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:

PDE3A mutations cause autosomal dominant hypertension with brachydactyly

Subjects

Abstract

Cardiovascular disease is the most common cause of death worldwide, and hypertension is the major risk factor1. Mendelian hypertension elucidates mechanisms of blood pressure regulation. Here we report six missense mutations in PDE3A (encoding phosphodiesterase 3A) in six unrelated families with mendelian hypertension and brachydactyly type E (HTNB)2. The syndrome features brachydactyly type E (BDE), severe salt-independent but age-dependent hypertension, an increased fibroblast growth rate, neurovascular contact at the rostral-ventrolateral medulla, altered baroreflex blood pressure regulation and death from stroke before age 50 years when untreated3,4. In vitro analyses of mesenchymal stem cell–derived vascular smooth muscle cells (VSMCs) and chondrocytes provided insights into molecular pathogenesis. The mutations increased protein kinase A–mediated PDE3A phosphorylation and resulted in gain of function, with increased cAMP-hydrolytic activity and enhanced cell proliferation. Levels of phosphorylated VASP were diminished, and PTHrP levels were dysregulated. We suggest that the identified PDE3A mutations cause the syndrome. VSMC-expressed PDE3A deserves scrutiny as a therapeutic target for the treatment of hypertension.

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

Figure 1: Pedigrees and PDE3A missense mutations in six unrelated families with autosomal dominant hypertension with BDE (HTNB).
Figure 2: The HTNB linkage interval and expression studies.
Figure 3: Michaelis-Menten kinetics, IC50 measurements and CFSE proliferation assays.
Figure 4: PDE3A Ser428 and Ser438 phosphorylation in HeLa cells, MSCs and MSC-derived VSMCs.
Figure 5: Peptide SPOT assay of the PDE3A Thr445Asn mutant.

Similar content being viewed by others

Accession codes

Accessions

NCBI Reference Sequence

References

  1. Hunter, D.J. & Reddy, K.S. Noncommunicable diseases. N. Engl. J. Med. 369, 1336–1343 (2013).

    Article  CAS  PubMed  Google Scholar 

  2. Schuster, H. et al. Severe autosomal dominant hypertension and brachydactyly in a unique Turkish kindred maps to human chromosome 12. Nat. Genet. 13, 98–100 (1996).

    Article  CAS  PubMed  Google Scholar 

  3. Schuster, H. et al. A cross-over medication trial for patients with autosomal-dominant hypertension with brachydactyly. Kidney Int. 53, 167–172 (1998).

    Article  CAS  PubMed  Google Scholar 

  4. Naraghi, R. et al. Neurovascular compression at the ventrolateral medulla in autosomal dominant hypertension and brachydactyly. Stroke 28, 1749–1754 (1997).

    Article  CAS  PubMed  Google Scholar 

  5. Lifton, R.P. Genetic dissection of human blood pressure variation: common pathways from rare phenotypes. Harvey Lect. 100, 71–101 (2004).

    PubMed  Google Scholar 

  6. Bilginturan, N., Zileli, S., Karacadag, S. & Pirnar, T. Hereditary brachydactyly associated with hypertension. J. Med. Genet. 10, 253–259 (1973).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Toka, O. et al. Childhood hypertension in autosomal-dominant hypertension with brachydactyly. Hypertension 56, 988–994 (2010).

    Article  CAS  PubMed  Google Scholar 

  8. Bähring, S. et al. Autosomal-dominant hypertension with type E brachydactyly is caused by rearrangement on the short arm of chromosome 12. Hypertension 43, 471–476 (2004).

    Article  CAS  PubMed  Google Scholar 

  9. Bähring, S. et al. Inversion region for hypertension and brachydactyly on chromosome 12p features multiple splicing and noncoding RNA. Hypertension 51, 426–431 (2008).

    Article  CAS  PubMed  Google Scholar 

  10. Gong, M. et al. Genome-wide linkage reveals a locus for human essential (primary) hypertension on chromosome 12p. Hum. Mol. Genet. 12, 1273–1277 (2003).

    Article  CAS  PubMed  Google Scholar 

  11. Maurice, D.H. et al. Cyclic nucleotide phosphodiesterase activity, expression, and targeting in cells of the cardiovascular system. Mol. Pharmacol. 64, 533–546 (2003).

    Article  CAS  PubMed  Google Scholar 

  12. Wakabayashi, S. et al. Involvement of phosphodiesterase isozymes in osteoblastic differentiation. J. Bone Miner. Res. 17, 249–256 (2002).

    Article  CAS  PubMed  Google Scholar 

  13. Michot, C. et al. Exome sequencing identifies PDE4D mutations as another cause of acrodysostosis. Am. J. Hum. Genet. 90, 740–745 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Lee, H. et al. Exome sequencing identifies PDE4D mutations in acrodysostosis. Am. J. Hum. Genet. 90, 746–751 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Maass, P.G. et al. A misplaced lncRNA causes brachydactyly in humans. J. Clin. Invest. 122, 3990–4002 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Insel, P.A., Stengel, D., Ferry, N. & Hanoune, J. Regulation of adenylate cyclase of human platelet membranes by forskolin. J. Biol. Chem. 257, 7485–7490 (1982).

    CAS  PubMed  Google Scholar 

  17. Knowles, R.G., Palacios, M., Palmer, R.M. & Moncada, S. Formation of nitric oxide from l-arginine in the central nervous system: a transduction mechanism for stimulation of the soluble guanylate cyclase. Proc. Natl. Acad. Sci. USA 86, 5159–5162 (1989).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Wechsler, J. et al. Isoforms of cyclic nucleotide phosphodiesterase PDE3A in cardiac myocytes. J. Biol. Chem. 277, 38072–38078 (2002).

    Article  CAS  PubMed  Google Scholar 

  19. Vandeput, F. et al. Selective regulation of cyclic nucleotide phosphodiesterase PDE3A isoforms. Proc. Natl. Acad. Sci. USA 110, 19778–19783 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Choi, Y.H. et al. Identification of a novel isoform of the cyclic-nucleotide phosphodiesterase PDE3A expressed in vascular smooth-muscle myocytes. Biochem. J. 353, 41–50 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Cone, J. et al. Comparison of the effects of cilostazol and milrinone on intracellular cAMP levels and cellular function in platelets and cardiac cells. J. Cardiovasc. Pharmacol. 34, 497–504 (1999).

    Article  CAS  PubMed  Google Scholar 

  22. Dunkerley, H.A. et al. Reduced phosphodiesterase 3 activity and phosphodiesterase 3A level in synthetic vascular smooth muscle cells: implications for use of phosphodiesterase 3 inhibitors in cardiovascular tissues. Mol. Pharmacol. 61, 1033–1040 (2002).

    Article  CAS  PubMed  Google Scholar 

  23. Linglart, A. et al. Recurrent PRKAR1A mutation in acrodysostosis with hormone resistance. N. Engl. J. Med. 364, 2218–2226 (2011).

    Article  CAS  PubMed  Google Scholar 

  24. Salem, H.K. & Thiemermann, C. Mesenchymal stromal cells: current understanding and clinical status. Stem Cells 28, 585–596 (2010).

    CAS  PubMed  Google Scholar 

  25. Begum, N., Hockman, S. & Manganiello, V.C. Phosphodiesterase 3A (PDE3A) deletion suppresses proliferation of cultured murine vascular smooth muscle cells (VSMCs) via inhibition of mitogen-activated protein kinase (MAPK) signaling and alterations in critical cell cycle regulatory proteins. J. Biol. Chem. 286, 26238–26249 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Schuster, H. et al. Autosomal dominant hypertension and brachydactyly in a Turkish kindred resembles essential hypertension. Hypertension 28, 1085–1092 (1996).

    Article  CAS  PubMed  Google Scholar 

  27. Hunter, R.W., Mackintosh, C. & Hers, I. Protein kinase C–mediated phosphorylation and activation of PDE3A regulate cAMP levels in human platelets. J. Biol. Chem. 284, 12339–12348 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Pozuelo Rubio, M., Campbell, D.G., Morrice, N.A. & Mackintosh, C. Phosphodiesterase 3A binds to 14-3-3 proteins in response to PMA-induced phosphorylation of Ser428. Biochem. J. 392, 163–172 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Castagna, M. et al. Direct activation of calcium-activated, phospholipid-dependent protein kinase by tumor-promoting phorbol esters. J. Biol. Chem. 257, 7847–7851 (1982).

    CAS  PubMed  Google Scholar 

  30. Graves, L.M. et al. Protein kinase A antagonizes platelet-derived growth factor–induced signaling by mitogen-activated protein kinase in human arterial smooth muscle cells. Proc. Natl. Acad. Sci. USA 90, 10300–10304 (1993).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Zhao, H., Guan, Q., Smith, C.J. & Quilley, J. Increased phosphodiesterase 3A/4B expression after angioplasty and the effect on VASP phosphorylation. Eur. J. Pharmacol. 590, 29–35 (2008).

    Article  CAS  PubMed  Google Scholar 

  32. Song, G.J., Fiaschi-Taesch, N. & Bisello, A. Endogenous parathyroid hormone–related protein regulates the expression of PTH type 1 receptor and proliferation of vascular smooth muscle cells. Mol. Endocrinol. 23, 1681–1690 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Hundsrucker, C. et al. Glycogen synthase kinase 3β interaction protein functions as an A-kinase anchoring protein. J. Biol. Chem. 285, 5507–5521 (2010).

    Article  CAS  PubMed  Google Scholar 

  34. Beca, S. et al. Phosphodiesterase type 3A regulates basal myocardial contractility through interacting with sarcoplasmic reticulum calcium ATPase type 2a signaling complexes in mouse heart. Circ. Res. 112, 289–297 (2013).

    Article  CAS  PubMed  Google Scholar 

  35. Lygren, B. et al. AKAP complex regulates Ca2+ re-uptake into heart sarcoplasmic reticulum. EMBO Rep. 8, 1061–1067 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Ahmad, F. et al. Regulation of SERCA2 activity by PDE3A in human myocardium: phosphorylation-dependent interaction of PDE3A1 with SERCA2. J. Biol. Chem. 290, 6763–6776 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Bassler, D., Kreutzer, K., McNamara, P. & Kirpalani, H. Milrinone for persistent pulmonary hypertension of the newborn. Cochrane Database Syst. Rev. CD007802 (2010).

  38. Pfitzer, G. Invited review: regulation of myosin phosphorylation in smooth muscle. J. Appl. Physiol. 91, 497–503 (2001).

    Article  CAS  PubMed  Google Scholar 

  39. Klopocki, E. et al. Deletion and point mutations of PTHLH cause brachydactyly type E. Am. J. Hum. Genet. 86, 434–439 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Bastepe, M. et al. Stimulatory G protein directly regulates hypertrophic differentiation of growth plate cartilage in vivo. Proc. Natl. Acad. Sci. USA 101, 14794–14799 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Chilco, P.J., Leopold, V. & Zajac, J.D. Differential regulation of the parathyroid hormone–related protein gene P1 and P3 promoters by cAMP. Mol. Cell. Endocrinol. 138, 173–184 (1998).

    Article  CAS  PubMed  Google Scholar 

  42. Drmanac, R. et al. Human genome sequencing using unchained base reads on self-assembling DNA nanoarrays. Science 327, 78–81 (2010).

    Article  CAS  PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Wang, K., Li, M. & Hakonarson, H. ANNOVAR: functional annotation of genetic variants from high-throughput sequencing data. Nucleic Acids Res. 38, e164 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Kamphans, T. & Krawitz, P.M. GeneTalk: an expert exchange platform for assessing rare sequence variants in personal genomes. Bioinformatics 28, 2515–2516 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Chen, K. et al. BreakDancer: an algorithm for high-resolution mapping of genomic structural variation. Nat. Methods 6, 677–681 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Spielmann, M. et al. Homeotic arm-to-leg transformation associated with genomic rearrangements at the PITX1 locus. Am. J. Hum. Genet. 91, 629–635 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Hofmann, N.A., Reinisch, A. & Strunk, D. Isolation and large scale expansion of adult human endothelial colony forming progenitor cells. J. Vis. Exp. 32, 1524 (2009).

    Google Scholar 

  49. Hundsrucker, C. et al. High-affinity AKAP7δ–protein kinase A interaction yields novel protein kinase A–anchoring disruptor peptides. Biochem. J. 396, 297–306 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Coin, I., Beyermann, M. & Bienert, M. Solid-phase peptide synthesis: from standard procedures to the synthesis of difficult sequences. Nat. Protoc. 2, 3247–3256 (2007).

    Article  CAS  PubMed  Google Scholar 

  51. Stefan, E. et al. Compartmentalization of cAMP-dependent signaling by phosphodiesterase-4D is involved in the regulation of vasopressin-mediated water reabsorption in renal principal cells. J. Am. Soc. Nephrol. 18, 199–212 (2007).

    Article  CAS  PubMed  Google Scholar 

  52. Schäfer, G. et al. Highly functionalized terpyridines as competitive inhibitors of AKAP-PKA interactions. Angew. Chem. Int. Edn Engl. 52, 12187–12191 (2013).

    Article  CAS  Google Scholar 

  53. Christian, F. et al. Small molecule AKAP–protein kinase A (PKA) interaction disruptors that activate PKA interfere with compartmentalized cAMP signaling in cardiac myocytes. J. Biol. Chem. 286, 9079–9096 (2011).

    Article  CAS  PubMed  Google Scholar 

  54. Aydin, A., Toliat, M.R., Bahring, S., Becker, C. & Nurnberg, P. New universal primers facilitate Pyrosequencing. Electrophoresis 27, 394–397 (2006).

    Article  CAS  PubMed  Google Scholar 

  55. Maass, P.G. et al. A cis-regulatory site downregulates PTHLH in translocation t(8;12)(q13;p11.2) and leads to Brachydactyly Type E. Hum. Mol. Genet. 19, 848–860 (2010).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank all family members for their cooperation. We thank M.-B. Köhler and M. Toliat for technical assistance. P.G.M., F.C.L., O.T. and S.B. received support from the Deutsche Forschungsgemeinschaft (DFG; BA1773/4-1, BA1773/4-2, MA5028/1-2 and MA5028/1-3) and grants-in-aid from the German Hypertension Society (Deutsche Hochdruckliga, DHL) and from the German Heart Research Foundation (F/24/13). E.K. was supported by the DFG (KL1415/4-2), the Else Kröner-Fresenius-Stiftung (2013_A145) and the German-Israeli Foundation (I-1210-286.13/2012). F.C.L. received support from the Lingen-Stiftung. F.V. and M.A.M. were supported by the US Department of Veterans Affairs (CARA-029-09F), the American Heart Association (10034439) and the University of Utah Research Foundation. James C. Melby referred one of the families. Dr. Melby died on 19 August 2007.

Author information

Authors and Affiliations

Authors

Contributions

N.B. first described this syndrome in 1973. F.C.L. and his laboratory have pursued this project since 1994. O.T., H.R.T., H. Schuster, J.J., J.T., H.H., R.H., L.O.H. and R.N. phenotyped the syndrome. D.C., M.G.B., G.P., M.H. and H.R.T. identified additional families with the syndrome. T.F.W., J.O., S.B., A.B. and F.R. performed microsatellite and SNP linkage analyses. M.G. and N.H. performed genotyping analyses within the families and also analyzed Chinese hypertensive families that showed linkage to the chromosome 12p locus. A.W., M.K., A.R., K.R. and T.L. performed cytogenetics. S.S. performed in situ mouse studies. S.M., P.M.K., D.P. and J.H. carried out Illumina whole-genome sequencing. A.A., P.G.M. and S.B. analyzed Complete Genomics whole-genome sequencing data, and A.A. identified the PDE3A mutation. H. Schulz statistically analyzed various data. C.L. and A.A. performed the confocal immunofluorescence imaging. F.Q., I.H., E.B.-K. and A.M. performed technical studies. K.M. and Y.W.-N. prepared MSCs. M.V. kindly provided unaffected MSCs and supported all the MSC investigations. Y.W.-N., A.A. and P.G.M. analyzed cell proliferation. F.V. and M.A.M. provided Flag-tagged PDE3A expression constructs and provided intellectual input. C.S. and E.K. performed ELISA assays on recombinant proteins and peptide SPOT assays. P.G.M. participated in all scientific aspects of the study and was personally responsible for the PDE3A functional assays, IC50 determinations and work with MSCs. P.G.M., F.C.L. and S.B. wrote the manuscript. The manuscript was the product of more than 20 years of research to which all authors have contributed.

Corresponding author

Correspondence to Friedrich C Luft.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 PDE3A conservation and whole-genome sequencing.

(a) Multiple-sequence alignment of the PDE3A peptide sequence near the identified mutations in five different species. The altered residues were in a highly conserved PDE3A domain. (b) Genomic coverage of Complete Genomics (CG) whole-genome sequencing of Turkish family members IV/6, IV/7, V/14 and V/30. A mean coverage of 57.39 ± 0.25× for each of the four samples was reached. (c) The detected insertion and deletion (indel) events compared to genome assembly hg19. (d) The substitution events are summarized. To use the complementary advantage of different genome sequencing platforms, we performed whole-genome sequencing of the same Turkish patient, whose DNA was also sequenced by CG, on the Illumina platform to screen for further small sequence variations, structural variants and the inversion breakpoints, which were found in fibroblasts and LCLs by interphase FISH. Among the more than 3 million SNVs and small indels, we identified the missense mutation in PDE3A that was also detected in CG sequencing. c.1334C>A in exon 4 of PDE3A (NM_000921; p.T445N) affects a highly conserved amino acid and was not observed in 1000 Genomes Project data nor among 5,000 exomes (Exome Variant Server, http://evs.gs.washington.edu/EVS/; accessed July 2013), in Sanger sequencing of all non-affected family members or in 200 unrelated Caucasian controls. (e) Electropherograms of the six different missense mutations in PDE3A that introduced amino acid changes at positions 445, 447 and 449; total number of samples analyzed; AFF, affected individuals; NON, unaffected controls.

Supplementary Figure 2 Magnetic resonance imaging (MRI) of VI/9.

(a) MRI of the hypertensive teenager with mild BDE (VI/9). The red arrows indicate the contacting vessels to the ventrolateral medulla (VLM) on the left and right sides. On the left side, the PICA is the offending artery indicating a type 1 neurovascular contact (nvc). The yellow arrows indicate the cranial nerves IX & X. (b) The red arrows indicate the contacting vessels to the VLM. (c) The upper image shows the left VLM. The red arrows indicate the PICA as an impressive loop contacting the VLM as type 1 nvc. The arrows in the middle image document the PICA attached to the right VLM. The coronal projection displays the vessel signal on both sides of the VLM.

Supplementary Figure 3 Pde3a in situ hybridization and characterization of human chondrogenically induced fibroblasts.

(a) Murine in situ hybridization with 3′ UTR probe in Pde3a detected light limb bud expression in embryos of 12.5 and 13.5 d. (b) The in situ antisense control probe in the coding sequence of Pde3a (cds) showed no expression in the developing limb buds. (c) Successful chondrogenic induction of human fibroblasts from buttock biopsies after 3 weeks in pellet culture. We used chondrogenically induced fibroblasts to focus on chondrogenic PTHLH expression. Because we were able to obtain more fibroblasts from non-affected and related controls than MSCs from the Turkish family, we performed PTHLH expression experiments on chondrogenically induced fibroblasts. After embedding and sectioning, Alcian blue or Safranin O staining detected extracellular matrix proteins that are characteristic for cartilage.

Supplementary Figure 4 cAMP and cGMP quantifications, and Lineweaver-Burke plots of Michaelis-Menten kinetics.

(a,b) HeLa cells were transiently transfected with the full-length wild-type PDE3A construct and the six full-length PDE3A mutant constructs; forskolin or l-arginine stimulation was performed to enhance intracellular cAMP or cGMP levels through adenylate or guanylate cyclase activation. Forty-eight hours after transfection, cells were collected to determine cAMP (a) and cGMP (b) levels in enzyme immunoassays. The data shown are the results of three independent experiments with minimum and maximum deviation. Significant differences were determined between wild-type PDE3A (wt) and the mutants, defining the mutations as gain-of-function mutations, causing increased cAMP hydrolysis (n = 4; P ≤ 0.001, two-tailed Student’s t test). cGMP levels were not altered in the presence of the PDE3A mutants (n = 6). (c) Western blotting detected successful expression of wild-type and mutant PDE3A proteins. Controls were cells transfected with empty vector. (dg) Linear regression analysis of transformed Michaelis-Menten data from Figure 3a of Flag-tagged PDE3A1-WT and PDE3A1-T445N. (hk) Linear regression analysis of transformed Michaelis-Menten data from Figure 3b of Flag-tagged PDE3A2-WT and PDE3A2-T445N, visualized as Lineweaver-Burke plots.

Supplementary Figure 5 IC50 measurement of cGMP.

cGMP competitively inhibits PDE3A, but nearly equivalent values were determined for Flag-tagged wild-type (WT) PDE3A1 and the tagged PDE3A1-T445N mutant. IC50 was 5.164 µM for WT and 4.78 µM for T445N (n = 3). Relevant differences were not found in three independent experiments.

Supplementary Figure 6 Functional CRE-luciferase assays in HeLa cells transiently expressing the six PDE3A mutants.

(a) HeLa cells were transfected with an empty vector (sc300-w/o; orange line), a full-length wild-type PDE3A expression construct (red line) and the six full-length PDE3A mutant expression plasmids (gray and black lines). HeLa cells were cotransfected with a cAMP-responsive element (CRE) regulating luciferase transcriptional activity under the influence of increasing forskolin concentrations to further elucidate the functional consequences of the PDE3A mutations (P < 0.002). A Renilla luciferase vector was used for standardization. The data describe the relative increase in luciferase activity normalized to the DMSO control of cells transfected with empty vector. The results are the means of three independent experiments (mean ± s.d.; n = 3; Wilcoxon-Mann-Whitney test, P < 0.002). The more hydrolyzed cAMP there was, the less luciferase expression was detected. In the presence of increasing forskolin concentrations enhancing cAMP levels, the PDE3A mutants showed a significant reduction in CRE-mediated luciferase activity as a result of the higher cAMP hydrolysis compared to wild-type PDE3A. (b) cGMP stimulation with increasing l-arginine concentrations showed that cGMP competitively inhibited cAMP hydrolysis with a significant difference between the mutants and wild-type PDE3A (P < 0.002). The more cGMP that was present, the less cAMP hydrolysis occurred and the greater the luciferase activity was.

Supplementary Figure 7 Characterization of MSCs and MSC-derived cells.

(a) FACS results of the affected patient VI/17 and one non-affected control. The analysis of surface markers detected CD105+, CD90+, CD73+, HLA-ABC+, CD31, CD34, CD45 and HLA-DR cells. (b) Plastic adherence and the multilineage potential of MSCs of one non-affected control and VI/17. MSCs and VSMCs from patient VI/9 and the second non-affected control were identically characterized and fulfilled all criteria of MSCs and VSMCs (data not shown). Immunocytochemical staining of the MSC-derived adipocytes, osteocytes and chondrocytes validated the multilineage potential of the MSCs. The fatty vacuoles show successful differentiation into adipocytes; calcium precipitates characterized MSC-derived osteocytes. Toluidine blue stained proteoglycans of sectioned chondrogenic tissue that was generated in micromass pellet cultures. Successful differentiation was similar in the MSCs from the control and affected patient VI/17. MSC characterizations fulfilled the criteria of the International Society for Cellular Therapy. (c) Myogenic differentiation of one non-affected control and MSCs from VI/17 into VSMCs after 21 d of differentiation. Semiquantitative immunofluorescence detected the smooth muscle markers smooth muscle actin (SMAα), calponin and transgelin (SM22α), which were highly expressed compared to in undifferentiated MSCs. The color spectrum indicates the expression level from low (black) to high (red).

Supplementary Figure 8 PDE3A overexpression in HeLa cells and carboxyfluorescein diacetate succinimidyl ester (CFSE) signal quantification in FACS analysis.

Intensity histogram of flow cytometry analysis of CFSE-labeled and transfected HeLa cells. After CFSE labeling, the cells were seeded and were transfected the next day with full-length wild-type PDE3A or full-length PDE3A expressing the six mutations. Measurements were performed at 24, 48 and 72 h after transfection to determine the proliferation rates. Because of mitosis and the distribution of CFSE, the signal intensity changed in the measured time of 72 h.

Supplementary Figure 9 Individual western blotting experiments.

(ah) Individual quantification of the western blot signals for each PDE3A mutant in the three independent experiments shown in Figure 4a,b. The signals were densitometrically quantified in relation to the loading control β-tubulin. Because of experimental variation and quality differences between the antibodies to phosphorylated protein, we observed variations in the investigated phosphorylation of Ser428 and Ser438 in PDE3A1 and PDE3A2. The antibody to PDE3A (Bethyl) recognizes motifs beginning at position 450. Perhaps the G449V substitution alters the neighboring protein domain and alters antibody binding, leading to the weak signals in the upper blots. However, the pooled results (Fig. 4a,b) of all mutants showed statistical significance compared to wild-type PDE3A (WT) in non-parametrical Mann-Whitney rank-sum testing.

Supplementary Figure 10 PKA and PKC phosphorylate PDE3A Ser438.

(a) Full-length wild-type (WT; NM_000921) and T445N PDE3A were transiently expressed in HeLa cells. All vectors encoded both PDE3A1 and PDE3A2, which arise from the use alternative translational start sites. In forskolin-stimulated cells, T445N PDE3A Ser438 phosphorylation was increased compared to WT and mock control (sc300 w/o). Use of the PKA inhibitor H89 showed less Ser438 phosphorylation. PDE3A-T445N phosphorylation was still stronger at Ser438 compared to controls. VASP Ser157 phosphorylation that is mediated by PKA was reduced in H89 PKA-inhibited cells. Non-phosphorylated VASP was not altered. (b) Use of PKC α, β, δ, ɛ and γ inhibitor bisindolylmaleimide I showed that PDE3A Ser438 was also phosphorylated by PKC. The combination of H89 with bisindolylmaleimide I further reduced Ser438 phosphorylation. Endogenous VASP was apparently not affected. (c) PMA stimulation alone and in combination with bisindolylmaleimide I and/or H89 inhibitors showed decreased PDE3A Ser438 and VASP Ser157 phosphorylation. Detection of non-phosphorylated VASP showed no obvious differences. However, the data suggest that other protein kinases also phosphorylate Ser438. Tubulin was used as a loading control; representative blots are shown for n = 3 experiments.

Supplementary Figure 11 Pyrosequencing of the PDE3A mutation encoding T445N in MSCs from VI/17.

The figure shows the raw data from PDE3A pyrosequencing of the patient harboring the T445N substitution. The data shown are from a reverse pyrosequencing approach. The yellow area indicates the position of the PDE3A mutation. The base “G” varies between 50 and 100%, and the base “T” varies between 0 and 50%. The data shown are from three independent replicates; error bars, s.d. The presence of each allele in the reaction is displayed as a percentage. Compared to the wild-type allele “G” (G = 99%, T = 1%), no significant differential expression was detected for the mutated PDE3A “T” allele of the patient harboring the T445N substitution (G = 55%, T = 45%). Preferred monoallelic expression, due to either the mutation or the inversion, was excluded.

Supplementary Figure 12 Expression of PTHrP and VASP in MSCs and VSMCs.

The color spectrum indicates expression levels from low (black) to high (red) in semiquantitative immunofluorescence. (a) PTHrP influences the proliferation of VSMCs (Song, G.J., Fiaschi-Taesch, N. & Bisello, A. Mol. Endocrinol.23, 1681-1680, 2009). Full-length PTHrP protein levels were increased, whereas PTHrP peptide levels were decreased, in MSCs and VSMCs from the patient harboring the T445N substitution. (b) Phosphorylation of VASP at Ser157 was reduced in VSMCs from the affected patients compared to controls; phosphorylated VASP was not detectable in MSCs.

Supplementary Figure 13 Peptide assays for the PDE3A mutants.

(a) Amino acid sequences of the synthesized peptide spots from Figure 5a. Serine residues (S; red) at position 428 or 438 were replaced by alanine (A; blue) or aspartic acid (D; green) or with prephosphorylated serine (pS; pink). Positions 445, 447 and 449 for the altered amino acids are marked in yellow; the altered amino acid is marked in gray. (b) Phosphorylation of two peptide spot membranes (30-mers representing PDE3A Ile421-Leu450; see Fig. 2a) without (– PKA) and with (+ PKA) PKA. The peptide spots for PDE3A-WT and PDE3A-T445N (black frame) are the same signals, shown in Figure 5a. (c) Amino acid sequences of the synthesized peptide spots from b. The color code is as indicated above. Serine residues at position 428 or 438 were replaced by alanine or aspartic acid or by prephosphorylated serine. Positions 445, 447 and 449 corresponding to the altered amino acids are shown in yellow; each altered amino acid is marked in gray.

Supplementary Figure 14 Individual measurements of peptide spot quantifications and further peptide sequences.

(ac) The individual signal quantifications for the peptide spot experiments shown as pooled results in Figure 5b. The summarized data for all mutations from lanes 6, 7 and 9 were statistically significant, as determined by non-parametrical two-tailed Mann-Whitney rank-sum testing (Fig. 5b). (d) Signal quantifications for the six PDE3A mutant peptides and wild-type PDE3A (WT) of lane 3 from Figure 5a and Supplementary Figure 13b (Ser428 and Ala438) showed no significant changes after PKA phosphorylation, indicating that Ser428 is not differentially phosphorylated by PKA. (e) Peptide assay to further determine the signal intensities at prephosphorylated Ser438 with alanine (A; blue) or aspartic acid (D; green) replacements or prephosphorylated serine (pS; pink) for the three different serine residues (S; red) at positions 428, 438 and 439 and the threonine residue (T; dark red) at position 440. (f) Amino acid sequences of the synthesized peptide spots from e. The serine residue at position 428 or 438 was replaced by alanine or aspartic acid or by prephosphorylated serine. The color code is as indicated above.

Supplementary Figure 15 PDE3A and its signal transductions.

Scheme for the involvement of PDE3A in VSMCs and chondrocytes. In VSMCs, the myosin light chain kinase (MLCK) and the myosin light chain phosphatase (MLCP) act in concert to phosphorylate and dephosphorylate the myosin light chain (MLC) for vascular contraction and relaxation, respectively (Pfitzer, G. J. Appl. Physiol 91, 497-503, 2001). PKA and protein kinase G (PKG) activate VASP by phosphorylation of Ser157 and Ser239, respectively. Reduced VASP Ser157 phosphorylation and higher PDE3A activity were associated with enhanced VSMC proliferation that accounts for vessel wall hyperplasia (Zhao, H., Guan, Q., Smith, C.J. & Quilley, J. et al. Eur. J. Pharmacol. 590, 29-35, 2008). Full-length PTHrP stimulated whereas PTHrP peptide (1–36) repressed VSMC proliferation (Song, G.J., Fiaschi-Taesch, N. & Bisello, A. Mol. Endocrinol. 23, 1681-1680, 2009). CREB binding to the PTHLH promoter regulates PTHLH expression in a cAMP-dependent manner. The encoded protein PTHrP transduces signals through the PTH1R (PTH/PTHrP) receptor in chondrocytes and activates the heterotrimeric G protein Gs, thereby stimulating adenylate cyclase (AC) to produce cAMP (Bastepe, M. et al. Proc. Natl. Acad. Sci. USA 101, 14794–14799, 2004). Because PDE3A hydrolyses cAMP, PKA is regulated in a cAMP-dependent manner and activates CREB, which in turn regulates PTHLH expression (Chilco, P.J., Leopold, V. & Zajac, J.D. Mol. Cell. Endocrinol.138, 173–184, 1998).

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–15 and Supplementary Tables 1–3. (PDF 11495 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Maass, P., Aydin, A., Luft, F. et al. PDE3A mutations cause autosomal dominant hypertension with brachydactyly. Nat Genet 47, 647–653 (2015). https://doi.org/10.1038/ng.3302

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ng.3302

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