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:

Mutations in KCNH1 and ATP6V1B2 cause Zimmermann-Laband syndrome

Abstract

Zimmermann-Laband syndrome (ZLS) is a developmental disorder characterized by facial dysmorphism with gingival enlargement, intellectual disability, hypoplasia or aplasia of nails and terminal phalanges, and hypertrichosis1,2,3,4. We report that heterozygous missense mutations in KCNH1 account for a considerable proportion of ZLS. KCNH1 encodes the voltage-gated K+ channel Eag1 (Kv10.1). Patch-clamp recordings showed strong negative shifts in voltage-dependent activation for all but one KCNH1 channel mutant (Gly469Arg). Coexpression of Gly469Arg with wild-type KCNH1 resulted in heterotetrameric channels with reduced conductance at positive potentials but pronounced conductance at negative potentials. These data support a gain-of-function effect for all ZLS-associated KCNH1 mutants. We also identified a recurrent de novo missense change in ATP6V1B2, encoding the B2 subunit of the multimeric vacuolar H+ ATPase, in two individuals with ZLS. Structural analysis predicts a perturbing effect of the mutation on complex assembly. Our findings demonstrate that KCNH1 mutations cause ZLS and document genetic heterogeneity for this disorder.

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

Figure 1: Heterozygous missense mutations in KCNH1 and ATP6V1B2 in individuals with ZLS.
Figure 2: Clinical features of ZLS-affected individuals with a KCNH1 or ATP6V1B2 mutation.
Figure 3: Structural impact of ZLS-associated KCNH1 and ATP6V1B2 mutations.
Figure 4: Voltage-dependent activation of human wild-type (WT) and mutant KCNH1 channels expressed in CHO cells.
Figure 5: Voltage-dependent activation of KCNH1 K+ currents resulting from coexpression of wild-type KCNH1 with mutant KCNH1 channel subunits in CHO cells using a 1:1 cDNA ratio.

Similar content being viewed by others

Accession codes

Primary accessions

NCBI Reference Sequence

Referenced accessions

NCBI Reference Sequence

Protein Data Bank

References

  1. Laband, P.F., Habib, G. & Humphreys, G.S. Hereditary gingival fibromatosis. Report of an affected family with associated splenomegaly and skeletal and soft-tissue abnormalities. Oral Surg. Oral Med. Oral Pathol. 17, 339–351 (1964).

    Article  CAS  PubMed  Google Scholar 

  2. Alavandar, G. Elephantiasis gingivae. Report of an affected family with associated hepatomegaly, soft tissue and skeletal abnormalities. J. All India Dent. Assoc. 37, 349–353 (1965).

    CAS  PubMed  Google Scholar 

  3. Chacon-Camacho, O.F., Vazquez, J. & Zenteno, J.C. Expanding the phenotype of gingival fibromatosis–mental retardation–hypertrichosis (Zimmermann-Laband) syndrome. Am. J. Med. Genet. 155A, 1716–1720 (2011).

    Article  PubMed  Google Scholar 

  4. Dávalos, I.P. et al. Zimmermann-Laband syndrome: further clinical delineation. Genet. Couns. 16, 283–290 (2005).

    PubMed  Google Scholar 

  5. Castori, M. et al. Clinical and genetic study of two patients with Zimmermann-Laband syndrome and literature review. Eur. J. Med. Genet. 56, 570–576 (2013).

    Article  PubMed  Google Scholar 

  6. Abo-Dalo, B. et al. Extensive molecular genetic analysis of the 3p14.3 region in patients with Zimmermann-Laband syndrome. Am. J. Med. Genet. 143A, 2668–2674 (2007).

    Article  CAS  PubMed  Google Scholar 

  7. Kim, H.G. et al. Candidate loci for Zimmermann-Laband syndrome at 3p14.3. Am. J. Med. Genet. 143A, 107–111 (2007).

    Article  PubMed  Google Scholar 

  8. Stefanova, M. et al. Zimmermann-Laband syndrome associated with a balanced reciprocal translocation t(3;8)(p21.2;q24.3) in mother and daughter: molecular cytogenetic characterization of the breakpoint regions. Am. J. Med. Genet. A. 117A, 289–294 (2003).

    Article  PubMed  Google Scholar 

  9. Abo-Dalo, B. et al. No mutation in genes of the WNT signaling pathway in patients with Zimmermann-Laband syndrome. Clin. Dysmorphol. 17, 181–185 (2008).

    Article  PubMed  Google Scholar 

  10. Petrovski, S. et al. Genic intolerance to functional variation and the interpretation of personal genomes. PLoS Genet. 9, e1003709 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. 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 

  12. Bauer, C.K. & Schwarz, J.R. Physiology of EAG K+ channels. J. Membr. Biol. 182, 1–15 (2001).

    Article  CAS  PubMed  Google Scholar 

  13. Garg, V., Sachse, F.B. & Sanguinetti, M.C. Tuning of EAG K+ channel inactivation: molecular determinants of amplification by mutations and a small molecule. J. Gen. Physiol. 140, 307–324 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Thouta, S. et al. Proline scan of the HERG channel S6 helix reveals the location of the intracellular pore gate. Biophys. J. 106, 1057–1069 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Wynia-Smith, S.L., Gillian-Daniel, A.L., Satyshur, K.A. & Robertson, G.A. hERG gating microdomains defined by S6 mutagenesis and molecular modeling. J. Gen. Physiol. 132, 507–520 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Veeramah, K.R. et al. Exome sequencing reveals new causal mutations in children with epileptic encephalopathies. Epilepsia 54, 1270–1281 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Yang, Y. et al. Multistate structural modeling and voltage-clamp analysis of epilepsy/autism mutation Kv10.2-R327H demonstrate the role of this residue in stabilizing the channel closed state. J. Neurosci. 33, 16586–16593 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Forgac, M. Vacuolar ATPases: rotary proton pumps in physiology and pathophysiology. Nat. Rev. Mol. Cell Biol. 8, 917–929 (2007).

    Article  CAS  PubMed  Google Scholar 

  19. Marshansky, V., Rubinstein, J.L. & Gruber, G. Eukaryotic V-ATPase: novel structural findings and functional insights. Biochim. Biophys. Acta 1837, 857–879 (2014).

    Article  CAS  PubMed  Google Scholar 

  20. Arai, S. et al. Rotation mechanism of Enterococcus hirae V1-ATPase based on asymmetric crystal structures. Nature 493, 703–707 (2013).

    Article  CAS  PubMed  Google Scholar 

  21. Fuster, D.G., Zhang, J., Xie, X.S. & Moe, O.W. The vacuolar-ATPase B1 subunit in distal tubular acidosis: novel mutations and mechanisms for dysfunction. Kidney Int. 73, 1151–1158 (2008).

    Article  CAS  PubMed  Google Scholar 

  22. Martin, S. et al. Eag1 potassium channel immunohistochemistry in the CNS of adult rat and selected regions of human brain. Neuroscience 155, 833–844 (2008).

    Article  CAS  PubMed  Google Scholar 

  23. Hemmerlein, B. et al. Overexpression of Eag1 potassium channels in clinical tumours. Mol. Cancer 5, 41 (2006).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  24. Pardo, L.A. & Stühmer, W. The roles of K+ channels in cancer. Nat. Rev. Cancer 14, 39–48 (2014).

    Article  CAS  PubMed  Google Scholar 

  25. Ouadid-Ahidouch, H. & Ahidouch, A. K+ channel expression in human breast cancer cells: involvement in cell cycle regulation and carcinogenesis. J. Membr. Biol. 221, 1–6 (2008).

    Article  CAS  PubMed  Google Scholar 

  26. Weber, C. et al. Silencing the activity and proliferative properties of the human EagI potassium channel by RNA interference. J. Biol. Chem. 281, 13030–13037 (2006).

    Article  CAS  PubMed  Google Scholar 

  27. Zhang, Y.Y. et al. BKCa and hEag1 channels regulate cell proliferation and differentiation in human bone marrow–derived mesenchymal stem cells. J. Cell. Physiol. 229, 202–212 (2014).

    Article  CAS  PubMed  Google Scholar 

  28. Ufartes, R. et al. Behavioural and functional characterization of Kv10.1 (Eag1) knockout mice. Hum. Mol. Genet. 22, 2247–2262 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Harakalova, M. et al. Dominant missense mutations in ABCC9 cause Cantú syndrome. Nat. Genet. 44, 793–796 (2012).

    Article  CAS  PubMed  Google Scholar 

  30. van Bon, B.W. et al. Cantú syndrome is caused by mutations in ABCC9. Am. J. Hum. Genet. 90, 1094–1101 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Brownstein, C.A. et al. Mutation of KCNJ8 in a patient with Cantú syndrome with unique vascular abnormalities—support for the role of KATP channels in this condition. Eur. J. Med. Genet. 56, 678–682 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  32. Cooper, P.E. et al. Cantú syndrome resulting from activating mutation in the KCNJ8 gene. Hum. Mutat. 35, 809–813 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Lipscombe, D., Helton, T.D. & Xu, W. L-type calcium channels: the low down. J. Neurophysiol. 92, 2633–2641 (2004).

    Article  CAS  PubMed  Google Scholar 

  34. Corrêa, J.D. et al. Phenytoin-induced gingival overgrowth: a review of the molecular, immune, and inflammatory features. ISRN Dent. 2011, 497850 (2011).

    PubMed  PubMed Central  Google Scholar 

  35. Mantegazza, M. et al. Voltage-gated sodium channels as therapeutic targets in epilepsy and other neurological disorders. Lancet Neurol. 9, 413–424 (2010).

    Article  CAS  PubMed  Google Scholar 

  36. Wynn, R.L. Calcium channel blockers and gingival hyperplasia—an update. Gen. Dent. 57, 105–107 (2009).

    PubMed  Google Scholar 

  37. Jacquinet, A. et al. Temple-Baraitser syndrome: a rare and possibly unrecognized condition. Am. J. Med. Genet. 152A, 2322–2326 (2010).

    Article  PubMed  Google Scholar 

  38. White, S.M. & Fahey, M. Report of a further family with dominant deafness-onychodystrophy (DDOD) syndrome. Am. J. Med. Genet. 155A, 2512–2515 (2011).

    Article  PubMed  Google Scholar 

  39. Campeau, P.M. & Hennekam, R.C. DOORS syndrome: phenotype, genotype and comparison with Coffin-Siris syndrome. Am. J. Med. Genet. 166C, 327–332 (2014).

    Article  PubMed  CAS  Google Scholar 

  40. Simons, C. et al. Mutations in the voltage-gated potassium channel gene KCNH1 cause Temple-Baraitser syndrome and epilepsy. Nat. Genet. 47, 73–77 (2015).

    Article  CAS  PubMed  Google Scholar 

  41. Nelson, R.D. et al. Selectively amplified expression of an isoform of the vacuolar H+-ATPase 56-kilodalton subunit in renal intercalated cells. Proc. Natl. Acad. Sci. USA 89, 3541–3545 (1992).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Puopolo, K. et al. Differential expression of the “B” subunit of the vacuolar H+-ATPase in bovine tissues. J. Biol. Chem. 267, 3696–3706 (1992).

    Article  CAS  PubMed  Google Scholar 

  43. van Hille, B. et al. Heterogeneity of vacuolar H+-ATPase: differential expression of two human subunit B isoforms. Biochem. J. 303, 191–198 (1994).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Yuan, Y. et al. De novo mutation in ATP6V1B2 impairs lysosome acidification and causes dominant deafness-onychodystrophy syndrome. Cell Res. 24, 1370–1373 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Mortensen, L.S. et al. KV10.1 opposes activity-dependent increase in Ca2+ influx into the presynaptic terminal of the parallel fibre–Purkinje cell synapse. J. Physiol. (Lond.) 593, 181–196 (2015).

    Article  CAS  Google Scholar 

  46. Poëa-Guyon, S. et al. The V-ATPase membrane domain is a sensor of granular pH that controls the exocytotic machinery. J. Cell Biol. 203, 283–298 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  47. Mindell, J.A. Lysosomal acidification mechanisms. Annu. Rev. Physiol. 74, 69–86 (2012).

    Article  CAS  PubMed  Google Scholar 

  48. Steinberg, B.E. et al. A cation counterflux supports lysosomal acidification. J. Cell Biol. 189, 1171–1186 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Van Dyke, R.W. Acidification of rat liver lysosomes: quantitation and comparison with endosomes. Am. J. Physiol. 265, C901–C917 (1993).

    Article  CAS  PubMed  Google Scholar 

  50. Kohl, T., Lörinczi, E., Pardo, L.A. & Stühmer, W. Rapid internalization of the oncogenic K+ channel KV10.1. PLoS ONE 6, e26329 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Ninkovic, M. et al. Physical and functional interaction of KV10.1 with Rabaptin-5 impacts ion channel trafficking. FEBS Lett. 586, 3077–3084 (2012).

    Article  CAS  PubMed  Google Scholar 

  52. Li, H. & Durbin, R. Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics 25, 1754–1760 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. McKenna, A. et al. The Genome Analysis Toolkit: a MapReduce framework for analyzing next-generation DNA sequencing data. Genome Res. 20, 1297–1303 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Cingolani, P. et al. A program for annotating and predicting the effects of single nucleotide polymorphisms, SnpEff: SNPs in the genome of Drosophila melanogaster strain w1118; iso-2; iso-3. Fly (Austin) 6, 80–92 (2012).

    Article  CAS  Google Scholar 

  55. Liu, X., Jian, X. & Boerwinkle, E. dbNSFP v2.0: a database of human non-synonymous SNVs and their functional predictions and annotations. Hum. Mutat. 34, E2393–E2402 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Biasini, M. et al. SWISS-MODEL: modelling protein tertiary and quaternary structure using evolutionary information. Nucleic Acids Res. 42, W252–W258 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Roy, A., Kucukural, A. & Zhang, Y. I-TASSER: a unified platform for automated protein structure and function prediction. Nat. Protoc. 5, 725–738 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Pettersen, E.F. et al. UCSF Chimera—a visualization system for exploratory research and analysis. J. Comput. Chem. 25, 1605–1612 (2004).

    CAS  PubMed  Google Scholar 

  60. Schuster, A.M., Glassmeier, G. & Bauer, C.K. Strong activation of ether-à-go-go–related gene 1 K+ channel isoforms by NS1643 in human embryonic kidney 293 and Chinese hamster ovary cells. Mol. Pharmacol. 80, 930–942 (2011).

    Article  CAS  PubMed  Google Scholar 

  61. Clay, J.R. Determining K+ channel activation curves from K+ channel currents often requires the Goldman-Hodgkin-Katz equation. Front. Cell. Neurosci. 3, 20 (2009).

    PubMed  PubMed Central  Google Scholar 

  62. Schönherr, R. et al. Individual subunits contribute independently to slow gating of bovine EAG potassium channels. J. Biol. Chem. 274, 5362–5369 (1999).

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

We are grateful to the patients and their families who contributed to this study. We thank I. Jantke, S. Cecchetti and S. Venanzi for skillful technical assistance, T. Kock for site-directed mutagenesis, A. Hasse for CHO cell transfection and injection, R. Bähring, J.M. Schröder and E. Neumann for help with the oocyte experiments, P. Meinecke for discussing clinical phenotypes and A. Podolska for help with ATP6V1B2 sequencing. G.B., L.S. and M.T. acknowledge CINECA for computational resources (whole-exome sequencing data and structural analyses). The KCNH1/heag1 clone was kindly provided by S.H. Heinemann (Friedrich Schiller University Jena). This work was supported by grants from the Deutsche Forschungsgemeinschaft (KO 4576/1-1 to F.K. and KU 1240/5-1 to K.K.), Istituto Superiore di Sanità (Ricerca Corrente 2013 to M.T.), Ministero della Salute (Ricerca Finalizzata RF-2010-2312766 to B.D.) and Ospedale Pediatrico Bambino Gesù (Gene-Rare to B.D.).

Author information

Authors and Affiliations

Authors

Contributions

F.K. performed whole-exome sequencing data analysis and validation, molecular screening and genotyping and wrote the manuscript. V.C. performed whole-exome sequencing data analysis and validation and wrote the manuscript. C.K.B. contributed the electrophysiological studies and wrote the manuscript. L.S. and G.B. performed the homology modeling and structural analysis. A.C. and M.A. contributed to the whole-exome sequencing data processing and analysis. E.F., S.P., M.L.D. and T.T.M.N. carried out the molecular screening and/or genotyping. P.G., G.C.K., V.L., D.M., L.D.V.N., P.T., S.M.W., B.D. and A.P. recruited and clinically characterized the study subjects and collected the biological samples. P.M.C. performed whole-exome sequencing data analysis, validation and genotyping. M.T. and K.K. conceived the project, analyzed and interpreted the data, and wrote the manuscript. All authors contributed to the final manuscript.

Corresponding authors

Correspondence to Marco Tartaglia or Kerstin Kutsche.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Sequence electropherograms of individuals with KCNH1 or ATP6V1B2 mutation.

Sequence electropherograms showing the de novo origin of the identified KCNH1 and ATP6V1B2 missense mutations in subjects 1–8 (upper and lower panels, indicated by red arrows). The heterozygous state of three mutations was documented in peripheral leukocytes, hair bulb and/or buccal cells of subjects 4, 5 and 7, indicating germline origin. An additional previously annotated (ExAC database) heterozygous KCNH1 variant, c.125T>C (p.Ile42Thr), was present in subject 5 and his healthy mother (indicated by blue arrows). By cloning the KCNH1 exon 7–containing amplicon of subject 2 followed by sequencing, we determined the haplotypes and found that the two identified de novo changes c.974C>A and c.1066G>C are in cis (wild-type allele and mutated KCNH1 allele in the middle panel; mutated nucleotides are framed).

Supplementary Figure 2 Multiple protein sequence alignments around the KCNH1 and ATP6V1B2 amino acid substitutions from different species.

Alignment of the regions flanking the detected missense variants in orthologous proteins, showing the evolutionary conservation of amino acids S325, G348, L352, V356, I467 and G469 in human KCNH1 (NP_002229.1) and of R485 in human ATP6V1B2 (NP_001684.2). Multiple alignments were gathered from http://www.ncbi.nlm.nih.gov/homologene/. Conserved residues have a red background, and non-conserved residues have a gray background. Amino acid sequence alignments demonstrate high (S325 and V356 in human KCNH1) or complete (G348, L352, I467 and G469 in human KCNH1 and R485 in human ATP6V1B2) evolutionary conservation of the altered residues.

Supplementary Figure 3 Voltage dependence of wild-type and mutant KCNH1 channel activation.

(a) KCNH1 channels were expressed in CHO cells, and families of wild-type (WT) and L352V current traces recorded with the depicted pulse protocols are shown. Zero current is indicated by dashed lines and arrowheads. (b) Mean (±s.e.m.) normalized instantaneous tail current amplitudes as a function of the preceding test pulse potential. Lines represent first-order Boltzmann functions fitted to the data points. The voltage dependence of channel activation was analyzed from instantaneous tail current measurements at +40 mV (for KCNH1 WT and I467V for all experiments and for G348R and S325Y/V356L for four experiments each) or at –20 mV (for L352V, G348R and S325Y/V356L). No significant differences were found between the potentials for half-maximal G348R or S325Y/V356L channel activation determined with the two different constant pulse potentials. (c) Table with means ± s.e.m. of the potential of half-maximal channel activation and the slope factor k derived from fits of first-order Boltzmann functions to the normalized instantaneous tail current amplitudes of the individual experiments. n, number of experiments; *, significantly different from WT with P < 0.05; ***, significantly different from WT with P < 0.001. Values were tested for significant differences compared to WT data with one-way ANOVA and post-hoc Bonferroni t test.

Source data

Supplementary Figure 4 Analysis of the activation and deactivation kinetics of wild-type and mutant KCNH1 channels expressed in CHO cells.

(a) The time course of channel activation was analyzed at +40 mV from experiments as shown in Figure 4 by fitting a double-exponential function to the current traces, yielding the fast and the slow time constant of current activation at +40 mV as well as the amplitudes of the two current components (Af and As). Note that the preceding holding potential was –80 mV, except for the mutant L352V, where a holding potential of –100 mV was used. Compared to wild-type KCNH1 channels, the time course of current activation was accelerated for all mutant channels. For G348R, I467V and L352V, both activation time constants were significantly decreased, and for the double mutant S325Y/V356L, the slow time constant decreased significantly in combination with a higher relative contribution of the faster activating current component. (b) Families of wild-type and G348R current traces recorded with the depicted deactivation protocol. Zero current is indicated by a dashed line. (c) The time course of KCNH1 channel deactivation was analyzed at –120 mV from experiments as shown in b. Compared to wild-type channel, the deactivation time course of all mutant channels was significantly slowed. Current decay upon hyperpolarization to –120 mV was fitted with a double-exponential function, yielding the fast and the slow time constant of current deactivation as well as the amplitudes of the two current components (Af and As). Af and As were extrapolated to segment start, because the first 0.5 to 1 ms of the pulse segment at –120 mV was not used for fitting to minimize contributions of capacitive currents. (a,c) The number of experiments is given in the lower bar plots; **, significantly different from wild-type channel with P < 0.01; ***, significantly different from wild-type channel with P < 0.001. Values were tested for significant differences compared to wild-type data with a two-tailed heteroscedastic t test and Bonferroni correction for multiple testing.

Source data

Supplementary Figure 5 Voltage dependence of S325Y and V356L KCNH1 channel activation.

(a,b) Mean values (±s.e.m.) of normalized current amplitudes (a) and whole-cell conductance (b) for S325Y (n = 14) and V356L (n = 11). Data points in a are connected by lines; the dark gray lines in b represent fits to the data points using equation (2). Corresponding data for the double mutant S325Y/V356L and for wild-type (WT) channels are shown for comparison as light gray and black lines, respectively (data from Fig. 4). Fit parameters are given in Supplementary Table 4.

Source data

Supplementary Figure 6 Expression and coexpression of wild-type (WT) and mutant (G469R) KCNH1 channels in Xenopus laevis oocytes.

(a) Families of current traces recorded with the depicted pulse protocol. For the G469R mutant, depolarizing pulses failed to induce voltage-dependent outward currents and coexpression of the mutant with wild-type channels suppressed the amplitude of KCNH1 outward currents at more positive potentials. Zero current is indicated by dashed lines and arrowheads. (b) Means (+s.e.m.) of the current amplitude recorded at the end of the test pulse to +40 mV. Absolute current values were normalized to the average current amplitude obtained for wild-type KCNH1 channels. ***, significantly different from WT with P < 0.001 (one-way ANOVA and post-hoc Bonferroni t test); n.s., not significantly different. The amount of injected cRNA is indicated, and the number of experiments is given in parentheses. Similar results were obtained in experiments using other batches of oocytes.

Source data

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–6 and Supplementary Tables 1–5. (PDF 3899 kb)

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kortüm, F., Caputo, V., Bauer, C. et al. Mutations in KCNH1 and ATP6V1B2 cause Zimmermann-Laband syndrome. Nat Genet 47, 661–667 (2015). https://doi.org/10.1038/ng.3282

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ng.3282

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