Congenital hypogonadotropic hypogonadism with split hand/foot malformation: a clinical entity with a high frequency of FGFR1 mutations

Carine Villanueva, Elka Jacobson-Dickman, Cheng Xu, Sylvie Manouvrier, Andrew A Dwyer, Gerasimos P Sykiotis, Andrew Beenken, Yang Liu, Johanna Tommiska, Youli Hu, Dov Tiosano, Marion Gerard, Juliane Leger, Valérie Drouin-Garraud, Hervé Lefebvre, Michel Polak, Jean-Claude Carel, Franziska Phan-Hug, Michael Hauschild, Lacey Plummer, Jean-Pierre Rey, Taneli Raivio, Pierre Bouloux, Yisrael Sidis, Moosa Mohammadi, Nicolas de Roux, Nelly Pitteloud, Carine Villanueva, Elka Jacobson-Dickman, Cheng Xu, Sylvie Manouvrier, Andrew A Dwyer, Gerasimos P Sykiotis, Andrew Beenken, Yang Liu, Johanna Tommiska, Youli Hu, Dov Tiosano, Marion Gerard, Juliane Leger, Valérie Drouin-Garraud, Hervé Lefebvre, Michel Polak, Jean-Claude Carel, Franziska Phan-Hug, Michael Hauschild, Lacey Plummer, Jean-Pierre Rey, Taneli Raivio, Pierre Bouloux, Yisrael Sidis, Moosa Mohammadi, Nicolas de Roux, Nelly Pitteloud

Abstract

Purpose: Congenital hypogonadotropic hypogonadism (CHH) and split hand/foot malformation (SHFM) are two rare genetic conditions. Here we report a clinical entity comprising the two.

Methods: We identified patients with CHH and SHFM through international collaboration. Probands and available family members underwent phenotyping and screening for FGFR1 mutations. The impact of identified mutations was assessed by sequence- and structure-based predictions and/or functional assays.

Results: We identified eight probands with CHH with (n = 3; Kallmann syndrome) or without anosmia (n = 5) and SHFM, seven of whom (88%) harbor FGFR1 mutations. Of these seven, one individual is homozygous for p.V429E and six individuals are heterozygous for p.G348R, p.G485R, p.Q594*, p.E670A, p.V688L, or p.L712P. All mutations were predicted by in silico analysis to cause loss of function. Probands with FGFR1 mutations have severe gonadotropin-releasing hormone deficiency (absent puberty and/or cryptorchidism and/or micropenis). SHFM in both hands and feet was observed only in the patient with the homozygous p.V429E mutation; V429 maps to the fibroblast growth factor receptor substrate 2α binding domain of FGFR1, and functional studies of the p.V429E mutation demonstrated that it decreased recruitment and phosphorylation of fibroblast growth factor receptor substrate 2α to FGFR1, thereby resulting in reduced mitogen-activated protein kinase signaling.

Conclusion: FGFR1 should be prioritized for genetic testing in patients with CHH and SHFM because the likelihood of a mutation increases from 10% in the general CHH population to 88% in these patients.

Figures

Figure 1. FGFR1 mutations underlie CHH with…
Figure 1. FGFR1 mutations underlie CHH with SHFM
(A) Pedigrees of the 7 CHH and SHFM families with FGFR1 mutations; probands are denoted by arrows, SB: stillborn, OB: olfactory bulbs. A: homozygous mutation. (B) Photographs and radiographs demonstrating severe skeletal anomalies of hands and feet among probands.
Figure 2. FGFR1 mutations identified in probands…
Figure 2. FGFR1 mutations identified in probands with CHH and SHFM
(A) Schematic of FGFR1 showing the locations of all published FGFR1 mutations associated with CHH (Kallmann syndrome or normosmic CHH, red circles), Hartsfield syndrome (blue squares), septo-optic-dysplasia (green triangles). SP: signal peptide, D1: immunoglobulin domain 1, AB: acid box domain, D2: immunoglobulin domain 2, D3: immunoglobulin domain 3, TM: transmembrane domain, F: FRS2α -binding domain, TKD: tyrosine kinase domain, stars: mutations described in this study. (B) FGFR1 mutations identified in probands with CHH and SHFM, all the substituted residues of missense mutations are conserved across vertebrates (cow, mouse, chicken, xenopus, and zebrafish) and in human FGFR2 (hFGFR2).
Figure 3. The V429E substitution in FGFR1…
Figure 3. The V429E substitution in FGFR1 impedes recruitment and phosphorylation of FRS2α and FGF2-induced MAPK signaling
(A) Analysis of the impact of the V429E mutation based on the NMR structure of the FRS2 phosphotyrosine binding (PTB) domain in complex with the juxtamembrane (JM) region peptide of FGFR1. The FRS2 PTB domain and FGFR1 JM peptide are shown as purple and green ribbons respectively, and side chains of the V429 of FGFR1 and L62, M105, V112 of FRS2 are rendered in sticks. The molecular surfaces of L62, M105, and V112 of FRS2 PTB are also shown, to highlight their hydrophobic contacts with V429 of FGFR1. (B) The V429E FGFR1c mutant fails to phosphorylate FRS2 in cell based assay. BaF3 were transfected with lentiviral vectors expressing WT or V429E FGFR1c and FRS2 phosphorylation was assessed upon FGF1 treatment by western blotting using anti phospho-FRS2-α specific antibodies . WT: wild type; EV: empty vector. (C) Analysis of the impact of the V429E mutation on the ability of FGFR1 to recruit FRS2α . The assay was based on FGFR2 V430E, which is equivalent to FGFR1 V429E. Increasing concentrations of FGFR2CDWT and FGFR2CDV430E (carrying the equivalent mutation to FGFR1-V429E) ranging from 12.5 to 400 nM were passed over a CM5 chip onto which FRS2α had been immobilized. As representative of the full dataset, binding responses obtained for injections of 200 nM of FGFR2CDWT or FGFR2CDV430E are shown. The rising and falling parts of the wild type curve (blue) represent the association and dissociation phases, respectively, of FGFR2CDWT-FRS2α binding over time. At 200 nM, FGFR2CDWT exhibits maximal binding of 55 response units (blue), whereas FGFR2CDV430E shows negligible binding (red). According to a steady-state equilibrium analysis of the full data sets (not shown), FGFR2CDWT binds FRS2α with a KD of 320 nM, whereas the FGFR2CDV430E negligible binding to FRS2α. (D) The V429E mutation reduces the ability of FGFR1 to phosphorylate FRS2 on Y196 in vitro. FGFR2WT and FGFR2V430E mutant kinases were allowed to phosphorylate FRS2 fragment PTB9-200 on Y196, a tyrosine phosphorylation site known to be required for Grb2 recruitment, at room temperature for 0, 1, 3, 5 10, 15, 20, 25, 30 minutes. Following tryptic digestion, samples were analyzed by Orbitrap mass spectrometry to quantify the phospho-Y196-containing tryptic peptide. FGFR2WT is shown in blue and FGFR2V430E is shown in red. (E-F) The V429E mutation is loss-of-function in the OCFRE reporter assay (FRS2α dependent MAPK signaling) and not different from WT in the NFAT reporter assay (FRS2α independent PLCγ /IP3/Ca2+ signaling). Data shown represent the means ± S.E.M. of 3 experiments. FGFR1 WT is shown in blue, FGFR1V429E in red, empty vector in black. Relative to the maximal stimulation of WT (%), ** p<0.001, NS: not significant. (G) Total abundance and maturation of recombinant FGFR1 proteins. COS-7 cells were transfected with FGFR1 constructs and the cell lysates were subjected to deglycosylation treatment followed by western blot analysis. PNGase-treated bands represent total protein abundance levels; 140kDa Endo H-treated bands represent the mature form while 100kDa Endo H-treated bands represent immature form of FGFR1. The experiment was performed three times, tested by Mann-Whitney test for statistic significance, no significant difference in overall expression and maturation index between WT and V429E. PNGase: Peptide N-Glycosidase F-treated; Endo H: endoglycosidase H-treated; WT: wild type; EV: empty vector. (H) Cell surface abundance of the transiently transfected FGFR1 mutant in COS-7 cells. Cell surface abundance levels were measured by a radiolabeled antibody binding assay and plotted as a percentage of wild type levels. The abundance level of cell surface expressed FGFR1V429E was significantly higher than FGFR1wt. Values shown are the means ± SEM of 3 experiments each performed in quadruplicate. The difference between the mutant and the wild type receptor expression was compared by Mann-Whitney test, * p<0.05.

References

    1. Tsai PS, Gill JC. Mechanisms of disease: Insights into X-linked and autosomal-dominant Kallmann syndrome. Nature clinical practice. Endocrinology & metabolism. 2006 Mar;2(3):160–171.
    1. Miraoui H, Dwyer AA, Sykiotis GP, et al. Mutations in FGF17, IL17RD, DUSP6, SPRY4, and FLRT3 are identified in individuals with congenital hypogonadotropic hypogonadism. American journal of human genetics. 2013 May 2;92(5):725–743.
    1. Dode C, Levilliers J, Dupont JM, et al. Loss-of-function mutations in FGFR1 cause autosomal dominant Kallmann syndrome. Nature genetics. 2003 Apr;33(4):463–465.
    1. Pitteloud N, Acierno JS, Jr., Meysing A, et al. Mutations in fibroblast growth factor receptor 1 cause both Kallmann syndrome and normosmic idiopathic hypogonadotropic hypogonadism. Proceedings of the National Academy of Sciences of the United States of America. 2006 Apr 18;103(16):6281–6286.
    1. Eswarakumar VP, Lax I, Schlessinger J. Cellular signaling by fibroblast growth factor receptors. Cytokine & growth factor reviews. 2005 Apr;16(2):139–149.
    1. Deng CX, Wynshaw-Boris A, Shen MM, Daugherty C, Ornitz DM, Leder P. Murine FGFR-1 is required for early postimplantation growth and axial organization. Genes & development. 1994 Dec 15;8(24):3045–3057.
    1. Tsai PS, Moenter SM, Postigo HR, et al. Targeted expression of a dominant-negative fibroblast growth factor (FGF) receptor in gonadotropin-releasing hormone (GnRH) neurons reduces FGF responsiveness and the size of GnRH neuronal population. Molecular endocrinology. 2005 Jan;19(1):225–236.
    1. Costa-Barbosa FA, Balasubramanian R, Keefe KW, et al. Prioritizing genetic testing in patients with Kallmann syndrome using clinical phenotypes. The Journal of clinical endocrinology and metabolism. 2013 May;98(5):E943–953.
    1. Jarzabek K, Wolczynski S, Lesniewicz R, Plessis G, Kottler ML. Evidence that FGFR1 loss-of-function mutations may cause variable skeletal malformations in patients with Kallmann syndrome. Advances in medical sciences. 2012;57(2):314–321.
    1. Simonis N, Migeotte I, Lambert N, et al. FGFR1 mutations cause Hartsfield syndrome, the unique association of holoprosencephaly and ectrodactyly. Journal of medical genetics. 2013 Sep;50(9):585–592.
    1. Duijf PH, van Bokhoven H, Brunner HG. Pathogenesis of split-hand/split-foot malformation. Human molecular genetics. 2003 Apr 1;12:R51–60. Spec No 1.
    1. Vilain C, Mortier G, Van Vliet G, et al. Hartsfield holoprosencephaly-ectrodactyly syndrome in five male patients: further delineation and review. American journal of medical genetics. Part A. 2009 Jul;149A(7):1476–1481.
    1. Bouvattier C, Maione L, Bouligand J, Dode C, Guiochon-Mantel A, Young J. Neonatal gonadotropin therapy in male congenital hypogonadotropic hypogonadism. Nature reviews. Endocrinology. 2012 Mar;8(3):172–182.
    1. Raivio T, Sidis Y, Plummer L, et al. Impaired fibroblast growth factor receptor 1 signaling as a cause of normosmic idiopathic hypogonadotropic hypogonadism. The Journal of clinical endocrinology and metabolism. 2009 Nov;94(11):4380–4390.
    1. Adzhubei IA, Schmidt S, Peshkin L, et al. A method and server for predicting damaging missense mutations. Nature methods. 2010 Apr;7(4):248–249.
    1. Kumar P, Henikoff S, Ng PC. Predicting the effects of coding non-synonymous variants on protein function using the SIFT algorithm. Nature protocols. 2009;4(7):1073–1081.
    1. Ferrer-Costa C, Orozco M, de la Cruz X. Sequence-based prediction of pathological mutations. Proteins. 2004 Dec 1;57(4):811–819.
    1. Schwarz JM, Rodelsperger C, Schuelke M, Seelow D. MutationTaster evaluates disease-causing potential of sequence alterations. Nature methods. 2010 Aug;7(8):575–576.
    1. Gonzalez-Perez A, Lopez-Bigas N. Improving the assessment of the outcome of nonsynonymous SNVs with a consensus deleteriousness score, Condel. American journal of human genetics. 2011 Apr 8;88(4):440–449.
    1. Schlessinger J, Plotnikov AN, Ibrahimi OA, et al. Crystal structure of a ternary FGF-FGFR-heparin complex reveals a dual role for heparin in FGFR binding and dimerization. Molecular cell. 2000 Sep;6(3):743–750.
    1. Bae JH, Lew ED, Yuzawa S, Tome F, Lax I, Schlessinger J. The selectivity of receptor tyrosine kinase signaling is controlled by a secondary SH2 domain binding site. Cell. 2009 Aug 7;138(3):514–524.
    1. Dhalluin C, Yan KS, Plotnikova O, et al. Structural basis of SNT PTB domain interactions with distinct neurotrophic receptors. Molecular cell. 2000 Oct;6(4):921–929.
    1. Ornitz DM, Xu J, Colvin JS, et al. Receptor specificity of the fibroblast growth factor family. The Journal of biological chemistry. 1996 Jun 21;271(25):15292–15297.
    1. Ibrahimi OA, Zhang F, Eliseenkova AV, Itoh N, Linhardt RJ, Mohammadi M. Biochemical analysis of pathogenic ligand-dependent FGFR2 mutations suggests distinct pathophysiological mechanisms for craniofacial and limb abnormalities. Human molecular genetics. 2004 Oct 1;13(19):2313–2324.
    1. Ichida M, Finkel T. Ras regulates NFAT3 activity in cardiac myocytes. The Journal of biological chemistry. 2001 Feb 2;276(5):3524–3530.
    1. Goetz R, Mohammadi M. Exploring mechanisms of FGF signalling through the lens of structural biology. Nature reviews. Molecular cell biology. 2013 Mar;14(3):166–180.
    1. Sykiotis GP, Plummer L, Hughes VA, et al. Oligogenic basis of isolated gonadotropin-releasing hormone deficiency. Proceedings of the National Academy of Sciences of the United States of America. 2010 Aug 24;107(34):15140–15144.
    1. Bailleul-Forestier I, Gros C, Zenaty D, Bennaceur S, Leger J, de Roux N. Dental agenesis in Kallmann syndrome individuals with FGFR1 mutations. International journal of paediatric dentistry / the British Paedodontic Society [and] the International Association of Dentistry for Children. 2010 Jul;20(4):305–312.
    1. Laitinen EM, Vaaralahti K, Tommiska J, et al. Incidence, phenotypic features and molecular genetics of Kallmann syndrome in Finland. Orphanet journal of rare diseases. 2011;6:41.
    1. Burgar HR, Burns HD, Elsden JL, Lalioti MD, Heath JK. Association of the signaling adaptor FRS2 with fibroblast growth factor receptor 1 (Fgfr1) is mediated by alternative splicing of the juxtamembrane domain. The Journal of biological chemistry. 2002 Feb 8;277(6):4018–4023.
    1. Twigg SR, Burns HD, Oldridge M, Heath JK, Wilkie AO. Conserved use of a non-canonical 5′ splice site (/GA) in alternative splicing by fibroblast growth factor receptors 1, 2 and 3. Human molecular genetics. 1998 Apr;7(4):685–691.
    1. Solomon BD, Pineda-Alvarez DE, Mercier S, Raam MS, Odent S, Muenke M. Holoprosencephaly flashcards: A summary for the clinician. American journal of medical genetics. Part C, Seminars in medical genetics. 2010 Feb 15;154C(1):3–7.
    1. Raivio T, Falardeau J, Dwyer A, et al. Reversal of idiopathic hypogonadotropic hypogonadism. The New England journal of medicine. 2007 Aug 30;357(9):863–873.
    1. Li C, Xu X, Nelson DK, Williams T, Kuehn MR, Deng CX. FGFR1 function at the earliest staes of mouse limb development plays an indispensable role in subsequent autopod morphogenesis. Development. 2005 Nov;132(21):4755–4764.
    1. Verheyden JM, Lewandoski M, Deng C, Harfe BD, Sun X. Conditional inactivation of Fgfr1 in mouse defines its role in limb bud establishment, outgrowth and digit patterning. Development. 2005 Oct;132(19):4235–4245.
    1. van Bokhoven H, Hamel BC, Bamshad M, et al. p63 Gene mutations in eec syndrome, limb-mammary syndrome, and isolated split hand-split foot malformation suggest a genotype-phenotype correlation. American journal of human genetics. 2001 Sep;69(3):481–492.
    1. Yang A, Schweitzer R, Sun D, et al. p63 is essential for regenerative proliferation in limb, craniofacial and epithelial development. Nature. 1999 Apr 22;398(6729):714–718.
    1. Sun X, Mariani FV, Martin GR. Functions of FGF signalling from the apical ectodermal ridge in limb development. Nature. 2002 Aug 1;418(6897):501–508.
    1. Ozen RS, Baysal BE, Devlin B, et al. Fine mapping of the split-hand/split-foot locus (SHFM3) at 10q24: evidence for anticipation and segregation distortion. American journal of human genetics. 1999 Jun;64(6):1646–1654.
    1. Falardeau J, Chung WC, Beenken A, et al. Decreased FGF8 signaling causes deficiency of gonadotropin-releasing hormone in humans and mice. The Journal of clinical investigation. 2008 Aug;118(8):2822–2831.

Source: PubMed

3
購読する