at –80 mV and the results of Kruskal–Wallis one-way ANOVA (H = 14.823, d.f. = 5) followed by Dunn’s method when appropriate are as follows: WT: –5.75; p.Met22Lys: –10.28, n.s. vs. WT; p.Tyr26Asn: –13.66, P < 0.05 vs. WT; p.Arg172Gln: –9.13, n.s. vs. WT; p.Lys362del: –13.69, P < 0.05 vs. WT; p.Ser865Arg, –12.58, P < 0.05 vs. WT. df, Total (d), protopore (e) and common gate (f) open probabilities were determined as described (see the Methods for details; WT, p.Arg172Gln and p.Ser865Arg total and common open probabilities are reproduced from Fig. 3; WT, n = 11; p.Met22Lys, n = 12; p.Tyr26Asn, n = 13; p.Arg172Gln, n =13; p.Lys362del, n = 12; p.Ser865Arg, n = 12). Data points were fit using a Boltzmann function (bold line), and 95% confidence intervals were determined using bootstrap sampling. The overall shift in activation of mutant ClC-2 channels apart from p.Ser865Arg mostly results from a higher common gate open probability, whereas the fast gate open probability of p.Ser865Arg is shifted to more positive potentials. g, Individual data and stable cell line clones used in the analysis shown in Fig. 3 and this figure.

Supplementary Figure 4 Mass spectrometry and confocal microscopy.

a,b, LC–MS/MS-based identification of the phosphorylation site of ClC-2 at Ser865. ClC-2 was identified with 43 peptides and sequence coverage of 52.8% Shown are annotated MS/MS spectra of the unmodified (a) and phosphorylated (b) peptide DSATSSSDTETTEVHALWGPHSR corresponding to amino acids 859–881 of ClC-2. The x axis shows m/z, the left y axis shows relative abundance and the right y axis shows absolute signal intensity. Unmodified peptide: 16 MS/MS counts, posterior error probability (PEP) 1.21 × 10–35, score 160.47; Ser865-phosphorylated peptide: 11 MS/MS counts, PEP 1.61 × 10–36, score 166.49. The phosphorylation site at Ser865 was fully localized, with a localization probability of 0.813811 (mass error of 0.28027 p.p.m.). One of two independent preparations (Methods) is shown; similar results were obtained for another two independent preparations using a different solvent. c,d, Live cell confocal microscopy of YFP-tagged WT and mutant ClC-2 and the surface membrane marker CFP-mem in H295R cells. c, The respective variant is noted on the left. Left, confocal image using the YFP channel; middle, CFP channel; right, overlay. Scale bars, 20 µm. Two independent clones of each plasmid were assessed, and representative images are shown. d, Correlation R between YFP and CFP fluorescence for each construct. WT, 0.29 ± 0.03 (n = 71); p.Met22Lys, 0.15 ± 0.02 (n = 81; P = 0.0001 vs. WT); p.Tyr26Asn, 0.25 ± 0.03 (n = 53; P > 0.9999 vs. WT); p.Arg172Gln, 0.24 ± 0.03 (n = 56; P = 0.5862 vs. WT); p.Lys362del, 0.29 ± 0.03 (n = 65; P > 0.9999 vs. WT); p.Ser865Arg, 0.30 ± 0.02 (n = 59; P > 0.9999 vs. WT); all mean ± s.e.m., Kruskal–Wallis test; Dunn’s multiple-comparisons test; Kruskal–Wallis statistic 28.47, d.f. = 5. Box, interquartile range; whiskers, 1.5 times the interquartile range; line, median; ***P < 0.001; n.s., not significant.

Supplementary Figure 5 CYP11B2 expression in H295R cells after transfection of non-functional CLCN2.

Relative expression levels of CYP11B2 determined by real-time PCR in H295R cells transfected with empty vector control, wild-type CLCN2 (blue) or two constructs with C-terminal deletions that affect ion channel function (yellow, corresponding deletions noted below the graph)30. Values were normalized to empty vector control. CYP11B2 expression is significantly lower after transfection of non-functional channels than after transfection of the wild-type channel (mean ± s.e.m.; WT, 16.35 ± 2.72; p.Arg751Ter, 1.46 ± 0.18, P = 0.0089 vs. WT; p.His573_Leu636del, 2.38 ± 0.50, P = 0.0077 vs. WT; **P < 0.01). n = 5, one-way ANOVA for all constructs, Dunnett’s multiple-comparisons test, F = 31.52, d.f. = 4. Box, interquartile range; whiskers, 1.5 times the interquartile range; line, median.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–5, Supplementary Tables 1–7 and 10, and Supplementary Note

Life Sciences Reporting Summary

Supplementary Table 8

Details of statistical analysis shown in Figs. 3 and 4, and Supplementary Fig. 3

Supplementary Table 9

Whole-transcriptome analysis of CLCN2-transfected H295R cells

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Scholl, U.I., Stölting, G., Schewe, J. et al. CLCN2 chloride channel mutations in familial hyperaldosteronism type II. Nat Genet 50, 349–354 (2018). https://doi.org/10.1038/s41588-018-0048-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-018-0048-5

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research