ArticlePDF Available

Digenic mutations account for variable phenotypes in idiopathic hypogonadotropic hypogonadism

Authors:

Abstract and Figures

Idiopathic hypogonadotropic hypogonadism (IHH) due to defects of gonadotropin-releasing hormone (GnRH) secretion and/or action is a developmental disorder of sexual maturation. To date, several single-gene defects have been implicated in the pathogenesis of IHH. However, significant inter- and intrafamilial variability and apparent incomplete penetrance in familial cases of IHH are difficult to reconcile with the model of a single-gene defect. We therefore hypothesized that mutations at different IHH loci interact in some families to modify their phenotypes. To address this issue, we studied 2 families, one with Kallmann syndrome (IHH and anosmia) and another with normosmic IHH, in which a single-gene defect had been identified: a heterozygous FGF receptor 1 (FGFR1) mutation in pedigree 1 and a compound heterozygous gonadotropin-releasing hormone receptor (GNRHR) mutation in pedigree 2, both of which varied markedly in expressivity within and across families. Further candidate gene screening revealed a second heterozygous deletion in the nasal embryonic LHRH factor (NELF) gene in pedigree 1 and an additional heterozygous FGFR1 mutation in pedigree 2 that accounted for the considerable phenotypic variability. Therefore, 2 different gene defects can synergize to produce a more severe phenotype in IHH families than either alone. This genetic model could account for some phenotypic heterogeneity seen in GnRH deficiency.
Content may be subject to copyright.
Research article
The Journal of Clinical Investigation http://www.jci.org
Digenic mutations account
for variable phenotypes in idiopathic
hypogonadotropic hypogonadism
Nelly Pitteloud,1 Richard Quinton,2,3 Simon Pearce,2,4 Taneli Raivio,1 James Acierno,1
Andrew Dwyer,1 Lacey Plummer,1 Virginia Hughes,1 Stephanie Seminara,1 Yu-Zhu Cheng,2,4
Wei-Ping Li,2,4 Gavin Maccoll,5 Anna V. Eliseenkova,6 Shaun K. Olsen,6 Omar A. Ibrahimi,6
Frances J. Hayes,1 Paul Boepple,1 Janet E. Hall,1 Pierre Bouloux,5
Moosa Mohammadi,6 and William Crowley Jr.1
1Reproductive Endocrine Unit of the Department of Medicine and Harvard Reproductive Endocrine Science Centers, Massachusetts General Hospital, Boston,
Massachusetts, USA. 2Department of Endocrinology and 3Royal Victoria Infirmary, School of Clinical Medical Sciences, and 4Institute for Human Genetics,
University of Newcastle upon Tyne, Newcastle upon Tyne, United Kingdom. 5Department of Endocrinology, Royal Free Hospital, London, United Kingdom.
6Department of Pharmacology, New York University School of Medicine, New York, New York, USA.
Idiopathic hypogonadotropic hypogonadism (IHH) due to defects of gonadotropin-releasing hormone
(GnRH) secretion and/or action is a developmental disorder of sexual maturation. To date, several single-
gene defects have been implicated in the pathogenesis of IHH. However, significant inter- and intrafamilial
variability and apparent incomplete penetrance in familial cases of IHH are difficult to reconcile with the
model of a single-gene defect. We therefore hypothesized that mutations at different IHH loci interact in some
families to modify their phenotypes. To address this issue, we studied 2 families, one with Kallmann syndrome
(IHH and anosmia) and another with normosmic IHH, in which a single-gene defect had been identified: a
heterozygous FGF receptor 1 (FGFR1) mutation in pedigree 1 and a compound heterozygous gonadotropin-releas-
ing hormone receptor (GNRHR) mutation in pedigree 2, both of which varied markedly in expressivity within and
across families. Further candidate gene screening revealed a second heterozygous deletion in the nasal embry-
onic LHRH factor (NELF) gene in pedigree 1 and an additional heterozygous FGFR1 mutation in pedigree 2
that accounted for the considerable phenotypic variability. Therefore, 2 different gene defects can synergize
to produce a more severe phenotype in IHH families than either alone. This genetic model could account for
some phenotypic heterogeneity seen in GnRH deficiency.
Introduction
Genetic analyses  of idiopathic  hypogonadotropic hypogonad-
ism (IHH), an important human disease model with implications 
for the discovery of genes responsible for human puberty, have 
provided considerable insight into the genes that control sexual 
maturation. IHH is a clinically and genetically heterogenous dis-
order resulting in gonadotropin-releasing hormone (GnRH) defi-
ciency that can be inherited as an X-linked, autosomal recessive, 
or autosomal dominant trait. IHH has  been considered to  be a 
monogenic disorder with several loci identified to date: Kallmann
syndrome 1 sequence (KAL1) (1–3), FGF receptor 1 (FGFR1) (4), proki-
neticin 2 (PROK2), and prokineticin receptor 2 (PROKR2) (5) under-
lie cases of Kallmann syndrome (KS), while gonadotropin-releasing
hormone receptor(GNRHR) (6), FGFR1 (7), and G protein–coupled
receptor 54 (GPR54) (8, 9) underlie normosmic IHH (nIHH). Addi-
tionally, nasal embryonic LHRH factor (NELF) has been implicated in 
the pathogenesis of KS (10). Despite these advances, conundrums 
remain in understanding the genetic basis of IHH. For example, 
there is a puzzling clinical heterogeneity of the reproductive and 
nonreproductive phenotypes both within and across IHH families 
carrying identical single gene mutations (4, 6, 11–14). Therefore, 
a given genotype at a single locus cannot reliably predict the phe-
notypic manifestations in any given member of affected families. 
Additionally, some mutations in genes accounting for IHH, espe-
cially FGFR1, apparently demonstrate incomplete penetrance (4). 
Finally, defects in the identified genetic loci account for only a 
small percentage (<30%) of cases. Thus, it is likely that other major 
IHH loci remain to be discovered and/or that the remaining (>70%) 
cases are caused by the interplay of several contributing genes.
We hypothesized that IHH often involves defects in more than 1 
gene. Herein we report evidence of IHH caused by the interaction 
of 2-gene defects (FGFR1 and NELFin pedigree 1 and GNRHR and 
FGFR1 in pedigree 2). In addition, in vitro biochemical character-
ization of FGFR1 and NELF mutants is provided.
Results
Pedigree 1. Pedigree 1 has several affected members (Figure 1). The 
proband (no. 1-03) was referred to an endocrinologist at age 21 for 
failure to undergo puberty. His presentation was consistent with a 
severe KS phenotype. He was unvirilized, had eunuchoidal propor-
tions (height, 176 cm; span, 185 cm), bilateral gynecomastia, micro-
phallus, prepubertal testes (2 ml; normal, >12 ml), a repaired cleft 
lip/palate, and clinodactyly. His serum gonadotropin levels were 
undetectable, testosterone (T) was 1 nmol/l, and inhibin B 74 pg/ml; 
otherwise, he  had normal pituitary function  and brain imaging. 
Nonstandard abbreviations used: D3, immunoglobulin-like domain 3; FGFR1, FGF 
receptor 1; GnRH, gonadotropin-releasing hormone; GNRHR, gonadotropin-releasing 
hormone receptor; IHH, idiopathic hypogonadotropic hypogonadism; KAL1, Kall-
mann syndrome 1 sequence; KS, Kallmann syndrome; NELF, nasal embryonic LHRH 
factor; nIHH, normosmic IHH; SPR, surface plasmon resonance; T, testosterone.
Conflict of interest: The authors have declared that no conflict of interest exists.
Citation for this article:J. Clin. Invest. doi:10.1172/JCI29884.
research article
The Journal of Clinical Investigation http://www.jci.org
Formal testing revealed hyposmia (score of 29/40, below fifth percen-
tile for his age) (15). Two subsequent years of gonadotropin therapy 
induced full virilization and sperm production. His father (no. 1-01) 
had a history of delayed puberty (growth spurt and full virilization 
after age 17) and congenital anosmia (score of 13/40), and his adult 
serum T was 18.1 nmol/l. The proband’s mother (no. 1-02) had clino-
dactyly and Duane ocular retraction syndrome and was menopausal. 
The proband’s sister (no. 1-05) exhibited midline defects, including 
a bifid nose and high arched palate; the brother (no. 1-04) exhibited 
clinodactyly only. The mother and both siblings had normal puberty 
and a normal sense of smell as determined by formal testing.
Mutational analysis of theFGFR1gene. The proband (no. 1-03) car-
ries a unique heterozygous mutation (c.1025 TC) in exon 7 pre-
dicted to substitute a leucine for serine at position 342 (p.L342S) 
in the immunoglobulin-like domain 3 (D3) of FGFR1 (Figure 1, 
Figure 2, A and B, and Figure 3G). This change was also found 
in the affected father (no. 1-01) and the affected sister (no. 1-05) 
(Figure 1) but not in 200 white controls.
Structural and biochemical analysis of the L342S mutation impli-
cate a loss of FGF8b signaling through FGFR1c in the pathology of
KS/IHH. The isoforms FGFR1b and FGFR1c are generated 
by alternative  splicing of exons 8A and 8B, respectively 
(16). To date, FGFR1mutations causing IHH have only 
been identified in exon 8B (17), implicating FGFR1c in the 
pathogenesis of  IHH. Because  L342 is highly conserved 
among the “c” splice isoforms of FGFR1–3 across species 
(Figure 2B), we used the FGFR2c-FGF8b crystallography 
model (18) to study L342S. The corresponding amino acid 
in FGFR2c, L343, is a key constituent in the hydrophobic 
groove of D3 and is extensively engaged by FGF8b (18) (Figure 3G). 
This leucine accounts for the unique binding specificity of FGF8b 
for the “c” isoforms of FGFR1–3. Consistent with these structural 
data, surface plasmon resonance (SPR) analysis revealed a dramatic 
loss (20-fold) in the affinity of the L342S mutant for FGF8b (Figure 
3, C and F), with only a small decrease (2-fold) in affinity for FGF1 
(Figure 3, A and D) and (3-fold) FGF2 (Figure 3, B and E).
L342S FGFR1c missense mutation is a loss-of-function mutation. L6 myo-
blasts transiently expressing WT FGFR1c were treated with FGF8b, 
which induced a 6-fold increase in LUC reporter gene expression (Fig-
ure 3H). In agreement with SPR results, the L342S FGFR1c was silent 
when expressed alone. As the KS subject harbored a heterozygous 
L342S mutation, we further coexpressed the WT FGFR1 and L342S 
in 1:1 and 1:2 ratios. The results are compatible with the hypothesis 
that this mutant acts as a dominant-negative mutation.
The variable degree of sexual maturation among family mem-
bers carrying the same FGFR1 mutation led to further candidate 
Figure 
Identification of FGFR1 (p.L342S)
and NELF (8-bp intronic deletion)
mutations in pedigree 1; identification
of GNRHR [p.Q106R] and [p.R262Q]
and FGFR1 (p.R470L) mutations in
pedigree 2. Only subjects harboring 2
gene defects have IHH. Probands are
identified by arrows; circles denote
females, squares denote males. Del,
NELF intronic deletion.
Figure 
Schematic showing location of the 2 FGFR1 mutations and
conservation of L342 and R470 residues across species and
FGFRs. (A) The FGFR1 gene is located on chromosome 8p.
FGFR1 contains 18 exons with intervening introns not drawn
due to scale. SP, signal peptide; TM, transmembrane domain;
TK, tyrosine kinase domain. (B) Comparison of L342 and R470
across species and within the FGFR family.
research article
The Journal of Clinical Investigation http://www.jci.org
gene screening and the identification of an additional deletion in 
NELF in the severely affected proband.
Mutational analysis of the NELFgene. A heterozygous 8-bp deletion 
ending 14 bp before exon 10 (c.1159-14_-22del) was identified in the 
proband (no. 1-03), his mother (no. 1-02), and his brother (no. 1-04) 
(Figure 4A). This deletion was not observed in 384 white controls.
RT-PCR. The HEK-293 cells transfected with the WT plasmid 
expressed NELFexons 8–11 at the expected size (291 bp). However, 
cells transfected with the 8-bp intronic deletion plasmid expressed 
an additional 257-bp splice form of NELF mRNA, lacking exon 
10 (Figure 4C). The missplicing of mRNA lacking exon 10 results 
in a premature stop codon that predicts a truncated NELF pro-
tein product (p.Y376X) rather than the full-length product, which 
comprises 528 residues.
Immunohistology of FNC-C4 cells. Colocalization of both NELF and 
GnRH1 in FNC-B4 cells was demonstrated by immunohistochemistry 
(Figure 4D) and RT-PCR (data not shown).
Genotype-phenotype correlations.  Proband no. 1-03  exhibited a 
severe KS phenotype with absent puberty, microphallus, hyposmia, 
undetectable serum luteinizing hormone and follicle stimulating 
hormone levels, hypogonadal T levels, and low inhibin B levels. He 
also had clinodactyly and cleft lip and palate. He harbored both a 
paternally derived heterozygous FGFR1 mutation (c.1025 TC, 
p.L342S), resulting in weaker binding to FGF8b, and a maternally 
derived heterozygous 8-bp intronic deletion of NELF, resulting in 
a splicing defect of exon 10 and premature stop codon (Figures 
1–4). While a single gene defect was associated with an attenuated 
phenotype (i.e., delayed puberty and anosmia in the father carrying 
L342S only), 2 mutant genes/gene products (FGFR1 and NELF) 
synergized to produce a more severe KS phenotype.
Pedigree 2. All members of the second, previously reported (14) 
pedigree were normosmic by formal smell testing, including 2 sis-
ters affected with nIHH (Figure 1). The proband (no. 2-03) present-
ed at 17 with primary amenorrhea, no breast development, short 
fourth metacarpals, and osteoporosis. She had undetectable serum 
gonadotropin and estradiol (E2) levels but otherwise normal pitu-
itary function and cranial imaging. As previously reported, pulsatile 
GnRH induced ovulation in the proband, but she had 5 consecutive 
miscarriages (14). Her sister also presented at 18 with primary amen-
orrhea, absent breast development, and scoliosis. She had 2 success-
Figure 
The L342S mutation dramatically reduces affinity of FGFR1c for FGF8b, implicating decreased FGF8b/FGFR1c signaling in the etiology of
KS/IHH. (AF) The L342S mutation reduces the affinity of FGFR1c for FGF8b. Varying concentrations of WT (AC) or L342S FGFR1c mutant
(DF) were injected over a CM5 chip onto which FGF1 (A and D), FGF2 (B and E), and FGF8b (C and F) were immobilized. Analyte concentra-
tions are indicated as follows: 31.25 nM in gray, 62.5 nM in violet, 125 nM in green, 250 nM in red, 500 nM in blue. (G) The location of L343 in
FGFR2c, the residue corresponding to L342 in FGFR1c, is mapped onto the FGF8b-FGFR2c structure (18). The L342S mutation should weaken
key hydrophobic contacts between F32, V36, and F93 of FGF8b and D3 of FGFR. Gray: molecular surface of FGFR; orange: FGF8b. The side
chains of selected residues are shown. The molecular surface of the hydrophobic groove of FGFR D3 (yellow) is rendered transparent so that
the side chain of L343 (the residue corresponding to L342 of FGFR1c) is visible. (H) L342S FGFR1 heterozygous mutation is a loss-of-function
mutation. WT and L342S FGFR1c were transiently transfected into L6 myoblasts with an FGFR1-responsive osteocalcin promoter luciferase
construct. FGF8b treatment of WT FGFR1c induced a 6-fold increase in LUC reporter gene expression, while the L342S FGFR1c alone remained
silent. The coexpression of the WT and L342S FGFR1c suggests that this mutation acts as a dominant negative.
research article
The Journal of Clinical Investigation http://www.jci.org
ful pregnancies on gonadotropin therapy. Her daughter (no. 2-06) 
underwent normal puberty and has scoliosis, and male twins (nos. 
2-07 and 2-08) were born without cryptorchidism or microphallus 
and have yet to undergo puberty (10 years old). The proband’s father 
has a history of delayed puberty, scoliosis, and bilateral hearing loss. 
The brother was born with a cardiac septal defect and 3 fused cervi-
cal bones and went through normal puberty. The mother is meno-
pausal but had normal reproductive function.
GNRHRmutational analysis. A compound heterozygous muta-
tion in GNRHR [Q106R] and [R262Q] was identified in the pro-
band (no. 2-03) and her sister (no. 2-05) (14). Both exhibited severe 
nIHH with absent puberty. Their father (no. 2-01) with a history 
of delayed puberty was heterozygous for R262Q, while the mother 
with no reproductive phenotype (no. 2-02) was heterozygous for 
Q106R. In addition, the brother (no. 2-04) and the niece (no. 2-06) 
were WT, while the twins (nos. 2-07 and 2-08) were heterozygous for 
Q106R and R262Q, respectively (Figure 1). In 1997, de Roux et al. 
described 2 siblings with partial nIHH as evidence by spontane-
ous testicular growth and active spermatogenesis in the male and 
Tanner IV breast development in the female 
carrying  the same compound  heterozygous 
[Q106R] and [R262Q] mutation in the GNRHR
gene. In vitro studies demonstrated that each 
mutant, [Q106R] and [R262Q], resulted in loss 
of function (6). The variable expressivity rang-
ing from absent puberty to a partial puberty 
across families carrying the same compound 
heterozygous GNRHR mutation led to further 
candidate gene screening, which identified an 
additional FGFR1 mutation in our pedigree.
FGFR1 mutational analysis.  We  identified 
a heterozygous mutation  in FGFR1 (c.1409 
GT) in exon 10 in both nIHH sisters. This 
nucleotide change is predicted to substitute an 
arginine for leucine at position 470 (p.R470L) 
(Figure 2A). Screening of the entire pedigree 
revealed this change only in the father with 
delayed puberty (no. 2-01) and the unaffected 
niece (no. 2-06) (Figure 1). R470 is conserved 
across species (Figure 2B), and R470L was not 
detected in 200 white controls.
The R470L mutation reduces the tyrosine kinase
activity of FGFR1. In the FGFR1 kinase struc-
ture, arginine 470 (R470) lies at the junction 
between the kinase domain  and  the  juxta-
membrane region (Figure 2A). The side chain 
of R470 makes 3 hydrogen bonds with D468; 
the latter  engages in a  hydrogen bond  with 
K536  from  the  αC  helix  (Figure  5A)  (19). 
Crystal structures of unphosphorylated and 
phosphorylated kinase domains show distinct 
and reversible movements of the αC helix dur-
ing the  kinase activation/inactivation cycle. 
R470 both facilitates the conformation of the 
juxtamembrane/kinase region and contributes 
to proper αC positioning. Therefore, the R470L 
mutation should negatively impact the tyrosine 
kinase activity of FGFR1. Indeed, comparison 
of WT FGFR1and the R470L mutant revealed a 
marked decrease in the tyrosine kinase activity 
in the mutant, indicating a loss of function (Figure 5B).
Genotype-phenotype correlations. The sisters severely affected with 
nIHH harbor  both a compound heterozygous mutation  in the 
GNRHR gene [Q106R] and [R262Q] and a heterozygous FGFR1
mutation (R470L) — therefore a triallelic pattern of inheritance. The 
additional FGFR1 mutation could explain the variable phenotypic 
expressivity seen between pedigrees harboring the same compound 
heterozygous GNRHR mutation (6). Interestingly, the father (no. 
2-01), with a combination of 2 heterozygous mutant alleles (GNRHR
and FGFR1), had delayed puberty, while other family members carry-
ing only 1 mutant allele display a normal reproductive phenotype. 
Discussion
IHH appears to follow the pattern of several disorders that were 
initially thought to be monogenic but have subsequently proven 
to be caused by or modulated by more than 1 gene defect (20, 21). 
In contrast to polygenic traits, these oligogenic disorders involve 
the synergistic action of mutant alleles at a small number of loci. 
This report expands our understanding of the genetics of IHH and 
Figure 
A 8-bp intronic NELF deletion results in missplicing of exon 10. (A) Direct DNA sequencing of
cloned PCR products of the proband’s genomic DNA revealed a heterozygous 8-bp deletion
in intron 9. This deletion is part of a direct 8-bp repeat (tgtggcct) and occurs 14 bp upstream
of the exon 10 acceptor (5) splice site. The lower case part of the sequence indicates introns
while the upper case indicates exons. (B) Predicted cDNAs in the exon 8–11 region of the
NELF gene. The proband was found to have an 8-bp deletion in intron 9 (black triangle), 14
bp upstream of exon 10. (C) The result of RT-PCR using HEK-293 cell mRNA from cells
transfected with WT NELF genomic construct containing exons 8–11 and the mutant NELF
construct (8-bp deletion in intron 9). The HEK-293 cells transfected with the WT NELF con-
struct show a normally spliced NELF exon 8–11 RT-PCR product, corresponding to the pre-
dicted size of 291 bp. The cells transfected with the mutant NELF construct, however, show
an additional band of 257 bp corresponding to the expected size of a transcript lacking exon
10. The RNA from cells transfected with either WT or mutant genomic construct show an
additional product of 794 bp (786 bp for the mutant construct), reflecting PCR amplification of
the residual plasmid DNA. (D) Colocalization of NELF and GnRH1 in the olfactory epithelial
cell line FNC-B4 by immunohistochemistry. Original magnification, ×150.
research article
The Journal of Clinical Investigation http://www.jci.org
suggests oligogenicity may underly other monogenic disorders 
also characterized by incomplete penetrance and variable pheno-
types within and between families.
IHH is currently described as a monogenic disorder resulting in 
defective GnRH secretion or action, with 7 loci implicated in the 
pathogenesis of the disease to date. However, the genotype-pheno-
type correlations from specific mutations in different loci have been 
imprecise. An illustrative example is FGFR1. Numerous heterozygous 
FGFR1 mutations underlie cases of KS (4, 13, 22–25) and nIHH (7, 
26), characterized by marked phenotypic variability both within and 
between families and apparent incomplete  penetrance.  Further-
more, although FGFR1 mutations are known as causing an auto-
somal dominant form of IHH, there is a report of 1 KS subject with 
a very severe phenotype harboring a homozygous FGFR1 mutation 
(4). These data suggest a dosage effect of the FGFR1 mutant alleles 
and the existence of modifier genes and/or environmental effects 
leading to phenotypic variability across and within families.
Herein, we report a proband who carries heterozygous muta-
tions in both FGFR1 and NELFand exhibits a severe KS pheno-
type with cleft palate (pedigree 1). The FGFR1 mutation, L342S, 
alters the spectrum of FGF binding by decreasing FGF8b binding 
specifically. These results were corroborated by our reporter gene 
assay, which showed the L342S mutation to be a loss-of-function 
mutation, potentially having a dominant-negative effect in the 
heterozygous state. In addition, these data implicate FGF8b as 
a key ligand for FGFR1c in the pathogenesis of KS. These find-
ings are in accordance with studies on mice with a hypomorphic 
Fgf 8 allele, which exhibited defected nasal cavity development and 
olfactory bulbs dysgenesis, a phenotype similar to that observed 
in the Fg fr1 conditional knockout mouse (27–29). Therefore, we 
believe that the L342S FGFR1 mutation contributed to both the 
KS and the craniofacial phenotype of the proband. However, the 
KS phenotype failed to segregate with the L342S mutation, as only 
the proband exhibited KS. Interestingly, he carried an additional 
heterozygous intronic NELF deletion resulting in missplicing of 
exon 10 and  a premature stop  codon. NELF plays a key role in 
GnRH and olfactory neuron outgrowth (30–32) and is colocalized 
with GnRH1 in human olfactory stem cells (Figure 4D). NELF has 
previously been implicated as a potential locus for IHH in an affect-
ed hypogonadotropic patient with a unique heterozygous muta-
tion (c.1438AG; p.T480A), though the biology of this mutant 
remains unexplored (10).  Interestingly, the mother  (no.  1-02), 
who carried the NELFmutation only, had Duane ocular retraction 
syndrome (OMIM 126800), a disorder of neuronal migration that 
has been associated with KS (33). Therefore, we believe that the 
heterozygous mutations in both the NELF and the FGFR1 gene 
synergized to cause severe KS in the proband, while individuals 
carrying only 1 of the 2 mutations do not present with KS.
Since 1997, GNRHR mutations have been known to cause a reces-
sive form of nIHH. However, unexplained variable expressivity of 
IHH across pedigrees carrying the same compound heterozygous 
mutation in GNRHR (6, 14) led us  to screen further candidate 
genes. As a result, we identified a triallelic mode of inheritance 
for IHH  in pedigree 2. Subjects carrying  only 1 mutant  allele 
(FGFR1 or GNRHR) had a normal reproductive phenotype. Fur-
ther, the individual (no. 1-01) carrying both heterozygous FGFR1
and GNRHR mutations had only a mild reproductive phenotype 
(delayed puberty), while the 2 sisters harboring both a loss-of-func-
tion compound heterozygous mutation in GNRHR and a loss-of-
function heterozygous FGFR1 mutation (R470L) had severe IHH 
phenotypes. Although IHH was initially thought to be transmit-
ted as a recessive trait, this pedigree indicates a triallelic digenic 
inheritance of IHH, suggesting that in some families, more than 
2 mutant alleles might be required to manifest IHH. Similar pat-
terns of mutation lead to Bardet-Biedl syndrome (34).
Only 30% of IHH subjects studied to date harbor a mutation 
in 1 of the genes known to cause IHH, and thus it is not possi-
ble to infer the frequency of oligogenicity in IHH. The concept 
of digenicity was evoked but not proven in a recent report that 
described a KS subject carrying both a KAL1 missense mutation 
and a PROKR2 missense mutation (5). However, the functionality 
of these mutants was not explored, and no data were presented on 
the pedigree of this subject.
The discovery of oligogenic traits in IHH raises both conceptual 
and practical issues. Because IHH is no longer considered to be a 
monogenic disorder, the transmission of a trait through families is 
no longer synonymous with the transmission of 1 specific mutant 
allele. Therefore, we must rethink our current approach in study-
ing the genetics of IHH, separating the segregation of a trait from 
the segregation of a specific mutant allele (21). The results of this 
study indicate that labeling mutant alleles as dominant or recessive 
is often an oversimplification. Oligogenicity also has implications 
for genetic counseling regarding IHH. Finally, the question remains 
Figure 
The R470L FGFR1 mutation is loss-of-function. (A) The RL sub-
stitution abolishes hydrogen bonds that play a role in the positioning
of the C-helix of FGFR1. The unphosphorylated “low-activity” form of
the FGFR1 kinase domain (PDB ID: 1FGK) is represented as a ribbon
diagram. The N-lobe of the FGFR1 kinase is shown in green, with
the exception of the αC helix, which is in blue. Red: C-lobe of FGFR1
kinase; gray: linker connecting the N- and C-lobes; yellow: activa-
tion loop in the C-lobe. Selected residues are shown, and hydrogen
bonds are represented as dashed lines. R470 indirectly contributes to
the proper positioning of αC in the kinase domain and hence kinase
regulation by engaging in 3 hydrogen bonds with D468, which in turn
engages in a hydrogen bond with K536 from αC helix. (B) The R470L
mutation reduces the tyrosine kinase activity of FGFR1.
research article
The Journal of Clinical Investigation http://www.jci.org
as to whether the variability of expressivity in IHH is controlled by a 
small number of major loci or a large number of minor loci.
The molecular basis of oligogenicity is poorly understood. To 
date, one of the best examples of the molecular basis of digenic 
inheritance is the direct interaction of the mutants retinal outer 
segment membrane protein  1 and retinal  degeneration  slow, 
which causes retinitis pigmentosa (35). The 2 mutants combine to 
prevent the formation of a tetrameric complex that is important 
for the integrity of the photoreceptors. Further examples in which 
synergistic mutations at different loci cause a disease are found in 
diseases characterized by mutations in receptor-ligand pairs, such 
as Hirschsprung disease (RET and GDNF) (36). In addition, indi-
rect interactions between 2 mutated proteins have been reported 
in severe insulin resistance where PPARG and PPP1R3A mutants 
are expressed in 2 different tissues (37).
In contrast, little is known about the molecular pathways where-
by the 2 pairs of mutant proteins (NELF and FGFR1; GnRHR 
and FGFR1) contribute to  the  IHH  phenotype. FGFR1 and 
NELF are both expressed in GnRH neurons (30, 38, 39). There-
fore, the 2 mutant proteins may act at different levels of the same 
intracellular pathway and may quantitatively contribute to its 
progressive dysfunction until a critical threshold is reached, thus 
producing the disease phenotype. Alternatively, these 2 proteins 
may participate in a multiprotein complex that becomes progres-
sively compromised by the additional mutations. In the other 
gene pair (FGFR1and GNRHR), the FGFR1 mutant is anticipated 
to reduce the number of GnRH neurons, as evidenced by a mouse 
model with targeted dominant-negative FGFR1 in the GnRH neu-
rons (40). These mice, although fertile, display a 30% decrease in 
the number of GnRH neurons in the hypothalamus. Conversely, 
the GNRHR mutants  result  in GnRH resistance  that would 
require a compensatory increase in GnRH to restore reproductive 
function. Such a putative compensatory mechanism would be 
impaired in the individuals with an additional FGFR1 mutation. 
Thus, deficient GnRH migration, secretion, and/or action could 
be caused by the interaction of these mutants in the same or dif-
ferent developmental pathways, leading to subtle amplification 
effects of each mutant protein.
This report expands our understanding of the genetics of this 
disorder by demonstrating that IHH can be caused by the com-
bination of gene defects. The complexity of this genetic model 
implies a revision of the genetic terminology of IHH. Furthermore, 
these data predict that mutational analysis in multiple genes for 
seemingly monogenic disorders will become increasingly frequent 
and may lead to greater accuracy in phenotypic predictions. In the 
case of IHH, studying the interaction of a small number of key 
loci will undoubtedly be key to understanding some of the broader 
variability of the timing of normal puberty.
Methods
Mutational analysis. Genetic  studies were approved by Partners Health-
care Human Research Committee, and all participants provided written 
informed consent.
Whole blood samples were obtained from subjects and genomic DNA 
extracted. Exons and  exon-intron  boundaries  were amplified using 
standard  PCR  techniques  for  GNRHR (Genbank  accession  number
NM_001012763 [ref. 41]); KAL1 (accession number NM_000216 [ref. 42]); 
GPR54 (accession number NM_032551 [ref. 9]); FGFR1 (accession num-
ber BC018128 [ref. 4]); and NELF (accession number ENSG00000165802 
[ref. 10]). Nonsense changes resulting in a truncated protein, frameshift, 
insertion, or deletion were categorized as definite mutations. Nucleotide 
changes, which were (a) absent from the Single Nucleotide Polymorphism 
database (http://www.ncbi.nlm.nih.gov/projects/SNP/)  and expressed 
sequence tags and (b) absent in at least 170 ethnically matched healthy 
controls were identified as mutations. All genes and proteins are described 
using standard nomenclature (43).
Control population. To determine whether the observed bp changes in the 
genes cited above were normal variants, a cohort of adult healthy white 
subjects (n ≥ 200) were screened. 
Structural analysis of the effects ofFGFR1mutations on FGFR1 function. Crystal 
structures of the FGFR1 kinase domain (Research Collaboratory for Struc-
tural Bioinformatics Protein Data Bank identification [PDB ID]: 1FGK) 
(19) and the extracellular ligand binding region of FGFR2c in complex 
with FGF8b and heparin oligosaccharide (PDB ID: 2FDB) (18) were used 
to map the potential effects of the mutations. Crystal structures were visu-
alized using program O (44). 
SPR analysis. Real-time biomolecular interactions between WT and L342S 
mutant FGFR1c extracellular domains with FGF1, FGF2, and FGF8 were 
characterized with a Biacore 3000 instrument as previously described (18).
Plasmids and in vitromutagenesis. Human WT FGFR1c cDNA (NM_000604) 
was subcloned into a pcDNA3.1+ expression vector (Invitrogen). TheFGFR1c
L342S mutation was introduced into the human FGFR1c cDNA by site-direct-
ed mutagenesis (QuikChange Site-Directed Mutagenesis Kit; Stratagene) and 
subcloned into pcDNA3.1+ using HindIII and XhoI restriction sites.
Reporter gene assay. To demonstrate that the heterozygous L342S muta-
tion was a loss-of-function mutation, we used an FGF reporter bioassay. WT 
FGFR1c and L342S FGFR1c were transfected along with an osteocalcin FGF 
response element (OCFRE) promoter/luciferase reporter (45) into L6 myo-
blasts, a cell line largely devoid of endogenous FGFRs and FGFs (46, 47). L6 
myoblasts were maintained in DMEM containing penicillin (100 U/l), strep-
tomycin (100 μg/l), and 10% FCS (Invitrogen). A total of 4×104 cells per well 
in 24-well plates were seeded and 24 hours later transiently transfected with 
a total of 400 ng of DNA consisting of 100 ng of WT FGFR1c cDNA, 30 ng 
of OCFRE reporter gene, and 0, 100, and 200 ng of the mutated FGFR1c 
(L342S) and empty vector using FuGENE 6 Reagent (Roche Diagnostics). 
Twenty-four hours after transfection, the cells were treated with increas-
ing doses of FGF8b (from 0 to 2,000 pM) in culture medium containing 
0.1% FCS and 1 μg/ml heparin. Following 16 hours of incubation, the cells 
were lysed, and luciferase activities were measured using Promega Luciferase 
Assay System (Promega). SinceRenilla and β-galactosidase were induced by 
FGFs, the luciferase data are reported directly. Experiments were performed 
in triplicate and repeated once. The data are reported as mean ± SD of a 
representative experiment performed in triplicate.
Kinase assay. The tyrosine autophosphorylation activity of the WT and 
R470L mutant FGFR1 kinase domains were quantified using a continuous 
spectrophotometric assay as previously described (48).
Generation of the NELFmutant construct. To assess the effect of the intronic 
deletion of the NELF gene on exon splicing, WT and mutant genomic NELF
constructs were made in the expression vector pCR3.1 and transfected into 
HEK-293 cells (Lipofectamine 2000; Invitrogen). Oligonucleotide primers 
(F5: CAGTGACCTGCAGAGCTC-3and R5: CCAGATCTTGGCTCCCTT-
GTG-3) spanning the genomic sequence from within exons 8 and 11 were 
used to amplify, via PCR, a fragment of WT DNA (794 bp) and mutant DNA 
(786 bp in the deleted allele) (Figure 4B). RT-PCR was expected to yield a 
291-bp product from a NELF cDNA with correct splicing of exons 9 and 10. 
Immunohistochemistry of FNC-B4 cells. FNC-B4 cells, isolated from human 
fetal olfactory epithelia, have characteristics of both secretory and migra-
tory GnRH neurons (39, 49). FNC-B4 cells were stained with either GnRH1 
or NELF antibodies (30, 49). Secondary antibodies conjugated with either 
Texas red or FITC were used to visualize GnRH1 and NELF staining.
research article
The Journal of Clinical Investigation http://www.jci.org
Acknowledgments
We thank David Ornitz for his kind  gift of the OCFRE lucifer-
ase reporter. This work was  funded by  NIH grants DE13686, 
HD15788, and HD028138.
Received for publication  July 27, 2006, and accepted in revised 
form November 27, 2006
Address correspondence to: Nelly Pitteloud, Reproductive Endo-
crine Unit, BHE 5, Massachusetts General Hospital, Boston, 02114 
Massachusetts, USA. Phone: (617) 724-1830; Fax: (617) 726-5357; 
E-mail: npitteloud@partners.org.
Moosa Mohammadi and William Crowley Jr. contributed equally 
to this work.
1. Franco, B., et al. 1991. A gene deleted in Kallmann’s 
syndrome shares homology with neural cell adhe-
sion and axonal  path-finding molecules. Nature.
353:529–536.
2. Legouis, R., et al. 1991. The candidate gene for the 
X-linked Kallmann syndrome  encodes a protein 
related to adhesion molecules. Cell.67:423–435.
3. Hardelin, J.P., et al. 1993.  Heterogeneity in  the 
mutations responsible for X chromosome- linked 
Kallmann syndrome. Hum. Mol. Genet.2:373–377.
4. Dode, C., et al. 2003. Loss-of-function mutations 
in FGFR1 cause  autosomal dominant Kallmann 
syndrome. Nat. Genet.33:463–465.
5. Dodé, C., et al. 2006. Kallmann syndrome:  muta-
tions in  the genes  encoding prokineticin-2 and 
prokineticin receptor-2. PLoS Genet.2:e175.
6. de Roux, N., et al. 1997. A family with hypogonado-
tropic hypogonadism and mutations in the gonad-
otropin-releasing hormone receptor. N. Engl. J. Med.
337:1597–1602.
7. Pitteloud, N., et al. 2006. Mutations in fibroblast 
growth factor receptor 1 cause both Kallmann syn-
drome and  normosmic idiopathic hypogonado-
tropic hypogonadism. Proc. Natl. Acad. Sci. U. S. A.
103:6281–6286.
8. de Roux, N., et al. 2003. Hypogonadotropic hypo-
gonadism due  to loss of  function of  the KiSS1-
derived peptide receptor GPR54.  Proc. Natl. Acad.
Sci. U. S. A.100:10972–10976.
9. Seminara, S.B., et al. 2003. The GPR54 gene as a reg-
ulator of puberty. N. Engl. J. Med.349:1614–1627.
10. Miura, K., Acierno, J.S., Jr., and Seminara, S.B. 2004. 
Characterization of the  human nasal embryonic 
LHRH factor gene, NELF, and a mutation screen-
ing among  65 patients with  idiopathic hypogo-
nadotropic hypogonadism (IHH).  J. Hum. Genet.
49:265–268.
11. de  Rou x,  N. ,  et  a l.  199 9.  Th e  sam e  mol ecula r 
defects of the  gonadotropin-releasing hormone 
receptor determine a variable degree of hypogo-
nadism in  affected  kindred. J. Clin. Endocrinol.
Metab.84:567–572.
12. Massin, N., et al. 2003. X chromosome-linked Kall-
mann syndrome: clinical heterogeneity in three sib-
lings carrying an intragenic deletion of the KAL-1 
gene. J. Clin. Endocrinol. Metab.88:2003–2008.
 13. Pitteloud, N.,  et al. 2005.  Reversible kallmann 
syndrome, delayed puberty, and isolated anosmia 
occurring in a single family with a mutation in the 
fibroblast growth  factor receptor 1  gene. J. Clin.
Endocrinol. Metab.90:1317–1322.
14. Seminara, S.B., et al. 2000. Successful  use of pul-
satile gonadotropin–releasing hormone  (GnRH) 
for ovulation induction and pregnancy in a patient 
with GnRH receptor mutations. J. Clin. Endocrinol.
Metab.85:556–562.
 1 5. Mays ton, M.J. ,  Har riso n,  L. M., and  S teph ens, 
J.A. 1999. A  neurophysiological study of mirror 
movements in adults and  children. Ann. Neurol.
45:583–594.
16. Eswarakumar, V.P., Lax,  I., and  Schlessinger, J. 
2005. Cellular signaling by fibroblast growth factor 
receptors. Cytokine Growth Factor Rev.16:139–149.
17. Zitzmann, M., and Nieschlag,  E. 2000. Hormone 
substitution in male  hypogonadism.  Mol. Cell.
Endocrinol.161:73–88.
18. Olsen, S.K., et al. 2006. Structural basis  by which 
alternative splicing modulates the organizer activ-
ity of FGF8 in the brain. Genes Dev.20:185–198.
19. Mohammadi, M., Schlessinger,  J., and Hubbard, 
S.R. 1996. Structure of the FGF receptor tyrosine 
kinase  domain  reveals  a  novel  autoinhibitory 
mechanism. Cell.86:577–587.
20. Carlton, V.E., et al. 2003. Complex inheritance of 
familial hypercholanemia with  associated muta-
tions in TJP2 and BAAT. Nat. Genet.34:91–96.
21. Badano, J.L., and K atsanis, N. 2002. Beyo nd Men-
del: an  evolving view  of human  genetic disease 
transmission. Nat. Rev. Genet.3:779–789.
22. Albuisson, J., et al. 2005. Kallmann syndrome:  14 
novel mutations in KAL1 and FGFR1 (KAL2). Hum.
Mutat.25:98–99.
23. Sato, N., et al. 2004. Clinical assessment and muta-
tion analysis of Kallmann syndrome 1 (KAL1) and 
fibroblast growth factor receptor  1 (FGFR1,  or 
KAL2)  in f ive fa milies  and 18 sporadic patients. 
J. Clin. Endocrinol. Metab.89:1079–1088.
24. Zenaty, D., et  al. 2006.  Paediatric phenotype  of 
Kallmann syndrome  due to mutations of fibro-
blast growth factor receptor 1 (FGFR1).  Mol. Cell.
Endocrinol.254255:78–83.
25. Pitteloud, N., et al. 2006. Mutations in f ibroblast 
growth factor receptor 1 cause Kallmann syndrome 
with a wide spectrum of reproductive phenotypes. 
Mol. Cell. Endocrinol.254255:60–69.
26. Trarbach, E.B., et al. 2006. Novel fibroblast growth 
factor receptor 1 mutations in patients with congeni-
tal hypogonadotropic hypogonadism with and with-
out anosmia. J. Clin. Endocrinol. Metab.91:4006–4012.
27. Hebert, J.M., Lin, M., Partanen, J., Rossant, J., and 
McConnell, S.K. 2003.  FGF signaling through 
FGFR1 is required for olfactory bulb morphogen-
esis. Development.130:1101–1111.
28. Meyers,  E.N., Lewandoski, M.,  and Martin,  G.R. 
1998. An Fgf8  mutant allelic series generated by 
Cre- and Flp-mediated recombination. Nat. Genet.
18:136–141.
29. Kawauchi, S., et al. 2005. Fgf8 expression defines a 
morphogenetic center required for olfactory neuro-
genesis and nasal cavity development in the mouse. 
Development.132:5211–5223.
 30. Kramer, P.R.,  and  Wray,  S.  2000.  Novel gene 
expressed in nasal region influences outgrowth of 
olfactory axons and migration of luteinizing hor-
mone-releasing hormone (LHRH) neurons. Genes
Dev.14:1824–1834.
31. Wray, S. 2002. Molecular mechanisms  for migra-
tion of placodally derived GnRH neurons. Chem.
Senses.27:569–572.
32. Gonzalez-Martinez, D., et al. 2004. Anosmin-1 mod-
ulates fibroblast growth factor receptor 1 signaling 
in human gonadotropin-releasing hormone olfac-
tory neuroblasts through a heparan sulfate-depen-
dent mechanism. J. Neurosci.24:10384–10392.
33. Cordonnier, M., Hanozet, V., Van Nechel, C., Fery, 
F., and  Aberkane, J.  1990. Bilateral  Duane syn-
drome associated with hypogonadotropic hypo-
gonadism and anosmia (Kallmann syndrome) [In 
French]. Bull. Soc. Belge Ophtalmol.239:29–35.
34. Beales, P.L., et al.  2001. Genetic and  mutational 
analyses of a large multiethnic Bardet-Biedl cohort 
reveal a minor involvement of BBS6 and delineate 
the critical intervals of other loci. Am. J. Hum. Genet.
68:606–616.
35. Kajiwara, K., Berson, E.L., and  Dryja, T.P. 1994. 
Digenic retinitis pigmentosa due to mutations at 
the unlinked peripherin/RDS and ROM1 loci. Sci-
ence.264:1604–1608.
36. Parisi,  M.A.,  and  Kapur,  R.P.  2000.  Genetics  of 
Hirschsprung disease. Curr. Opin. Pediatr.12:610–617.
37. Savage, D.B., et  al. 2002. Digenic  inheritance of 
severe insulin resistance in a human pedigree. Nat.
Genet.31:379–384.
38. Gill, J.C., Moenter, S.M., and Tsai, P.S. 2004. Devel-
opmental regulation  of gonadotropin-releasing 
hormone neurons by Fgf signaling. Endocrinology. 
145:3830–3839.
39. Romanelli, R.G., et al. 2004. Expression and func-
tion of gonadotropin-releasing hormone (GnRH) 
receptor in human olfactory GnRH-secreting neu-
rons: an autocrine GnRH loop underlies neuronal 
migration. J. Biol. Chem.279:117–126.
40. Tsai,  P.S. ,  et  al.  2 005.  Targ eted  ex press ion  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. Mol. Endo-
crinol.19:225–236.
41. Beranova, M., et al. 2001. Prevalence,  phenotypic 
spectrum, and modes of inheritance of gonado-
tropin-releasing  hormone  receptor  mutations 
in idiopathic hypogonadotropic  hypogonadism. 
J. Clin. Endocrinol. Metab.86:1580–1588.
42. Oliveira, L.M., et al. 2001. The importance of auto-
somal genes in Kallmann syndrome: genotype-phe-
notype correlations and neuroendocrine character-
istics. J. Clin. Endocrinol. Metab.86:1532–1538.
43. Antonarakis,  S.E. 1998.  Recommendations for  a 
nomenclature system for human gene mutations. 
Nomenclature Working Group.Hum. Mutat.11:1–3.
44. Mohammadi, M., et al. 1996. Identification of six 
novel autophosphor ylation sites  on  fibroblast 
growth factor receptor 1 and elucidation of their 
importance  in  receptor  activation  and  signal 
transduction. Mol. Cell. Biol.16:977–989.
45. Newberry, E.P., Boudreaux, J.M., and Towler, D.A. 
1996. The rat osteocalcin fibroblast growth factor 
(FGF)-responsive element: an okadaic acid-sensi-
tive, FGF-selective transcriptional response motif. 
Mol. Endocrinol.10:1029–1040.
46. Olwin, B.B., and  Hauschka, S.D. 1989.  Cell type 
and tissue distribution of the  fibroblast growth 
factor receptor. J. Cell. Biochem.39:443–454.
47. Roghani,  M.,  et  al.  1994.  Hepari n inc reases  the 
affinity of basic fibroblast growth factor  for its 
receptor but  is not  required for binding.  J. Biol.
Chem.269:3976–3984.
48. Barker, S.C., et al. 1995. Characterization of pp60c-
src tyrosine kinase  activities using a  continuous 
assay: autoactivation of the enzyme is an intermo-
lecular autophosphorylation process. Biochemistry.
34:14843–14851.
49. Vannelli, G.B., et al.  1995. Neuroblast long-term 
cell cultures from human fetal  olfactory epithe-
lium respond to odors. J. Neurosci.15:4382–4394.
... Genital anomalies such as micropenis may be accompanied by a pathogenic variant in EHMT1 by itself, but mutation in the NSMF (OMIM #608137) may also be related to genital anomalies. NSMF is known to contribute to the HH phenotype in an oligogenic pattern, but monoallelic deleterious NSMF variants are not sufficient to cause HH. 15,16) Therefore, we considered that EHMT1 contributed to the genital anomaly in our patient. ...
Article
Full-text available
Kleefstra syndrome is caused by chromosome 9q34.3 deletion or heterozygous mutations in the euchromatin histone methyl transferase 1 ( EHMT1 ) gene. It can be accompanied by intellectual disability, distinctive facial features, microcephaly, psychiatric disorders, hypotonia in childhood, hearing loss, heart defects, renal defects, epilepsy, speech anomalies, and obesity. Furthermore, genital anomalies are present in 30%–40% of male patients with Kleefstra syndrome, but their mechanisms have not been elucidated. Herein, we report a patient with Kleefstra syndrome presenting with micropenis. The patient was transferred to Kyungpook National University Children's Hospital for management of imperforate anus on the day of birth. Physical examination revealed micropenis with stretched penile length of 0.9 cm and facial dysmorphisms, including hypertelorism and anteverted nares. Chromosomal microarray revealed 424-kb heterozygous deletion at chromosome 9q34.3 (arr[hg19] 9q34.3 (140,234,315-140,659,055)x1). Among the involved main OMIM genes, phenotypically relevant genes were EHMT1 and NSMF . Endocrinological investigation showed low basal gonadotropin and testosterone levels. Anterior pituitary hormones and steroid hormone levels were in the normal range. Testicular function was normal based on human chorionic gonadotropin stimulation test. The patient experienced improvement in penile length growth with intramuscular testosterone enanthate injection initiated at 4 months of age. The purpose of this study is to describe the etiology, endocrine laboratory tests, and treatment of micropenis in Kleefstra syndrome.
... It has been shown that the chemotactic property of PROK2 allows for the migration of neural precursors to the olfactory bulb (OB) and regulates its morphogenesis [23]. Thus, the PROK2/PROKR2 axis in the OB plays a role in the development and migration/survival of GnRH neurons [23][24][25], and mutations in PROK2 and PROK2 genes lead to a defect in OB neuron migration associated with Kallmann syndrome [26]. The induction of PROK2 expression in response to pathological insults, such as ischemia and stroke, contributes to inflammation and negatively affects the pathophysiological response. ...
Article
Full-text available
Prokineticins are a family of small proteins with diverse roles in various tissues, including the brain. However, their specific effects on different cerebral cell types and blood–brain barrier (BBB) function remain unclear. The aim of this study was to investigate the effects of PROK1 and PROK2 on murine cerebral cell lines, bEnd.3, C8.D30, and N2a, corresponding to microvascular endothelial cells, astrocytes and neurons, respectively, and on an established BBB co-culture model. Western blot analysis showed that prokineticin receptors (PROKR1 and PROKR2) were differentially expressed in the considered cell lines. The effect of PROK1 and PROK2 on cell proliferation and migration were assessed using time-lapse microscopy. PROK1 decreased neural cells’ proliferation, while it had no effect on the proliferation of endothelial cells and astrocytes. In contrast, PROK2 reduced the proliferation of all cell lines tested. Both PROK1 and PROK2 increased the migration of all cell lines. Blocking PROKRs with the PROKR1 antagonist (PC7) and the PROKR2 antagonist (PKR-A) inhibited astrocyte PROK2-mediated migration. Using the insert co-culture model of BBB, we demonstrated that PROKs increased BBB permeability, which could be prevented by PROKRs’ antagonists.
... The nonsense CHD7 has never been reported in the literature. It could explain the CHARGE phenotype observed in this patient, while the PROK2 variant has already been described in the literature in association with cHH/KS [16,39], and its pathogenicity was supported by functional studies [1]. ...
Article
Full-text available
Congenital hypogonadotropic hypogonadism (cHH)/Kallmann syndrome (KS) is a rare genetic disorder with variable penetrance and a complex inheritance pattern. Consequently, it does not always follow Mendelian laws. More recently, digenic and oligogenic transmission has been recognized in 1.5–15% of cases. We report the results of a clinical and genetic investigation of five unrelated patients with cHH/KS analyzed using a customized gene panel. Patients were diagnosed according to the clinical, hormonal, and radiological criteria of the European Consensus Statement. DNA was analyzed using next-generation sequencing with a customized panel that included 31 genes. When available, first-degree relatives of the probands were also analyzed to assess genotype–phenotype segregation. The consequences of the identified variants on gene function were evaluated by analyzing the conservation of amino acids across species and by using molecular modeling. We found one new pathogenic variant of the CHD7 gene (c.576T>A, p.Tyr1928) and three new variants of unknown significance (VUSs) in IL17RD (c.960G>A, p.Met320Ile), FGF17 (c.208G>A, p.Gly70Arg), and DUSP6 (c.434T>G, p.Leu145Arg). All were present in the heterozygous state. Previously reported heterozygous variants were also found in the PROK2 (c.163del, p.Ile55*), CHD7 (c.c.2750C>T, p.Thr917Met and c.7891C>T, p.Arg2631*), FLRT3 (c.1106C>T, p.Ala369Val), and CCDC103 (c.461A>C, p.His154Pro) genes. Molecular modeling, molecular dynamics, and conservation analyses were performed on three out of the nine variants identified in our patients, namely, FGF17 (p.Gly70Arg), DUSP6 (p.Leu145Arg), and CHD7 p.(Thr917Met). Except for DUSP6, where the L145R variant was shown to disrupt the interaction between β6 and β3, needed for extracellular signal-regulated kinase 2 (ERK2) binding and recognition, no significant changes were identified between the wild-types and mutants of the other proteins. We found a new pathogenic variant of the CHD7 gene. The molecular modeling results suggest that the VUS of the DUSP6 (c.434T>G, p.Leu145Arg) gene may play a role in the pathogenesis of cHH. However, our analysis indicates that it is unlikely that the VUSs for the IL17RD (c.960G>A, p.Met320Ile) and FGF17 (c.208G>A, p.Gly70Arg) genes are involved in the pathogenesis of cHH. Functional studies are needed to confirm this hypothesis.
... In the past few years, the traditional Mendelian view of congenital hypogonadotropic hypogonadism as a monogenic disorder has been revised after the identification of oligogenic forms in about 20% of patients. 91 Furthermore, reversal of congenital hypogonadotropic hypogonadism occurs in about 10-20% of patients with congenital hypogonadotropic hypogonadism, which challenges the view that the condition is lifelong. 92 More than 60 genes are associated with congenital hypogonadotropic hypogonadism (table). ...
Article
Puberty is a major maturational event; its mechanisms and timing are driven by genetic determinants, but also controlled by endogenous and environmental cues. Substantial progress towards elucidation of the neuroendocrine networks governing puberty has taken place. However, key aspects of the mechanisms responsible for the precise timing of puberty and its alterations have only recently begun to be deciphered, propelled by epidemiological data suggesting that pubertal timing is changing in humans, via mechanisms that are not yet understood. By integrating basic and clinical data, we provide a comprehensive overview of current advances on the physiological basis of puberty, with a particular focus on the roles of kisspeptins and other central transmitters, the underlying molecular and endocrine mechanisms, and the pathways involved in pubertal modulation by nutritional and metabolic cues. Additionally, we have summarised molecular features of precocious and delayed puberty in both sexes, as revealed by clinical and genetic studies. This Review is a synoptic up-to-date view of how puberty is controlled and of the pathogenesis of major pubertal alterations, from both a clinical and translational perspective. We also highlight unsolved challenges that will seemingly concentrate future research efforts in this active domain of endocrinology.
Article
Full-text available
Human sexual and reproductive development is regulated by the hypothalamic-pituitary-gonadal (HPG) axis, which is primarily controlled by the gonadotropin-releasing hormone (GnRH) acting on its receptor (GnRHR). Dysregulation of the axis leads to conditions such as congenital hypogonadotropic hypogonadism (CHH) and delayed puberty. The pathophysiology of GnRHR makes it a potential target for treatments in several reproductive diseases and in congenital adrenal hyperplasia. GnRHR belongs to the G protein-coupled receptor family and its GnRH ligand, when bound, activates several complex and tissue-specific signaling pathways. In the pituitary gonadotrope cells, it triggers the G protein subunit dissociation and initiates a cascade of events that lead to the production and secretion of the luteinizing hormone (LH) and follicle-stimulating hormone (FSH) accompanied with the phospholipase C, inositol phosphate production, and protein kinase C activation. Pharmacologically, GnRHR can be modulated by synthetic analogues. Such analogues include the agonists, antagonists, and the pharmacoperones. The agonists stimulate the gonadotropin release and lead to receptor desensitization with prolonged use while the antagonists directly block the GnRHR and rapidly reduce the sex hormone production. Pharmacoperones include the most recent GnRHR therapeutic approaches that directly correct the misfolded GnRHRs, which are caused by genetic mutations and hold serious promise for CHH treatment. Understanding of the GnRHR’s genomic and protein structure is crucial for the most appropriate assessing of the mutation impact. Such mutations in the GNRHR are linked to normosmic hypogonadotropic hypogonadism and lead to various clinical symptoms, including delayed puberty, infertility, and impaired sexual development. These mutations vary regarding their mode of inheritance and can be found in the homozygous, compound heterozygous, or in the digenic state. GnRHR expression extends beyond the pituitary gland, and is found in reproductive tissues such as ovaries, uterus, and prostate and non-reproductive tissues such as heart, muscles, liver and melanoma cells. This comprehensive review explores GnRHR’s multifaceted role in human reproduction and its clinical implications for reproductive disorders.
Article
Objective: To study if a pituitary or ovarian defect contributes to subfertility of the female Nsmf knockout (KO) mice, an animal model of the hypogonadotropic hypogonadism gene NSMF. Design: Analysis of hypothalamic, pituitary and ovarian gene expression at baseline, serum gonadotropin levels before and after GnRH stimulation, ovarian response and implantation following superovulation, gonadotropin effects following ovariectomy, and ovarian NSMF protein expression. Setting: University research laboratory PATIENTS: None; mice were utilized INTERVENTIONS: GnRH stimulation, superovulation, and ovariectomy in separate experiments MAIN OUTCOME MEASURES: Gene expression in the hypothalamus, pituitary, and ovary; ovarian response and implantation following superovulation; serum gonadotropins following GnRH stimulation and ovariectomy; western blot to measure ovarian NSMF expression.
Article
Olfaction, as one of our 5 senses, plays an important role in our daily lives. It is connected to proper nutrition, social interaction, and protection mechanisms. Disorders affecting this sense consequently also affect the patients’ general quality of life. Because the underlying genetics of congenital olfactory disorders (COD) have not been thoroughly investigated yet, this systematic review aimed at providing information on genes that have previously been reported to be mutated in patients suffering from COD. This was achieved by systematically reviewing existing literature on 3 databases, namely PubMed, Ovid Medline, and ISI Web of Science. Genes and the type of disorder, that is, isolated and/or syndromic COD were included in this study, as were the patients’ associated abnormal features, which were categorized according to the affected organ(-system). Our research yielded 82 candidate genes/chromosome loci for isolated and/or syndromic COD. Our results revealed that the majority of these are implicated in syndromic COD, a few accounted for syndromic and isolated COD, and the least underly isolated COD. Most commonly, structures of the central nervous system displayed abnormalities. This study is meant to assist clinicians in determining the type of COD and detecting potentially abnormal features in patients with confirmed genetic variations. Future research will hopefully expand this list and thereby further improve our understanding of COD.
Article
The mechanism underlying mirrored activity/movements in normal individuals is unknown. To investigate this, we studied 11 adults and 39 children who performed sequential finger-thumb opposition or repetitive index finger abduction. Surface electromyographic (EMG) activity recorded from the left and right first dorsal interosseous muscles (1DI) during unilateral sequential finger-thumb opposition (voluntarily activated muscle, 1DIvol) showed mirrored EMG activity (homologous muscle of the opposite hand, 1DImm) that decreased with increasing age. The time of onset of involuntary compared with voluntary EMG activity was variable but could start at the same time. A significant increase in E2 (transcortical component) size of the cutaneomuscular reflex recorded from the 1DImm indicated increased excitability of the motor cortex ipsilateral to the 1DIvol during active index finger abduction compared with the 1DIvol relaxed. Transcranial magnetic stimulation, using the Bistim technique, indicated that the transcallosal inhibitory pathway in children may not operate in the same way as in the adult. Cross-correlation analysis did not detect shared synaptic input to motoneuron pools innervating homologous left and right hand muscles. We conclude that the mirrored movements/activity observed in healthy adults and children are produced by simultaneous activation of crossed corticospinal pathways originating from both left and right motor cortices.
Article
Bardet-Biedl syndrome (BBS) is a rare autosomal recessive disorder characterized primarily by obesity, polydactyly, retinal dystrophy, and renal disease. The significant genetic and clinical heterogeneity of this condition have substantially hindered efforts to positionally clone the numerous BBS genes, because the majority of available pedigrees are small and the disorder cannot be assigned to any of the six known BBS loci. Consequently, the delineation of critical BBS intervals, which would accelerate the discovery of the underlying genetic defect(s), becomes difficult, especially for loci with minor contributions to the syndrome. We have collected a cohort of 163 pedigrees from diverse ethnic backgrounds and have evaluated them for mutations in the recently discovered BBS6 gene (MKKS) on chromosome 20 and for potential assignment of the disorder to any of the other known BBS loci in the human genome. Using a combination of mutational and haplotype analysis, we describe the spectrum of BBS6 alterations that are likely to be pathogenic; propose substantially reduced critical intervals for BBS2, BBS3, and BBS5; and present evidence for the existence of at least one more BBS locus. Our data also suggest that BBS6 is a minor contributor to the syndrome and that some BBS6 alleles may act in conjunction with mutations at other BBS loci to cause or modify the BBS phenotype.
Article
Kallmann syndrome associates hypogonadotropic hypogonadism and anosmia and is probably due to a defect in the embryonic migration of olfactory and GnRH-synthesizing neurons. The Kallmann gene had been localized to Xp22.3. In this study 67 kb of genomic DNA, corresponding to a deletion interval containing at least part of the Kallmann gene, were sequenced. Two candidate exons, identified by multiparameter computer programs, were found in a cDNA encoding a protein of 679 amino acids. This candidate gene (ADMLX) is interrupted in its 3' coding region in the Kallmann patient, in which the proximal end of the KAL deletion interval was previously defined. A 5' end deletion was detected in another Kallmann patient. The predicted protein sequence shows homologies with the fibronectin type III repeat. ADMLX thus encodes a putative adhesion molecule, consistent with the defect of embryonic neuronal migration.