Abstract

Disorders of sex development (DSD) are a group of clinical conditions with variable presentation and genetic background. Females with or without development of secondary sexual characters and presenting with primary amenorrhea (PA) and a 46,XY karyotype are one of the classified groups in DSD. In this study, we aimed to determine the genetic mutations in 25 females with PA and a 46,XY karyotype to show correlations with their phenotypes. Routine Sanger sequencing with candidate genes like SRY, AR, SRD5A2, and SF1, which are mainly responsible for 46,XY DSD in adolescent females, was performed. In a cohort of 25 patients of PA with 46,XY DSD, where routine Sanger sequencing failed to detect the mutations, next-generation sequencing of a targeted gene panel with 81 genes was used for the molecular diagnosis. The targeted sequencing identified a total of 21 mutations including 8 novel variants in 20 out of 25 patients with DSD. The most frequently identified mutations in our series were in AR (36%), followed by SRD5A2 (20%), SF1 (12%), DHX37 (4%), HSD17B3 (4%), and DMRT2 (4%). We could not find any mutation in the DSD-related genes in five (20%) patients due to complex molecular mechanisms in 46,XY DSD, highlighting the possibility of new DSD genes which are yet to be discovered in these disorders. In conclusion, genetic testing, including cytogenetics and molecular genetics, is important for the diagnosis and management of 46,XY DSD cases.

Introduction

The disorders of sex development (DSD) are congenital conditions in which development of chromosomal, gonadal, or anatomical sex is atypical (Strauss et al., 2017). Adolescent females with DSD and presenting with primary amenorrhea (PA) have variable phenotypic presentations depending on the defects of gonadal development (testis/ovary) (e.g. complete gonadal dysgenesis known as Sweyer syndrome) or conditions that lead to defective hormone biosynthesis or action of androgens (e.g. androgen insensitivity syndrome, AIS) (Buonocore et al., 2019). Among DSD, the incidence of phenotypic females with a 46,XY karyotype and presenting with PA is about 14%, as estimated by the practice committee of American Society of Reproductive Medicine (Practice Committee of the American Society for Reproductive Medicine, 2008). These conditions are estimated to occur with a frequency of 1:10 000 individuals in the population (Wang et al., 2018).

Development of sex is mainly dependent on genetic sex, which in turn decides the gonadal sex, and as genital development goes by way of the genetic and gonadal sex, the phenotypic sex of the individual is determined. Several genes are involved in the complex process of sex determination in humans and malfunctioning of any one of them leads to DSD (Paliwal et al., 2011). Development of gonads is bipotential during the first 7–8 weeks of mammalian embryogenesis, and subsequently differentiation occurs to give rise to testes or ovaries in the individuals with a 46,XY or 46,XX karyotype, respectively (Paliwal et al., 2011). The individuals with normal 46,XX (female) karyotype and normal ovaries should develop female external and internal genitalia. But, there are individuals who have 46,XY (male) karyotype with testes or dysgenetic gonads and labeled as female or ambiguous genitalia at the time of birth, and subsequently grow up as females. Such females present with PA with or without development of normal secondary sexual characters at the time of puberty. This can be a social stigma for the individual as well as the family members and may cause a psychological burden for them. Despite having the Y chromosome, such females may not develop normal testes and normal male internal and external genitalia. The reason for non-development of testes is a defect in the genes that determine the sex of these patients, due to which the bipotential gonads fail to develop into either testes or ovaries. The sex-determining genes are expressed in the developing genital ridges and the differentiation of gonads into one sex or the other is a complex process due to the interaction of the various gene products responsible for gonadal and genital differentiation (MacLaughlin and Donahoe, 2004). If the phenotypic females develop normal testes, the location of such testicular tissue is at an abnormal region and there may be defective androgen production or failure of conversion of testosterone into its active form, dihydrotestosterone (DHT), by the enzyme steroid 5 alpha-reductase (SRD5A2 gene defects) or, more often, there may be defective action of androgen due to defective androgen receptors (AR gene defects). This leads to a problem with masculinization of the external genitalia giving rise to the abnormal phenotype appearing as female (Practice Committee of the American Society for Reproductive Medicine, 2008). Thus, the majority of the females with an XY karyotype and PA are diagnosed as having AIS (Minto et al., 2003).

However, revelation of molecular mechanisms behind the other potential causes of PA could be possible due to recent advancements like next-generation sequencing (NGS) (Maimoun et al., 2011). A genetic diagnostic rate of 43% has been reported by using a targeted DSD gene panel in a large patient cohort of 278 patients with 46,XY DSD (Eggers et al., 2016). In view of the importance of the genes responsible for DSD, this study was undertaken with an aim to identify the underlying molecular changes by using NGS of a targeted gene panel to investigate a cohort of 25 patients with PA and 46,XY DSD.

Materials and methods

Ethical statement

This study was approved by the Institutional Ethics Committee for research on human patients of the ICMR-National Institute of Immunohaematology, Parel, Mumbai, India.

Clinical assessment of the patients

We have studied 25 patients of PA with 46,XY DSD and no menarche by the age of 13 or later, regardless of the presence of normal growth and development, with or without the appearance of the secondary sexual characters. The patients’ details like age, height, clinical features, breast development, pubic hairs, axillary hairs, hormone profile, and ultrasound details were recorded in the case record sheet. Detailed family histories were collected, and pedigree charts were drawn. Pedigree charts of the patients with novel variants and patients with previously reported mutations are shown in Supplementary Figs S1A, B, C, D, E, F, G, and H and S2A, B, C, D, E, F, G, H, I, J, K, and L, respectively. The study protocols were approved by the Institutional Ethics committee for human patients. Informed consent was obtained from all adult participants and in case of adolescents, the consent was obtained from their parents prior to peripheral blood collection.

Conventional cytogenetics and FISH

The whole peripheral blood collected in heparin vacutainers was cultured (37°C for 72 h) in RPMI 1640 medium (9 ml) (Gibco, New York, USA), supplemented with bovine serum (1 ml) (Gibco), l-glutamine (0.1 ml), phytohemagglutinin (Gibco), and antibiotics (penicillin and streptomycin; Gibco). The cultures were treated with a hypotonic solution (0.075 M KCl), fixed in methanol:acetic acid (3:1), and dropped on pre-chilled slides. The chromosomal preparation was subjected to GTG banding and karyotyped according to International System of Human Cytogenetic Nomenclature (ISCN 2016) (Moorhead et al., 1960; Seabright, 1971).

The chromosomal preparations were passed through a series of alcohol (70%, 80%, and 100%) for 5 min at room temperature and then air dried. The centromeric and locus-specific probes for the X and Y chromosomes (Vysis, Abbott, USA) were used for FISH, and hybridization was carried out using a standard protocol (Davies, 1993). The FISH probes were prepared as per the manufacturer’s instructions and applied. The co-denaturation of the prepared slides was performed at 73°C for 4 min. After co-denaturation, slides were covered with coverslips by applying rubber cement and kept for hybridization at 37°C for 24 h. The post-hybridization washing was performed using wash solution of different stringency (low and high) at 72°C. The slides were counterstained with DAPI (4′,6-diamidino-2-phenylindole, 7.5 μl/slide; Sigma-Aldrich, St. Louis, MO, USA), and analyzed under a fluorescent microscope (Nikon 90i). The images were captured with a CCD camera and analyzed with Applied Spectral Image System (ASI, Israel).

Molecular study

Genomic DNA was extracted from peripheral blood using a human blood DNA extraction kit (Qiagen GmbH, Hilden, Germany) as per the standard protocol. Sanger sequencing was carried out for SRY, AR, SRD5A2, and SF1(NR5A1) genes and primers were designed using Primer 3 software (Supplementary Table SI). The annealing temperatures for the amplification of the SRY, AR, SRD5A2, and SF1 genes are shown in Supplementary Table SI. Sequencing was carried out using an ABI-3130 genetic analyzer, the analysis was performed using Chromas Lite software (Technelysium Pty Ltd, South Brisbane, Australia), and nucleotide sequences were compared with the wild-type DNA sequences.

Next-generation sequencing

NGS was carried out for a DSD gene panel covering 81 DSD-related genes (Supplementary Table SII). We performed targeted exome sequencing at Med-Genome Labs Pvt. Ltd. (Bangalore, India) using the custom capture kit. Sequencing (Illumina platform, San Diego, CA, USA) with a mean coverage of 80–100× were performed. Using the BWA program, the sequences were then aligned to the human reference genome (GRCh38.p13) and analyzed using Picard and GATK version 3.6. The VEP program was used to annotate the gene variants. Furthermore, the clinically relevant mutations were annotated using datasets such as ClinVar, OMIM (update on 20 February 2020), GWAS, HGMD (v2019.4), and SwissVar. Furthermore, the common variants were filtered based on allele frequency in 1000 Genome Phase 3, gnomAD (v2.1), EVS, dbSNP (v151), 1000 Genome, and an internal Indian population database.

Bioinformatics analysis

Primer 3 software (online available) version 4.1.0 was used to design the primers (https://primer3.ut.ee/). SRY, AR, SRD5A2, and SF1 gene references were obtained from the Ensembl genome browser (https://asia.ensembl.org/index.html) and NCBI (https://www.ncbi.nlm.nih.gov/). Sequencing data were analyzed with a reference genome for screening the mutations using the Clustal Omega tool (https://www.ebi.ac.uk/Tools/msa/clustalo/) (Hinxton, Cambridgeshire, UK). The pathogenicity of the novel missense mutations was confirmed using online bioinformatics tools such as Polyphen-2 (http://genetics.bwh.harvard.edu/pph2/), SIFT (http://sift.jcvi.org/), and MutationTaster (www.mutationtaster.org/). The variants were further classified based on ACMG guidelines (Richards et al., 2015).

Protein modeling

The given protein sequences of AR, SRD5A2, DMRT2, and SF1/NR5A1 were considered for model building using different software based on their sequence identity and query coverage. To begin with, the protein sequence of the wild-type SRD5A2 was modeled using a homology modeling-based method on the SWISSMODEL online server (Guex et al., 2009). To understand the impact of the mutation on the protein structure, the mutant protein was also modeled using the same server. DMRT2, AR, and SF1/NR5A1 were modeled using the Robetta online server, which is a deep learning-based method for model building (Kim et al., 2004). All of the generated models were subjected to energy minimization using the GROMACS standalone server (Berendsen et al., 1995), wherein topology files were generated using OPLS AA/L force field. All of the 3D models were subjected to 50 000 steps during energy minimization. Next, these optimized models were subjected to structure validation using ProSA-Web (Wiederstein and Sippl, 2007) and PROCHECK-Ramachandran plot (Laskowski et al., 1993) online servers, for the local energy and phi-psi angle analysis, respectively.

Protein–ligand docking

The modeled SRD5A2 protein was examined for docking with NADPH using Hex standalone software (Ghoorah et al., 2013). The 3D ligand of the NADPH was extracted from PDB ID: 7BW1 (Xiao et al., 2020). Regarding docking, the wild-type SRD5A2 protein was loaded in to HEX software. Within the software, ‘correlation type’ was set as shape + Electro + DARS. Furthermore, ‘Post-Processing’ was DARS minimization. The grid dimension was maintained at 0.6 Å. Finally, post-docking generated 10 poses that were subjected to CHIMERA (Pettersen et al., 2004) for structure visualization and LIGPLOT (Laskowski and Swindells, 2011) for amino acid interaction analysis. Similar steps were followed for the docking of the mutant protein with the NADPH ligand.

Results

Clinical evaluation

The study included 25 unrelated patients with DSD who mainly presented as PA. The common clinical features associated with these cases were absence of secondary sexual characters, ambiguous genitalia, absent Müllerian structures, and gonadal abnormality in the form of either absent ovaries or streak gonads or normal or undescended testis in phenotypic females (Table I). The pedigree analysis of all the patients showed that there was no history of parental consanguinity except for Patients P13 (Supplementary Fig. S2H) and P20 (Supplementary Fig. S1H). None of the parents or siblings had similar complaints to that of the 46,XY DSD patients except for Patients P4 (Supplementary Fig. S1B) and P5 (Supplementary Fig. S2C), where siblings were also affected. The age ranged from 13 to 31 years with a mean age of 19.2 years. The clinical examination revealed normal female external genitalia in 56% of cases and ambiguous genitalia (pseudo-vaginal perineal hypospadias or clitoromegaly) in 44% of cases (Fig. 1). Clinical evaluation of breast development in XY DSD cases with PA showed Tanner stage 1 (48%), stage 2 (20%) and stage 3 (32%), with no patients showing stage 4 or 5. The uterus was found to be absent in 76% of cases. Gonads were found to be testis in 44% of cases, streak gonads in 24% of cases, and absent in 32% of cases (Fig. 1). The levels of FSH and testosterone were recorded and are shown in Table I.

Frequency of clinical features in 46,XY DSD females with PA. External genitalia: normal or ambiguous genitalia. Breast development Tanner staging: B1—no glandular breast tissue palpable; B2—breast bud palpable under the areola (first pubertal sign in females); B3—breast tissue palpable outside areola; no areolar development; B4—areola elevated above the contour of the breast, forming a ‘double scoop’ appearance. Pubic hair development Tanner staging: P1—no hair; P2—downy hair; P3—scant terminal hair; P4—terminal hair that fills the entire triangle overlying the pubic region; uterus—present or absent; gonads: streak, absent, or testis. DSD, disorders of sex development; PA, primary amenorrhea.
Figure 1.

Frequency of clinical features in 46,XY DSD females with PA. External genitalia: normal or ambiguous genitalia. Breast development Tanner staging: B1—no glandular breast tissue palpable; B2—breast bud palpable under the areola (first pubertal sign in females); B3—breast tissue palpable outside areola; no areolar development; B4—areola elevated above the contour of the breast, forming a ‘double scoop’ appearance. Pubic hair development Tanner staging: P1—no hair; P2—downy hair; P3—scant terminal hair; P4—terminal hair that fills the entire triangle overlying the pubic region; uterus—present or absent; gonads: streak, absent, or testis. DSD, disorders of sex development; PA, primary amenorrhea.

Table I

Clinical and cytogenetic features of DSD females with PA.

Patient IDAgeKaryotypeExternal genitaliaSecondary sex characters Tanner stageMüllerian structuresGonadsFSH (mIU/ml)TEST (ng/ml)
P12146,XYNFB3P1A2AbsentOvaries Ab0.97847
P21946,XYNFB2P3A3AbsentOvaries Ab4.07893.9
P31946,XY (90%), 45,X (10%)NFB3P3A3AbsentTestis U Abd4.02NK
P42346,XYNFB3P4A3AbsentOvaries AbNKNK
P51546,XYNFB3P2A2AbsentTestis U SIRNKNK
P61846,XYNFB3P2A1AbsentTestis U AbdNKNK
P73146,XYNFB3P3A3AbsentTestis U AbdNKNK
P82346,XYNFB3P2A2AbsentTestis U SIR6.883492.73
P91446,XYNFB2P1A1AbsentOvaries Ab4.01
P101946,XYAGB2P3A3AbsentTestis Rt D, Lt U11.13.9
P111346,XYAGB1P2A2AbsentTestis U Groin, Lt Ab11.23
P121346,XYAGB1P1A1AbsentTestis U InguinalNK346
P132846,XYAGB1P4A4AbsentTestis DNK4.1
P141546,XYAGB1P2A1AbsentTestis DNKNK
P151546,XYAGB1P2A2AbsentSmall ovaries114NK
P162946,XYNFB2P3A1AbsentOvaries Rt N, Lt Ab83.61NK
P171446,XYAGB1P1A1AbsentOvaries Ab21.426.14
P183146,XYAGB1P3A2AbsentOvaries Ab86.56NK
P191746,XYAGB3P4A4Uterus hypoplasticOvaries Ab106.19NK
P202146,XYAGB2P2A2AbsentOvaries Ab58.33NK
P211446,XYNFB1P1A2Uterus hypoplasticOvaries Ab53.81NK
P221746,XYAGB1P1A1Uterus hypoplasticStreak117NK
P231446,XX (60%), 46,XY (40%)NFB1P1A2Uterus hypoplasticStreak>150NK
P241946,XYNFB1P2A1Uterus hypoplasticStreak63.09NK
P251646,XY (90%), 45,X (10%)NFB1P1A1Uterus hypoplasticStreak27.8NK
Patient IDAgeKaryotypeExternal genitaliaSecondary sex characters Tanner stageMüllerian structuresGonadsFSH (mIU/ml)TEST (ng/ml)
P12146,XYNFB3P1A2AbsentOvaries Ab0.97847
P21946,XYNFB2P3A3AbsentOvaries Ab4.07893.9
P31946,XY (90%), 45,X (10%)NFB3P3A3AbsentTestis U Abd4.02NK
P42346,XYNFB3P4A3AbsentOvaries AbNKNK
P51546,XYNFB3P2A2AbsentTestis U SIRNKNK
P61846,XYNFB3P2A1AbsentTestis U AbdNKNK
P73146,XYNFB3P3A3AbsentTestis U AbdNKNK
P82346,XYNFB3P2A2AbsentTestis U SIR6.883492.73
P91446,XYNFB2P1A1AbsentOvaries Ab4.01
P101946,XYAGB2P3A3AbsentTestis Rt D, Lt U11.13.9
P111346,XYAGB1P2A2AbsentTestis U Groin, Lt Ab11.23
P121346,XYAGB1P1A1AbsentTestis U InguinalNK346
P132846,XYAGB1P4A4AbsentTestis DNK4.1
P141546,XYAGB1P2A1AbsentTestis DNKNK
P151546,XYAGB1P2A2AbsentSmall ovaries114NK
P162946,XYNFB2P3A1AbsentOvaries Rt N, Lt Ab83.61NK
P171446,XYAGB1P1A1AbsentOvaries Ab21.426.14
P183146,XYAGB1P3A2AbsentOvaries Ab86.56NK
P191746,XYAGB3P4A4Uterus hypoplasticOvaries Ab106.19NK
P202146,XYAGB2P2A2AbsentOvaries Ab58.33NK
P211446,XYNFB1P1A2Uterus hypoplasticOvaries Ab53.81NK
P221746,XYAGB1P1A1Uterus hypoplasticStreak117NK
P231446,XX (60%), 46,XY (40%)NFB1P1A2Uterus hypoplasticStreak>150NK
P241946,XYNFB1P2A1Uterus hypoplasticStreak63.09NK
P251646,XY (90%), 45,X (10%)NFB1P1A1Uterus hypoplasticStreak27.8NK

FSH: normal range 2.5–10 mIU/ml. External genitalia: secondary sex characters tanner stage: B: breast, P: pubic hairs, A: axillary hairs. TEST: normal range; females 0.2–0.8 ng/ml and males 4–11 ng/ml. AG, ambiguous genitalia; D, descended; DSD, disorders of sex development; Lt, left; N, normal; NF, normal female; NK, not known; Ovaries Ab, ovaries absent; P, patient id; PA, primary amenorrhea; Rt, right; SIR, superficial inguinal ring; TEST: testosterone; U Abd: undescended abdominal.

Table I

Clinical and cytogenetic features of DSD females with PA.

Patient IDAgeKaryotypeExternal genitaliaSecondary sex characters Tanner stageMüllerian structuresGonadsFSH (mIU/ml)TEST (ng/ml)
P12146,XYNFB3P1A2AbsentOvaries Ab0.97847
P21946,XYNFB2P3A3AbsentOvaries Ab4.07893.9
P31946,XY (90%), 45,X (10%)NFB3P3A3AbsentTestis U Abd4.02NK
P42346,XYNFB3P4A3AbsentOvaries AbNKNK
P51546,XYNFB3P2A2AbsentTestis U SIRNKNK
P61846,XYNFB3P2A1AbsentTestis U AbdNKNK
P73146,XYNFB3P3A3AbsentTestis U AbdNKNK
P82346,XYNFB3P2A2AbsentTestis U SIR6.883492.73
P91446,XYNFB2P1A1AbsentOvaries Ab4.01
P101946,XYAGB2P3A3AbsentTestis Rt D, Lt U11.13.9
P111346,XYAGB1P2A2AbsentTestis U Groin, Lt Ab11.23
P121346,XYAGB1P1A1AbsentTestis U InguinalNK346
P132846,XYAGB1P4A4AbsentTestis DNK4.1
P141546,XYAGB1P2A1AbsentTestis DNKNK
P151546,XYAGB1P2A2AbsentSmall ovaries114NK
P162946,XYNFB2P3A1AbsentOvaries Rt N, Lt Ab83.61NK
P171446,XYAGB1P1A1AbsentOvaries Ab21.426.14
P183146,XYAGB1P3A2AbsentOvaries Ab86.56NK
P191746,XYAGB3P4A4Uterus hypoplasticOvaries Ab106.19NK
P202146,XYAGB2P2A2AbsentOvaries Ab58.33NK
P211446,XYNFB1P1A2Uterus hypoplasticOvaries Ab53.81NK
P221746,XYAGB1P1A1Uterus hypoplasticStreak117NK
P231446,XX (60%), 46,XY (40%)NFB1P1A2Uterus hypoplasticStreak>150NK
P241946,XYNFB1P2A1Uterus hypoplasticStreak63.09NK
P251646,XY (90%), 45,X (10%)NFB1P1A1Uterus hypoplasticStreak27.8NK
Patient IDAgeKaryotypeExternal genitaliaSecondary sex characters Tanner stageMüllerian structuresGonadsFSH (mIU/ml)TEST (ng/ml)
P12146,XYNFB3P1A2AbsentOvaries Ab0.97847
P21946,XYNFB2P3A3AbsentOvaries Ab4.07893.9
P31946,XY (90%), 45,X (10%)NFB3P3A3AbsentTestis U Abd4.02NK
P42346,XYNFB3P4A3AbsentOvaries AbNKNK
P51546,XYNFB3P2A2AbsentTestis U SIRNKNK
P61846,XYNFB3P2A1AbsentTestis U AbdNKNK
P73146,XYNFB3P3A3AbsentTestis U AbdNKNK
P82346,XYNFB3P2A2AbsentTestis U SIR6.883492.73
P91446,XYNFB2P1A1AbsentOvaries Ab4.01
P101946,XYAGB2P3A3AbsentTestis Rt D, Lt U11.13.9
P111346,XYAGB1P2A2AbsentTestis U Groin, Lt Ab11.23
P121346,XYAGB1P1A1AbsentTestis U InguinalNK346
P132846,XYAGB1P4A4AbsentTestis DNK4.1
P141546,XYAGB1P2A1AbsentTestis DNKNK
P151546,XYAGB1P2A2AbsentSmall ovaries114NK
P162946,XYNFB2P3A1AbsentOvaries Rt N, Lt Ab83.61NK
P171446,XYAGB1P1A1AbsentOvaries Ab21.426.14
P183146,XYAGB1P3A2AbsentOvaries Ab86.56NK
P191746,XYAGB3P4A4Uterus hypoplasticOvaries Ab106.19NK
P202146,XYAGB2P2A2AbsentOvaries Ab58.33NK
P211446,XYNFB1P1A2Uterus hypoplasticOvaries Ab53.81NK
P221746,XYAGB1P1A1Uterus hypoplasticStreak117NK
P231446,XX (60%), 46,XY (40%)NFB1P1A2Uterus hypoplasticStreak>150NK
P241946,XYNFB1P2A1Uterus hypoplasticStreak63.09NK
P251646,XY (90%), 45,X (10%)NFB1P1A1Uterus hypoplasticStreak27.8NK

FSH: normal range 2.5–10 mIU/ml. External genitalia: secondary sex characters tanner stage: B: breast, P: pubic hairs, A: axillary hairs. TEST: normal range; females 0.2–0.8 ng/ml and males 4–11 ng/ml. AG, ambiguous genitalia; D, descended; DSD, disorders of sex development; Lt, left; N, normal; NF, normal female; NK, not known; Ovaries Ab, ovaries absent; P, patient id; PA, primary amenorrhea; Rt, right; SIR, superficial inguinal ring; TEST: testosterone; U Abd: undescended abdominal.

Cytogenetic analysis

Chromosomal analysis revealed a 46,XY karyotype in each of the 25 patients. A FISH study showed a XY karyotype in 22 cases and mosaic patterns for 2 cases with X/XY [46,XY (90%), 45,X (10%)] and 1 case with XX/XY [46,XX (60%), 46,XY (40%)].

Molecular study

The targeted sequencing identified a total of 21 mutations in 20 out of 25 patients with DSD. The overall incidence was found to be 84%. The most frequently identified mutations in our series were in the AR gene (nine patients, 36%), followed by SRD5A2 (five patients, 20%), SF1 (three patients, 12%), DHX37 (one patient, 4%), HSD17B3 (one patient, 4%), and DMRT2 genes (one patient, 4%) (Fig. 2). In our study, we found that 8 of the 21 mutations (38.09%) were novel variants, of which three were in the AR gene, two were in the SRD5A2 gene, and one each was in the SF1, DMRT2 and HSD17B3 genes.

Different gene mutations in 25 DSD patients with PA. The frequency of gene mutations: AR (36%), SRD5A2 (20%), SF1 (12%), DHX37 (4%), HSD17B3 (4%), DMRT2 (4%), and no mutation in (20%) patients. DSD, disorders of sex development; PA, primary amenorrhea.
Figure 2.

Different gene mutations in 25 DSD patients with PA. The frequency of gene mutations: AR (36%), SRD5A2 (20%), SF1 (12%), DHX37 (4%), HSD17B3 (4%), DMRT2 (4%), and no mutation in (20%) patients. DSD, disorders of sex development; PA, primary amenorrhea.

In the present study, we found three mutations, c.175C>T (p.Gln59Ter), c.1443C>G (p.Tyr481Ter), and c.1567G>T, (p.Glu523Ter) in Exon 1 of the AR gene, two missense mutations, c.1742A>C (p.Lys581Thr) and c.1762G>C (p.Ala588Pro) in Exon 2 of the AR gene, and three mutations, c.2226G>A (p.Trp742Ter), c.2255G>A (p.Trp752Ter), and c.2301del (p.Asp768IlefsTer21) in Exon 5 of the AR gene, while one patient had a missense mutation c.2323C>T (p.Arg775Cys) in Exon 6 of the AR gene (Table II).

Table II

Spectrum of gene mutations identified in AR, SRD5A2, SF1, DHX37, DMRT2, and HSD17B3 genes in 46,XY DSD subjects.

Patient IDMutationProtein changeExon/mutation siteType of mutationPolyphen2SIFTMutationTaster 2/LRTACMG guidelinesNovel/reportedReferences for reported mutationClinVar accession numbers
AR gene mutations
P1c.175C>Tp.Gln59TerExon 1/NTDHemizygousDamaging by MutationTaster 2PathogenicReportedHolterhus et al. (2005) and Zoppi et al. (1993)SCV001960994
Non-sense
P2c.1443C>Gp.Tyr481TerExon 1/NTDHemizygousDamaging by MutationTaster 2PathogenicReportedPhilibert et al. (2010)SCV001960995
Non-sense
P3c.1567G>Tp.Glu523TerExon 1/NTDHemizygousDamaging by MutationTaster 2PathogenicNovelSCV001960990
Non-sense
P4c.1742A>Cp.Lys581ThrExon 2/DBDHemizygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTLikely PathogenicNovelSCV001960993
Missense
P5c.1762G>Cp.Ala588ProExon 2/DBDHemizygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTLikely PathogenicReportedLek et al. (2018) and Stenson et al. (2017)SCV001960996
Missense
P6c.2226G>Ap.Trp742TerExon 5/LBDHemizygousDamaging by MutationTaster 2PathogenicNovelSCV002097635
Non-sense
P7c.2255G>Ap.Trp752TerExon 5/LBDHemizygousDamaging by MutationTaster 2PathogenicReportedGottlieb et al. (2012)SCV001960997
Non-sense
P8c.2301delp.Asp768IlefsTer21Exon 5/LBDHemizygousDamaging by MutationTaster 2PathogenicReportedGottlieb et al. (2012)SCV001960998
Frameshift
P9c.2323C>Tp.Arg775CysExon 6/LBDHemizygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTPathogenic by ClinvarReportedAudí et al. (2010), Gottlieb et al. (2012), Akcay et al. (2014), and Wang et al. (2017)SCV001960999
Missense
SRD5A2 gene mutations
P10c.169G>Tp.Glu57TerExon 1HomozygousDamaging by MutationTaster 2PathogenicReportedShabir et al. (2015)SCV001960989
Non-sense
P11c.81_94delp.Ala28LeufsTer103Exon 1HomozygousDamaging by MutationTaster 2PathogenicNovelSCV001960987
Frame shift
P12c.169G>Tp.Glu57TerExon 1Non-senseDamaging by MutationTaster 2PathogenicReportedShabir et al. (2015)SCV001960988
c.691C>Tp.His231TyrExon 4CompoundProbably damagingDamaging by SIFTLikely PathogenicNovel
heterozygous
Missense
P13c.589G>Ap.Glu197LysExon 4HomozygousProbably damagingDamaging by SIFTPathogenicReportedGui et al. (2019)SCV001960991
Missense
P14c.737G>Ap.Arg246GlnExon 5HomozygousProbably damagingDamaging by SIFTLikely PathogenicReportedYang et al. (2012), Shabir et al. (2015), and Cheng et al. (2015)SCV001960992
Missense
SF1 gene mutations
P15c.19G>Tp.Glu7TerExon 2/DBDHeterozygousDamaging by MutationTaster 2PathogenicReportedFabbri-Scallet et al. (2018)SCV002097636
Nonsense
P16c.97T>Cp.Cys33ArgExon 2/DBDHeterozygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTLikely PathogenicNovelSCV002097637
Missense
P17c.250C>Tp.Arg84CysExon 4/DBDHeterozygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTLikely PathogenicReportedFabbri-Scallet et al. (2018)SCV001961002
Missense
DHX37 gene mutation
P18c.1877C>Tp.Ser626LeuExon 15HeterozygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTLikely PathogenicReportedMcElreavey et al. (2020)SCV001961000
Missense
DMRT2 gene mutation
P19c.71A>Cp.Glu24AlaExon 3HeterozygousDamaging by SIFTUncertain SignificanceNovelSCV001961001
Missense
HSDB17 gene mutation
P20c.385 + 5G>AIntron 4/5′Splice-siteDamaging by MutationTaster 2Uncertain SignificanceNovelSCV002097638
Patient IDMutationProtein changeExon/mutation siteType of mutationPolyphen2SIFTMutationTaster 2/LRTACMG guidelinesNovel/reportedReferences for reported mutationClinVar accession numbers
AR gene mutations
P1c.175C>Tp.Gln59TerExon 1/NTDHemizygousDamaging by MutationTaster 2PathogenicReportedHolterhus et al. (2005) and Zoppi et al. (1993)SCV001960994
Non-sense
P2c.1443C>Gp.Tyr481TerExon 1/NTDHemizygousDamaging by MutationTaster 2PathogenicReportedPhilibert et al. (2010)SCV001960995
Non-sense
P3c.1567G>Tp.Glu523TerExon 1/NTDHemizygousDamaging by MutationTaster 2PathogenicNovelSCV001960990
Non-sense
P4c.1742A>Cp.Lys581ThrExon 2/DBDHemizygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTLikely PathogenicNovelSCV001960993
Missense
P5c.1762G>Cp.Ala588ProExon 2/DBDHemizygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTLikely PathogenicReportedLek et al. (2018) and Stenson et al. (2017)SCV001960996
Missense
P6c.2226G>Ap.Trp742TerExon 5/LBDHemizygousDamaging by MutationTaster 2PathogenicNovelSCV002097635
Non-sense
P7c.2255G>Ap.Trp752TerExon 5/LBDHemizygousDamaging by MutationTaster 2PathogenicReportedGottlieb et al. (2012)SCV001960997
Non-sense
P8c.2301delp.Asp768IlefsTer21Exon 5/LBDHemizygousDamaging by MutationTaster 2PathogenicReportedGottlieb et al. (2012)SCV001960998
Frameshift
P9c.2323C>Tp.Arg775CysExon 6/LBDHemizygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTPathogenic by ClinvarReportedAudí et al. (2010), Gottlieb et al. (2012), Akcay et al. (2014), and Wang et al. (2017)SCV001960999
Missense
SRD5A2 gene mutations
P10c.169G>Tp.Glu57TerExon 1HomozygousDamaging by MutationTaster 2PathogenicReportedShabir et al. (2015)SCV001960989
Non-sense
P11c.81_94delp.Ala28LeufsTer103Exon 1HomozygousDamaging by MutationTaster 2PathogenicNovelSCV001960987
Frame shift
P12c.169G>Tp.Glu57TerExon 1Non-senseDamaging by MutationTaster 2PathogenicReportedShabir et al. (2015)SCV001960988
c.691C>Tp.His231TyrExon 4CompoundProbably damagingDamaging by SIFTLikely PathogenicNovel
heterozygous
Missense
P13c.589G>Ap.Glu197LysExon 4HomozygousProbably damagingDamaging by SIFTPathogenicReportedGui et al. (2019)SCV001960991
Missense
P14c.737G>Ap.Arg246GlnExon 5HomozygousProbably damagingDamaging by SIFTLikely PathogenicReportedYang et al. (2012), Shabir et al. (2015), and Cheng et al. (2015)SCV001960992
Missense
SF1 gene mutations
P15c.19G>Tp.Glu7TerExon 2/DBDHeterozygousDamaging by MutationTaster 2PathogenicReportedFabbri-Scallet et al. (2018)SCV002097636
Nonsense
P16c.97T>Cp.Cys33ArgExon 2/DBDHeterozygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTLikely PathogenicNovelSCV002097637
Missense
P17c.250C>Tp.Arg84CysExon 4/DBDHeterozygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTLikely PathogenicReportedFabbri-Scallet et al. (2018)SCV001961002
Missense
DHX37 gene mutation
P18c.1877C>Tp.Ser626LeuExon 15HeterozygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTLikely PathogenicReportedMcElreavey et al. (2020)SCV001961000
Missense
DMRT2 gene mutation
P19c.71A>Cp.Glu24AlaExon 3HeterozygousDamaging by SIFTUncertain SignificanceNovelSCV001961001
Missense
HSDB17 gene mutation
P20c.385 + 5G>AIntron 4/5′Splice-siteDamaging by MutationTaster 2Uncertain SignificanceNovelSCV002097638

DSD, disorders of sex development.

Table II

Spectrum of gene mutations identified in AR, SRD5A2, SF1, DHX37, DMRT2, and HSD17B3 genes in 46,XY DSD subjects.

Patient IDMutationProtein changeExon/mutation siteType of mutationPolyphen2SIFTMutationTaster 2/LRTACMG guidelinesNovel/reportedReferences for reported mutationClinVar accession numbers
AR gene mutations
P1c.175C>Tp.Gln59TerExon 1/NTDHemizygousDamaging by MutationTaster 2PathogenicReportedHolterhus et al. (2005) and Zoppi et al. (1993)SCV001960994
Non-sense
P2c.1443C>Gp.Tyr481TerExon 1/NTDHemizygousDamaging by MutationTaster 2PathogenicReportedPhilibert et al. (2010)SCV001960995
Non-sense
P3c.1567G>Tp.Glu523TerExon 1/NTDHemizygousDamaging by MutationTaster 2PathogenicNovelSCV001960990
Non-sense
P4c.1742A>Cp.Lys581ThrExon 2/DBDHemizygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTLikely PathogenicNovelSCV001960993
Missense
P5c.1762G>Cp.Ala588ProExon 2/DBDHemizygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTLikely PathogenicReportedLek et al. (2018) and Stenson et al. (2017)SCV001960996
Missense
P6c.2226G>Ap.Trp742TerExon 5/LBDHemizygousDamaging by MutationTaster 2PathogenicNovelSCV002097635
Non-sense
P7c.2255G>Ap.Trp752TerExon 5/LBDHemizygousDamaging by MutationTaster 2PathogenicReportedGottlieb et al. (2012)SCV001960997
Non-sense
P8c.2301delp.Asp768IlefsTer21Exon 5/LBDHemizygousDamaging by MutationTaster 2PathogenicReportedGottlieb et al. (2012)SCV001960998
Frameshift
P9c.2323C>Tp.Arg775CysExon 6/LBDHemizygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTPathogenic by ClinvarReportedAudí et al. (2010), Gottlieb et al. (2012), Akcay et al. (2014), and Wang et al. (2017)SCV001960999
Missense
SRD5A2 gene mutations
P10c.169G>Tp.Glu57TerExon 1HomozygousDamaging by MutationTaster 2PathogenicReportedShabir et al. (2015)SCV001960989
Non-sense
P11c.81_94delp.Ala28LeufsTer103Exon 1HomozygousDamaging by MutationTaster 2PathogenicNovelSCV001960987
Frame shift
P12c.169G>Tp.Glu57TerExon 1Non-senseDamaging by MutationTaster 2PathogenicReportedShabir et al. (2015)SCV001960988
c.691C>Tp.His231TyrExon 4CompoundProbably damagingDamaging by SIFTLikely PathogenicNovel
heterozygous
Missense
P13c.589G>Ap.Glu197LysExon 4HomozygousProbably damagingDamaging by SIFTPathogenicReportedGui et al. (2019)SCV001960991
Missense
P14c.737G>Ap.Arg246GlnExon 5HomozygousProbably damagingDamaging by SIFTLikely PathogenicReportedYang et al. (2012), Shabir et al. (2015), and Cheng et al. (2015)SCV001960992
Missense
SF1 gene mutations
P15c.19G>Tp.Glu7TerExon 2/DBDHeterozygousDamaging by MutationTaster 2PathogenicReportedFabbri-Scallet et al. (2018)SCV002097636
Nonsense
P16c.97T>Cp.Cys33ArgExon 2/DBDHeterozygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTLikely PathogenicNovelSCV002097637
Missense
P17c.250C>Tp.Arg84CysExon 4/DBDHeterozygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTLikely PathogenicReportedFabbri-Scallet et al. (2018)SCV001961002
Missense
DHX37 gene mutation
P18c.1877C>Tp.Ser626LeuExon 15HeterozygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTLikely PathogenicReportedMcElreavey et al. (2020)SCV001961000
Missense
DMRT2 gene mutation
P19c.71A>Cp.Glu24AlaExon 3HeterozygousDamaging by SIFTUncertain SignificanceNovelSCV001961001
Missense
HSDB17 gene mutation
P20c.385 + 5G>AIntron 4/5′Splice-siteDamaging by MutationTaster 2Uncertain SignificanceNovelSCV002097638
Patient IDMutationProtein changeExon/mutation siteType of mutationPolyphen2SIFTMutationTaster 2/LRTACMG guidelinesNovel/reportedReferences for reported mutationClinVar accession numbers
AR gene mutations
P1c.175C>Tp.Gln59TerExon 1/NTDHemizygousDamaging by MutationTaster 2PathogenicReportedHolterhus et al. (2005) and Zoppi et al. (1993)SCV001960994
Non-sense
P2c.1443C>Gp.Tyr481TerExon 1/NTDHemizygousDamaging by MutationTaster 2PathogenicReportedPhilibert et al. (2010)SCV001960995
Non-sense
P3c.1567G>Tp.Glu523TerExon 1/NTDHemizygousDamaging by MutationTaster 2PathogenicNovelSCV001960990
Non-sense
P4c.1742A>Cp.Lys581ThrExon 2/DBDHemizygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTLikely PathogenicNovelSCV001960993
Missense
P5c.1762G>Cp.Ala588ProExon 2/DBDHemizygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTLikely PathogenicReportedLek et al. (2018) and Stenson et al. (2017)SCV001960996
Missense
P6c.2226G>Ap.Trp742TerExon 5/LBDHemizygousDamaging by MutationTaster 2PathogenicNovelSCV002097635
Non-sense
P7c.2255G>Ap.Trp752TerExon 5/LBDHemizygousDamaging by MutationTaster 2PathogenicReportedGottlieb et al. (2012)SCV001960997
Non-sense
P8c.2301delp.Asp768IlefsTer21Exon 5/LBDHemizygousDamaging by MutationTaster 2PathogenicReportedGottlieb et al. (2012)SCV001960998
Frameshift
P9c.2323C>Tp.Arg775CysExon 6/LBDHemizygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTPathogenic by ClinvarReportedAudí et al. (2010), Gottlieb et al. (2012), Akcay et al. (2014), and Wang et al. (2017)SCV001960999
Missense
SRD5A2 gene mutations
P10c.169G>Tp.Glu57TerExon 1HomozygousDamaging by MutationTaster 2PathogenicReportedShabir et al. (2015)SCV001960989
Non-sense
P11c.81_94delp.Ala28LeufsTer103Exon 1HomozygousDamaging by MutationTaster 2PathogenicNovelSCV001960987
Frame shift
P12c.169G>Tp.Glu57TerExon 1Non-senseDamaging by MutationTaster 2PathogenicReportedShabir et al. (2015)SCV001960988
c.691C>Tp.His231TyrExon 4CompoundProbably damagingDamaging by SIFTLikely PathogenicNovel
heterozygous
Missense
P13c.589G>Ap.Glu197LysExon 4HomozygousProbably damagingDamaging by SIFTPathogenicReportedGui et al. (2019)SCV001960991
Missense
P14c.737G>Ap.Arg246GlnExon 5HomozygousProbably damagingDamaging by SIFTLikely PathogenicReportedYang et al. (2012), Shabir et al. (2015), and Cheng et al. (2015)SCV001960992
Missense
SF1 gene mutations
P15c.19G>Tp.Glu7TerExon 2/DBDHeterozygousDamaging by MutationTaster 2PathogenicReportedFabbri-Scallet et al. (2018)SCV002097636
Nonsense
P16c.97T>Cp.Cys33ArgExon 2/DBDHeterozygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTLikely PathogenicNovelSCV002097637
Missense
P17c.250C>Tp.Arg84CysExon 4/DBDHeterozygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTLikely PathogenicReportedFabbri-Scallet et al. (2018)SCV001961002
Missense
DHX37 gene mutation
P18c.1877C>Tp.Ser626LeuExon 15HeterozygousProbably damagingDamaging by SIFT, MutationTaster 2, LRTLikely PathogenicReportedMcElreavey et al. (2020)SCV001961000
Missense
DMRT2 gene mutation
P19c.71A>Cp.Glu24AlaExon 3HeterozygousDamaging by SIFTUncertain SignificanceNovelSCV001961001
Missense
HSDB17 gene mutation
P20c.385 + 5G>AIntron 4/5′Splice-siteDamaging by MutationTaster 2Uncertain SignificanceNovelSCV002097638

DSD, disorders of sex development.

The SRD5A2 gene was found to be the second most frequently mutated gene in our study associated with DSD. A total of six mutations in SRD5A2 were identified in five patients. Out of these, three were in Exon 1, two were in Exon 4, and one was in Exon 5. The mutation c.169G>T (p.Glu57Ter) was identified in two cases. The mutations c.81_94del, c.589G>A (p.Glu197Lys), c.691C>T (p.His231Tyr) and c.737G>A (p.Arg246Gln) were identified in one case each (Table II).

Three mutations were identified in the SF1 (NR5A1) gene; c.19G>T (p.Glu7Ter), c.97T>C (p.Cys33Arg) in Exon 2 and c.250C>T (p.Arg84Cys) in Exon 4.

One mutation was found in each of the DHX37, DMRT2, and HSD17B3 genes; c.1877C>T (p.Ser626Leu), c.71A>C (p.Glu24Ala), and c.385 + 5G>A, respectively (Table II).

The electropherograms of Sanger sequencing of all the novel and reported mutations are shown in Fig. 3 and Supplementary Figs S3 and S4, respectively.

Electropherograms of novel mutations in SRD5A2, AR, SF1, DMRT2, and HSD17B3 genes in our cohort of 46,XY DSD females with PA. Electropherograms showing: (A) a missense mutation in the SRD5A2 gene at Exon 4 c.691C>T; (B) a frame shift mutation in the SRD5A2 gene at Exon 1 c.81_94del; (C) a nonsense mutation in the AR gene at Exon 1 c.1567 G>T; (D) a missense mutation in the AR gene at Exon 2 c.1742A>C; (E) a nonsense mutation in the AR gene at Exon 5 c.2226G>A; (F) a missense mutation in the SF1 gene at Exon 2 c.97T>C; (G) a missense mutation in the DMRT2 gene at Exon 3 c.71A>C; and (H) a spice-site mutation in the HSD17B3 gene at Intron 4/5' splice-site c.385 + 5G>A. DSD, disorders of sex development; PA, primary amenorrhea.
Figure 3.

Electropherograms of novel mutations in SRD5A2, AR, SF1, DMRT2, and HSD17B3 genes in our cohort of 46,XY DSD females with PA. Electropherograms showing: (A) a missense mutation in the SRD5A2 gene at Exon 4 c.691C>T; (B) a frame shift mutation in the SRD5A2 gene at Exon 1 c.81_94del; (C) a nonsense mutation in the AR gene at Exon 1 c.1567 G>T; (D) a missense mutation in the AR gene at Exon 2 c.1742A>C; (E) a nonsense mutation in the AR gene at Exon 5 c.2226G>A; (F) a missense mutation in the SF1 gene at Exon 2 c.97T>C; (G) a missense mutation in the DMRT2 gene at Exon 3 c.71A>C; and (H) a spice-site mutation in the HSD17B3 gene at Intron 4/5' splice-site c.385 + 5G>A. DSD, disorders of sex development; PA, primary amenorrhea.

Protein modeling of the novel variants

DMRT2 p.Glu24Ala

The 3D structure of the DMRT2 protein was generated using the Robetta server and considered for hydrogen bond analysis using CHIMERA standalone software. In particular, Glu24 of the wild-type DMRT2 protein shows a salt bridge formation with Arg31 (Fig. 4A). However, after the p.Glu24Ala substitution, because the methyl side chain of alanine is non-reactive, there is no salt bridge formation with Arg31 (Fig. 4B). Thus, the mutation results in the loss of salt bridge formation, which could certainly have a local impact on the flexibility of the loop region of the protein. A literature review confirmed that any substitution of charged residue with alanine has an adverse effect on protein stability (Pakula et al., 1986; Cunningham and Wells, 1989; Yang et al., 2009).

Protein modeling of DMRT2, AR, SF1/NR5A1, and SRD5A2. (A) The wild-type DMRT2 protein in ribbon format showing salt bridge formation between Glu24 and Arg31; (B) the mutant-type DMRT2 protein substitution of Glu with Ala results in the loss of salt bridge formation; (C) AR protein showing the region of truncation after the formation of stop codon at position Glu523 (orange), Trp742 (cyan), and Trp752 (blue); (D) AR protein wild-type form showing two interactions; (E) AR protein mutant form, due to the Lys581Thr substitution, showing three interactions; (F) SF1/NR5A1 wild type shown in ribbon format with Cys33 forming a disulphide bridge with Cys30 and also, Cys33 interacts with the hydrophobic Phe37 amino acid; (G) SF1/NR5A1 mutant-type Arg33 after its substitution interacts with Cys30, Val15, and Phe37; (H) SRD5A2 protein after the frame shift; (I) Wild-type SRD5A2 docked with NADP; and (J) Mutant SRD5A2 type docked with NADP.
Figure 4.

Protein modeling of DMRT2, AR, SF1/NR5A1, and SRD5A2. (A) The wild-type DMRT2 protein in ribbon format showing salt bridge formation between Glu24 and Arg31; (B) the mutant-type DMRT2 protein substitution of Glu with Ala results in the loss of salt bridge formation; (C) AR protein showing the region of truncation after the formation of stop codon at position Glu523 (orange), Trp742 (cyan), and Trp752 (blue); (D) AR protein wild-type form showing two interactions; (E) AR protein mutant form, due to the Lys581Thr substitution, showing three interactions; (F) SF1/NR5A1 wild type shown in ribbon format with Cys33 forming a disulphide bridge with Cys30 and also, Cys33 interacts with the hydrophobic Phe37 amino acid; (G) SF1/NR5A1 mutant-type Arg33 after its substitution interacts with Cys30, Val15, and Phe37; (H) SRD5A2 protein after the frame shift; (I) Wild-type SRD5A2 docked with NADP; and (J) Mutant SRD5A2 type docked with NADP.

AR p.Glu523Ter, AR p.Trp742Ter

The 3D model of the protein sequence of AR showed potential termination residues at positions like 523, 742, and 752. Position 523 harbors glutamic acid whereas Positions 742 and 752 have tryptophans. Mutations at Positions 523, 742, and 752 introducing stop codons could terminate transcription of the ligand-binding domain (LBD). Thus, truncation removes the ligand-binding helical secondary structure, which is critically involved in receptor dimerization (Xiao et al., 2020) (Fig. 4C).

AR p.Lys581Thr

Another 3D model of the protein sequence of AR in wild-type form showed two interactions at Lys581 where there was a single bond formation with Cys 577 and Lys585 (Fig. 4D). After the mutation substituting threonine, there are three interactions observed wherein Thr581 is showing two interactions with Cys577 and a single interaction with Lys585 (Fig. 4E). This mutation could also be associated with the loss of an androgen binding site.

SF1 p.Cys33Arg

The generated wild-type model of SF1/NR5A1 has cysteine at Position 33. As a result, it forms a disulphide bridge with Cys30 and an additional interaction with Phe37. After the substitution of Cys with Arg, there is a loss of disulphide bridge formation (Fig. 4F and G). However, the hydrogen bond formation is still observed for Cys33 with Phe37, Cys30, and Val15.

SRD5A2 p.Ala28LeufsTer103

The given protein sequence of SRD5A2 was submitted to the SWISSMODEL server. Based on the template search, PDB ID: 7BW1 was listed as a potential 3D structure with a query coverage of 96% and sequence identity of 99.6%. This template was used for the modeling of the wild type and the mutant SRD5A2 protein. The SRD5A2 gene has a reported mutation wherein Ala28 is substituted by Leu due to the deletion of bases between 81 and 94. As a result, there is a frameshift from amino acid Position 29 and formation of a stop codon at Position 103, which results in a premature protein truncation. The frame shift changes the transmembrane and beta-strand region formation after the 29th position (Fig. 4H).

Furthermore, the structure validation report of all the modeled structures from PROCHECK and ProSA-Web report confirmed energetically stable structures.

SRD5A2 p.His231Tyr

Protein–ligand docking showed that NADP was docked within the active site pocket based on the template of 7BW1. When wild-type SRD5A2 was docked with NADP, the binding energy was −730.10 kcal/mol. The mutant with NADP showed a binding energy of −717.16 kcal/mol. Thus, a lower binding energy was observed after the substitution of histidine with tyrosine. Moreover, the histidine residue at Position 231 showcased hydrogen bond formation with NADP. However, after mutation, this hydrogen bonding ceases as per their hydrogen bond report (Fig. 4I and J). Based on the LIGPLOT report, there is an absence of hydrogen bonding between Tyr and NADP. Instead, there is an external bond formation observed with a light blue color (Fig. 5A and B).

SRD5A2 protein: LIGPLOT report. LIGPLOT reports of SRD5A2 protein. (A) Wild-type SRD5A2 in association with NADP and (B) mutant SRD5A2 in association with NADP.
Figure 5.

SRD5A2 protein: LIGPLOT report. LIGPLOT reports of SRD5A2 protein. (A) Wild-type SRD5A2 in association with NADP and (B) mutant SRD5A2 in association with NADP.

HSDB17 intron 4′5 splice site

The HSDB17 mutation was not investigated by protein modeling as it created a change in one of the mRNA splice sites.

Discussion

DSD are rare congenital disorders, where there is discordance between chromosomal, gonadal, and the phenotypic sex of an individual. Although DSD are a major disorder of pediatric concern, many cases remain undiagnosed until puberty and the majority present as PA at the time of puberty (Mendonca et al., 2009). This study included a total of 25 patients, presumed to be females, with 46,XY DSD, who presented with PA with or without the development of secondary sexual characters and varying degrees of virilization. To the best of our knowledge, this is the first Indian study on such a large number of cases. We have systematically analyzed the genotypes of the cases and related them with their phenotypes, and we note the extreme variability in the clinical spectrum and genetic background of these cases. In the present study, we found that the phenotype of the DSD patients with PA is mainly female external genitalia in 56% of patients, while ambiguous genitalia in the form of pseudo-vaginal perineal hypospadias or clitoromegaly was also observed in 44% of cases, making the exact clinical diagnosis of these conditions difficult. Clinical evaluation of internal genitalia in these cases showed that majority of these cases (76%) were without Müllerian structures, and many cases (44%) had testes as gonads. The most common cause of the PA with DSD in this cohort was found to be AIS (36%) followed by steroid 5 alpha-reductase deficiency (25%), both of which result in defective androgen activity, and this was similarly found in a previous report (Costagliola et al., 2021). The cytogenetic investigations play an important role in establishing the diagnosis of DSD and should be advised immediately to those patients who are suspected to be having clinical features of 46,XY DSD. This is important, as post-pubertal complete AIS patients are prone to tumor development in undescended abdominal testes (Kathrins and Kolon, 2016). As per the literature, the estimated risk of malignancy is 15–50% in gonadal dysgenesis and <1–15% in disorders of androgen synthesis or action (Wisniewski et al., 2019). The 46,XY karyotype in females with PA and DSD have been reported with various frequencies (2–31%) in different populations (Ghosh et al., 2018). In the present study, in all the 25 DSD patients, the karyotype was found to be 46,XY with no additional chromosomal abnormalities. The chromosomal abnormalities using a combination of conventional G-banded cytogenetics and FISH revealed that 88% of the cases with PA (22/25) had a pure XY karyotype and the remaining 12% (3/25) had a mosaic karyotype. The frequencies of chromosomal abnormalities identified in our study were similar to the frequencies reported in the literature (García-Acero et al., 2020). Hence, the combination of cytogenetics tools is important in the accurate diagnosis of the sex abnormalities.

Various studies have been carried out in the past using targeted NGS, where mutations have been reported with different frequencies, commonly in AR, SRD5A2, SF1, DHX37, HSD17B3, SRY, DMRT1, DMRT2, and other DSD-related genes with low frequency (Wang et al., 2018; Costagliola et al., 2021; Yu et al., 2021). Our study is the single largest study, where we have evaluated molecular genetic characterization of PA cases with DSD, providing 80% of these cases with an exact molecular diagnosis with a targeted NGS approach among the Indian population. Reaching a specific diagnosis is important, not only to know the pattern of inheritance and chances of other family members being affected, but also for identifying the associated features and to know the risk of tumors in the long term (Achermann et al., 2015). The overall incidence of mutations in the AR, SRD5A2, and SF1 genes in our study was 72% (18/25) and was found to be similar as that of the study reported by Yu et al. (78.3%, 47/60). We have identified 8 novel variants out of the 21 variants reported (38.1%) in our study, which is similar to a frequency (41.7%, 25/60) previously reported in the literature (Yu et al., 2021). Our extensive literature search revealed that AR gene mutations ranging from a single nucleotide variation to complete gene deletion including intronic mutations (Brinkmann et al., 1996; Boehmer et al., 2001) are responsible for variable phenotypes in 46,XY DSD patients, leading to either complete or partial AIS (Nagaraja et al., 2019). The androgen receptor gene is present at the chromosomal location Xq11-12 and contains a total of eight exons. The AR gene encodes a protein of 919 amino acids, which is a member of the nuclear receptor super family and comprises an N-terminal domain (NTD) (Exon 1), a DNA-binding domain (DBD) (Exons 2 and 3), a hinge region (Exons 3 and 4), and an LBD (Exons 4–8). A defective AR protein can cause androgen insensitivity, known as AIS (OMIM # 300068) (Audí et al., 2010). We have identified nine cases (36%) with AR gene mutations, which confirms the diagnosis as AIS in these patients. So far, 32 different mutations in the AR gene have been reported in the Indian population (Nagaraja et al., 2019). Our single-centre study will increase these data with nine mutations in AIS patients. Out of the 32 mutations, most were found in the LBD region (14 mutations) and the least were found in the Hinge region (two mutations). Similarly, we also found most (4/9) of the mutations to be in the LBD region. In a study by Yu et al. (2021), the frequency of AR gene mutations was found to be 35% (21/60). We also identified the AR gene as the most common mutated (42.8%, 9/21) gene in our Indian population. In the present study, three novel variants in the AR gene were identified. The hemizygous nonsense variant c.1567G>T (p.Glu523Ter), found in Exon 1 of the AR gene, results in a stop codon and causes the premature truncation of the protein at Codon 523. Exon 1 of the AR gene encodes the NTD, and it mediates the transactivation function, the disruption of which compromises the biological function of the AR gene (Saranya et al., 2016). In silico prediction of the variant was found to be damaging by MutationTaster2 and it was found to be a pathogenic variant as per the ACMG guidelines. The second novel variant observed in our study was the missense hemizygous mutation c.1742A>C (p.Lys581Thr) leading to the replacement of the amino acid lysine with threonine at Position 581 in Exon 2 of the AR gene. Exons 2 and 3 of the AR gene encode the DBD (residues 539–628), which contain two peculiar zinc finger modules necessary to bind the DNA and stabilize the protein (Gulía et al., 2018). Many studies have reported a relatively high number of mutations that can damage the structure and function of this domain, causing AIS. This mutation was found to be likely pathogenic and in silico prediction of the variant was probably damaging by polyphen-2 and damaging by SIFT. Another novel variant observed was the nonsense mutation c.2226G>A (p.Trp742Ter) in Exon 5 replacing tryptophan at the amino acid Position 742 by a stop codon, leading to the premature truncation of protein. This mutation was found to be pathogenic and damaging by MutationTaster2. Exon 5 of the AR gene encodes the LBD of the protein. Receptor dimerization of the AR protein can occur due to strong interactions between the LBDs and weak interactions between DBDs (Brinkmann et al., 1996).

The second gene associated with 46,XY DSD in our study was the steroid 5 alpha-reductase 2 gene (SRD5A2) (cytogenetic location: 2p23.1). This gene consists of five exons and is responsible for conversion of testosterone to its active form DHT. Mutations in the SRD5A2 gene lead to pseudo-vaginal perineoscrotal hypospadias (OMIM # 264600) (Andersson et al., 1991), a rare autosomal-recessive (both the parents are carriers) disorder. We found that 6 out of 21 (28.6%) mutations were in the SRD5A2 gene and these were the second most frequent in our cohort. In a review by Nagaraja et al. (2019), a total of 19 mutations were in the SRD5A2 gene from various studies among the Indian population and they were found to be the second most frequently reported among the genes causing 46,XY DSD. Yu et al. (2021) also reported SRD5A2 gene mutations to be the second most common mutation (21.7%, 13/60) in an Asian population. In our study, we found three mutations in Exon 1, two in Exon 4, and one in Exon 5. Maimoun et al. (2011) reported a predominance of mutations in Exon 1 (35.8%), Exon 4 (21.7%), and Exon 3 (11.3%), while Exon 5 (9.4%) was found to be in relatively preserved regions. Various Asian studies have revealed that the majority of mutations were confined to Exons 1 and 5 of the SRD5A2 gene (Thigpen et al., 1992; Nie et al., 2011; Shabir et al., 2015), and this was similarly observed in Indian studies, where most of the mutations were observed in Exons 1 and 5, indicating that these exons are the hot spot regions of the gene (Eunice et al., 2008; Sahu et al., 2009; Nagaraja et al., 2010). Furthermore, we have identified two novel variants in the SRD5A2 gene. The novel homozygous mutation found in Exon 1 of the SRD5A2 gene was c.81_94del, with substitution of the amino acid alanine at Position 28, with leucine causing a frameshift from amino acid Position 29 and formation of a stop codon at Position 103, which resulted in the premature truncation of the protein. A mutated enzyme of 103 residues is therefore formed instead of a protein of 245 amino acids. This variant was found to be pathogenic as per ACMG guidelines. Point mutations leading to the formation of premature stop codons have been described in previous studies in patients with 5 alpha-reductase Type 2 deficiency (Hiort et al., 2002; Vilchis et al., 2008), but there are very few cases reported with mutations in the coding region of the SRD5A2 gene leading to premature termination of the enzyme. The mutations in Exon 1 of SRD5A2 were observed in different ethnic and geographical backgrounds (Mazen et al., 2003; Maimoun et al., 2011), but there are very few reports described with the presence of heterozygous mutations (Thigpen et al., 1992).

Another novel mutation observed in our study is the compound heterozygous mutation c.691C>T in Exon 4 that replaces the amino acid histidine with tyrosine at amino acid Position 231. This mutation was found to be likely pathogenic as per the ACMG guidelines and was predicted to be probably damaging by Polyphen 2. The positively charged histidine at Position 231 shows hydrogen bond formation with NADP. However, after mutation, this hydrogen bonding ceases, as per the CHIMERA report. Based on the LIGPLOT report, similar to the CHIMERA report, the hydrogen bonding was absent between the tyrosine and NADP (Fig. 5A and B). Instead, there is an external bond which may disrupt the cofactor binding, which is quite evident from the mutation.

The other gene that is responsible for 46,XY DSD is SF1 (Steroidogenic Factor 1) (OMIM # 184757), located at 9q33, also known as the NR5A1 gene. It is an orphan nuclear receptor that consists of 461 amino acids and is structurally similar to the other nuclear receptors (Oba et al., 1996; Wong et al., 1996). It encodes a nuclear transcription factor regulating the expression of several genes that participate in sexual development. In humans, heterozygous SF1 mutations in XY individuals lead to adrenal and gonadal failure (Köhler and Achermann, 2010), cryptorchidism (Wada et al., 2005), micro-penis and infertility (Bashamboo et al., 2010). Fabbri-Scallet et al. (2018) reviewed 238 published cases of both 46,XX and XY DSD and observed SF1 gene mutations at a frequency of 12% (25/205) in 46,XY DSD cases, with a predominance of mutations in the DBD (35%) and LBD (42.3%) domains. The frequency of SF1 gene mutations leading to DSD in our study was found to be 14.3% (3/21), similar to the study by Fabbri-Scallet et al. We found three heterozygous mutations in the DBD encoding region of SF1: two in Exon 2 and one in Exon 4. Nagaraja et al. (2019) reviewed four cases of SF1 gene mutations in gonadal dysgenesis among the Indian population, out of which three were reported by in male patients and one was reported by Paliwal et al. (2011) in a female patient. We identified one novel variant c.97T>C (p.Cys33Arg), a likely pathogenic, missense mutation in the DBD region of the protein. In silico prediction of this variant is probably damaging by polyphen-2 and damaging by SIFT. Fabbri-Scarlet et al. (2018) reported a mutation on same position; however, the amino acid change reported was a serine instead of a cysteine.

The less frequently identified gene mutations observed in our study were in one patient each with DMRT2, DHX37, and HSD17B3 (4.76%, 1/21) mutations and diagnosis of 46,XY DSD. Previous studies also reported these mutations as less frequent mutations among 46,XY DSD females (Yu et al., 2021). We have identified a novel heterozygous missense variant in Exon 3 of the DMRT2 gene in a female with PA and it was not reported previously in 46,XY DSD cases. This variant resulted in the amino acid substitution of alanine for glutamic acid (E) at Codon 24 (p.Glu24Ala). In silico prediction of the variant was damaging by SIFT. Although this variant is classified as a variant of uncertain significance as per ACMG guidelines, parental testing is essential and may change the classification. Unfortunately, we could not perform the parental sequencing, as the parents were not available for the molecular study. The DMRT2 gene (encoding doublesex and mab-3-related transcription factor 2) has six exons and is believed to be involved in gonadal development. To date, only eight clinically significant DMRT2 gene mutations have been reported from ClinVar and Humasvar with links to dbSNP (https://www.genecards.org/).

The DEAH-box RNA helicase (DHX37), essential for ribosome biogenesis, is found specifically in association with 46,XY gonadal dysgenesis (OMIM # 273250) and 46,XY testicular regression syndrome (TRS). The pathogenic mutations in the DHX37 gene (OMIM # 617362) are responsible for phenotypes ranging from phenotypic females to males with bilateral or unilateral undescended testis. The exact prevalence of TRS is not known but it has been found to affect ∼1:2000 boys (Pirgon and Dündar, 2012). In our study, we found one patient with a heterozygous missense variation, c.1877C>T, in Exon 15 of the DHX37 gene that resulted in the amino acid substitution of leucine for serine at Codon 626 (p.Ser626Leu). A similar variation was reported by McElreavey et al. (2020) in a male baby with TRS.

Another novel intronic splice-site variant, c.385 + 5G>A, was found in the HSD17B3 gene in a single patient with DSD and PA in our study. Homozygous or compound heterozygous mutations in the HSD17B3 gene lead to 17-beta hydroxysteroid dehydrogenase III (17β-HSD3) deficiency (OMIM # 264300). In the largest DSD European database, patients with 17β-HSD3 deficiency represent ∼4% of total 46,XY DSD patients (Hughes, 2008). Mendonca et al. (2017), found 37 mutations in the HSD17B3 gene reported in the literature and they were found to be missense mutations, nonsense mutations, exonic deletions, duplications, intronic splice sites, and amplification mutations. This disorder is characterized by hypoplastic to normal internal genitalia (epididymis, vas deferens, seminal vesicles, and ejaculatory ducts) with female external genitalia and the absence of a prostate.

Genotype–phenotype correlation in patients with DSD and clinical relevance

It is difficult to establish a clear genotype-phenotype correlation in patients with 46,XY DSD as there are multiple genes involved in the etiopathogenesis of DSD (Maimoun et al., 2011). Those females who present with PA and karyotype 46,XY are phenotypically female since the abnormal gonadal tissue in these cases fails to produce Müllerian inhibiting factor and testosterone and, even if they produce testosterone, there is a defective androgen receptor for its action or it fails to be converted into the active form, DHT. The genotype–phenotype correlation was assessed in 25 unrelated patients with 46,XY DSD and presenting with PA with molecular defects in AR, SRD5A2, SF1, DHX37, DMRT2, and HSD17B3 genes (Tables I and II). All the patients from P1 to P9 with a molecular diagnosis of AR gene mutations were phenotypic females with 46,XY karyotypes (except Patient P3 who had a mosaic karyotype), with variable degrees of Tanner staging of breast, pubic and axillary hair development, absent uterus and cryptorchidism of one or both gonads. As all the patients with AR gene mutations, including the novel variants, have normal female phenotypes and presented at the time of puberty with PA, the diagnosis of partial androgen insensitivity syndrome (PAIS) is less likely. We were not able to establish the inheritance pattern of novel and reported mutations as the parents were not available or refused to participate in the study and this is the limitation of the present study. In Patients P1, P2, P4, and P9, testicular tissue was not located with ultrasound findings, needing repeat evaluation or confirmation by MRI. Elevated levels of testosterone were observed in three patients (P1, P2, and P8), which is typically seen in AIS. In the other six cases, although testosterone levels were not known to define a specific clinical diagnosis, molecular evaluation confirmed the diagnosis of AIS in these cases. The management of such cases should include creation of a functional vagina and removal of cryptorchid gonads as these may lead to development of malignancy. Gonadectomy can be delayed in those with AIS as gonadal tumors rarely occur before puberty in such patients and the limited pubertal development results from endogenous gonadal hormone production rather than exogenous hormone treatment (Speroff and Fritz, 2005). The cryptorchid testis has a relatively high incidence of neoplasia; therefore, a gonadectomy is suggested before exogenous hormone treatment. The prevalence of a germ cell tumor is 15% in PAIS and the risk increases after puberty, reaching up to 33% after the age of 50 (Araujo-Melendez et al., 2020). In our study, nine patients had AIS, and of these, only one patient, P7 who was a 31-year-old female, was diagnosed with gonadal malignancy. While evaluating the clinical details of the other patients, malignancy was not seen, which could be associated with younger age at the time of presentation; these patients need to be followed up to determine their status of malignancy.

All the patients with SRD5A2 gene mutations (Patients P10–P14) had a male karyotype, ambiguous genitalia, and variable degrees of under-virilization ranging from pseudo-vaginal hypospadias to clitoromegaly and variable positions of either one or both the gonads. Müllerian structures were found to be absent in all the patients. Although the ratios of testosterone to DHT were not available for the exact clinical diagnosis of these cases due to insufficient data, NGS was found to be important in confirming the diagnosis of steroid 5 alpha-reductase deficiency.

As far as the SF1 gene is concerned, all three cases showed a 46,XY karyotype; two of these cases, P15 and P17, had ambiguous genitalia, and one case, P16, had female external genitalia. The secondary sexual characters were typically underdeveloped with variable Tanner staging of development. The uterus was found to be absent in all three cases and either or both the ovaries were found to be small or absent in Patients P15 and P16. In Patient P17, a gonadectomy revealed the gonads to be testicular tissue. All the three patients with SF1 gene mutations had elevated levels of FSH and were categorized as hypergonadotropic hypogonadism.

A clinical diagnosis was difficult in Patients P18, P19, and P20, as all of them had a male karyotype, ambiguous genitalia, and absent uterus and ovaries with elevated levels of FSH. These patients were identified with mutations in the DHX37, DMRT2, and HSD17B3 genes, respectively, providing them with an exact molecular diagnosis, which will help in genetic counseling (Berra et al., 2011). Gonadal tumors occur in up to 25% of women with a Y chromosome; unlike cases of complete AIS, their gonads do not secrete hormones and should therefore be removed at the time of diagnosis. Although the management of patients with DSD is difficult due to the complex molecular mechanisms, the evaluation of tumor risk can be aided by the advances in genotyping for Y-chromosomal material, which is not evident in traditional karyotyping. Thus, the benefits of reaching a specific diagnosis are not only important for understanding the tumor risk but for identifying associated features, inheritance, chances of other family members being affected, and an understanding of the natural history of a condition. In our cohort of 25 patients, we could not find any mutation in the DSD-related genes in five patients (20%, 5/25) due to complex molecular mechanisms in 46,XY DSD; this suggests new DSD genes that are yet to be discovered in these disorders. Even though there are already known genes in DSD patients, it is difficult to establish the genotype–phenotype co-relation (Wang et al., 2017). Furthermore, the pathogenicity of novel variants was determined by multiple variant effect prediction tools and protein modeling and docking studies; however, in vitro functional studies are important to ascertain the pathogenicity of these variants, which can be considered as the limitation of this study.

Conclusion

The present study concludes that adolescent females with DSD and presenting with PA and absent secondary sexual characters should be investigated for chromosomal abnormalities along with routine hormonal and ultrasound investigations. Cytogenetically confirmed 46,XY adolescent females should be investigated for genetic mutations at a molecular level to arrive at a specific diagnosis, as the clinical signs and symptoms are found to be variable in these cases. Targeted NGS using a gene panel related to sex development will help in identifying an exact molecular diagnosis in these patients and should be preferred over routine Sanger sequencing. Our study highlights the importance of molecular diagnosis and genetic counseling of DSD patients, which will ultimately help in appropriate gender assignment of these cases.

Supplementary data

Supplementary data are available at Molecular Human Reproduction online.

Data availability

All data generated or analyzed during this study are included in this article and its supplementary information files. The novel variants identified have been reported to ClinVar in a two-spreadsheet submission (Accession no: SCV001960987–SCV001961002 and SCV002097635–SCV002097638).

Acknowledgments

We are grateful to ICMR-NIIH for providing financial support. We thank all the patients for participating in our study and the technical staff of Department of Cytogenetics, ICMR-NIIH.

Authors’ roles

V.K., S.K.C., S.D., and B.R.V. contributed to the study conception and design, carried out the experiments, and analyzed the data. V.K. performed the clinical assessment, diagnosis of patients, and experimental work. S.K.C. performed the bioinformatic analysis and in silico validation. V.K., S.D., and J.G. collected and analyzed the data and carried out the validation work. V.K. and S.K.C. prepared the article draft. B.R.V. edited and finalized the article. All the authors read and approved the article before submission.

Funding

This study was funded by the Indian Council of Medical Research through an Intramural grant provided to ICMR-NIIH (Intramural/ICMR-NIIH/2018-19).

Conflict of interest

The authors declare that there are no conflicts of interest in connection with this article.

References

Achermann
JC
,
Domenice
S
,
Bachega
TA
,
Nishi
MY
,
Mendonca
BB.
Disorders of sex development: effect of molecular diagnostics
.
Nat Rev Endocrinol
2015
;
11
:
478
488
.

Akcay
T
,
Fernandez‐Cancio
M
,
Turan
SE
,
Güran
,
Audi
L
,
Bereket
AB.
AR and SRD 5A2 gene mutations in a series of 51 Turkish 46, XY DSD children with a clinical diagnosis of androgen insensitivity
.
Andrology
2014
;
2
:
572
578
.

Andersson
S
,
Berman
DM
,
Jenkins
EP
,
Russell
DW.
Deletion of steroid 5α-reductase 2 gene in male pseudohermaphroditism
.
Nature
1991
;
354
:
159
161
.

Araujo-Melendez
M
,
Verduzco-Aguirre
H
,
Morales
JJ
,
Martinez-Benitez
B
,
Castillejos-Molina
R
,
Fuentes
A
,
Salama
M
,
Bourlon
MT.
Disorders of sex development and malignant germ cell tumors
.
Oncology (Williston Park, NY)
2020
;
34
:
421
426
.

Audí
L
,
Fernández-Cancio
M
,
Carrascosa
A
,
Andaluz
P
,
Torán
N
,
Piró
C
,
Vilaró
E
,
Vicens-Calvet
E
,
Gussinyé
M
,
Albisu
MA
et al. ;
Grupo de Apoyo al Síndrome de Insensibilidad a los Andrógenos (GrApSIA)
.
Novel (60%) and recurrent (40%) androgen receptor gene mutations in a series of 59 patients with a 46, XY disorder of sex development
.
J Clin Endocrinol Metab
2010
;
95
:
1876
1888
.

Bashamboo
A
,
Ferraz-de-Souza
B
,
Lourenço
D
,
Lin
L
,
Sebire
NJ
,
Montjean
D
,
Bignon-Topalovic
J
,
Mandelbaum
J
,
Siffroi
JP
,
Christin-Maitre
S
et al.
Human male infertility associated with mutations in NR5A1 encoding steroidogenic factor 1
.
Am J Hum Genet
2010
;
87
:
505
512
.

Berendsen
HJ
,
van der Spoel
D
,
van Drunen
R.
GROMACS: a message-passing parallel molecular dynamics implementation
.
Comput Phys Commun
1995
;
91
:
43
56
.

Berra
M
,
Williams
EL
,
Muroni
B
,
Creighton
SM
,
Honour
JW
,
Rumsby
G
,
Conway
GS.
Recognition of 5a-reductase-2 deficiency in an adult female 46XY DSD clinic
.
Eur J Endocrinol
2011
;
164
:
1019
1025
.

Boehmer
AL
,
Brinkmann
O
,
Brüggenwirth
H
,
van Assendelft
C
,
Otten
BJ
,
Verleun-Mooijman
MC
,
Niermeijer
MF
,
Brunner
HG
,
Rouwé
CW
,
Waelkens
JJ
et al.
Genotype versus phenotype in families with androgen insensitivity syndrome
.
J Clin Endocrinol Metab
2001
;
86
:
4151
4160
.

Brinkmann
A
,
Jenster
G
,
Ris-Stalpers
C
,
van der Korput
H
,
Brüggenwirth
H
,
Boehmer
A
,
Trapman
J.
Molecular basis of androgen insensitivity
.
Steroids
1996
;
61
:
172
175
.

Buonocore
F
,
Clifford-Mobley
O
,
King
TFJ
,
Striglioni
N
,
Man
E
,
Suntharalingham
JP
,
Del Valle
I
,
Lin
L
,
Lagos
CF
,
Rumsby
G
et al.
Next-generation sequencing reveals novel genetic variants (SRY, DMRT1, NR5A1, DHH, DHX37) in adults with 46, XY DSD
.
J Endocr Soc
2019
;
3
:
2341
2360
.

Cheng
J
,
Lin
R
,
Zhang
W
,
Liu
G
,
Sheng
H
,
Li
X
,
Zhou
Z
,
Mao
X
,
Liu
L.
Phenotype and molecular characteristics in 45 Chinese children with 5α‐reductase type 2 deficiency from South China
.
Clin Endocrinol (Oxf)
2015
;
83
:
518
526
.

Costagliola
G
,
Cosci O di Coscio
M
,
Masini
B
,
Baldinotti
F
,
Caligo
MA
,
Tyutyusheva
N
,
Sessa
MR
,
Peroni
D
,
Bertelloni
S.
Disorders of sexual development with XY karyotype and female phenotype: clinical findings and genetic background in a cohort from a single centre
.
J Endocrinol Invest
2021
;
44
:
145
151
.

Cunningham
BC
,
Wells
JA.
High-resolution epitope mapping of hGH-receptor interactions by alanine-scanning mutagenesis
.
Science
1989
;
244
:
1081
1085
.

Davies
KE
(ed).
Human Genetic Disease Analysis: A Practical Approach
. Oxford, UK: Oxford University Press, Vol. 124, 1993. 

Eggers
S
,
Sadedin
S
,
Van Den Bergen
JA
,
Robevska
G
,
Ohnesorg
T
,
Hewitt
J
,
Lambeth
L
,
Bouty
A
,
Knarston
IM
,
Tan
TY
et al.
Disorders of sex development: insights from targeted gene sequencing of a large international patient cohort
.
Genome Biol
2016
;
17
:
1
21
.

Eunice
M
,
Philibert
P
,
Kulshreshtha
B
,
Audran
F
,
Paris
F
,
Khurana
ML
,
Pulikkanath
PE
,
Kucheria
K
,
Sultan
C
,
Ammini
AC.
Molecular diagnosis of 5α‐reductase‐2 gene mutation in two Indian families with male pseudohermaphroditism
.
Asian J Androl
2008
;
10
:
815
818
.

Fabbri‐Scallet
H
,
de Mello
MP
,
Guerra‐Júnior
G
,
Maciel‐Guerra
AT
,
de Andrade
JG
,
de Queiroz
CM
,
Monlleó
IL
,
Struve
D
,
Hiort
O
,
Werner
R.
Functional characterization of five NR5A1 gene mutations found in patients with 46, XY disorders of sex development
.
Hum Mutat
2018
;
39
:
114
123
.

García-Acero
M
,
Moreno-Niño
O
,
Suárez-Obando
F
,
Molina
M
,
Manotas
MC
,
Prieto
JC
,
Forero
C
,
Céspedes
C
,
Pérez
J
,
Fernandez
N
et al.
Disorders of sex development: genetic characterization of a patient cohort
.
Mol Med Rep
2020
;
21
:
97
106
.

Ghoorah
AW
,
Devignes
MD
,
Smaïl‐Tabbone
M
,
Ritchie
DW.
Protein docking using case‐based reasoning
.
Proteins
2013
;
81
:
2150
2158
.

Ghosh
S
,
Roy
S
,
Pal
P
,
Dutta
A
,
Halder
A.
Cytogenetic analysis of patients with primary amenorrhea in Eastern India
.
J Obstet Gynaecol
2018
;
38
:
270
275
.

Gottlieb
B
,
Beitel
LK
,
Nadarajah
A
,
Paliouras
M
,
Trifiro
M.
The androgen receptor gene mutations database: 2012 update
.
Hum Mutat.
2012
;
33
:
887
894
.

Guex
N
,
Peitsch
MC
,
Schwede
T.
Automated comparative protein structure modeling with SWISS‐MODEL and Swiss‐PdbViewer: a historical perspective
.
Electrophoresis
2009
;
30
:
S162
S173
.

Gui
B
,
Song
Y
,
Su
Z
,
Luo
F-H
,
Chen
L
,
Wang
X
,
Chen
R
,
Yang
Y
,
Wang
J
,
Zhao
X
et al.
New insights into 5α-reductase type 2 deficiency based on a multi-centre study: regional distribution and genotype–phenotype profiling of SRD5A2 in 190 Chinese patients
.
J Med Genet
2019
;
56
:
685
692
.

Gulía
C
,
Baldassarra
S
,
Zangari
A
,
Briganti
V
,
Gigli
S
,
Gaffi
M
,
Signore
F
,
Vallone
C
,
Nucciotti
R
,
Costantini
FM
et al.
Androgen insensitivity syndrome
.
Eur Rev Med Pharmacol Sci
2018
;
22
:
3873
3887
.

Hiort
O
,
Schütt
SM
,
Bals‐Pratsch
M
,
Holterhus
PM
,
Marschke
C
,
Struve
D.
A novel homozygous disruptive mutation in the SRD5A2‐gene in a partially virilized patient with 5α‐reductase deficiency
.
Int J Androl
2002
;
25
:
55
58
.

Holterhus
PM
,
Werner
R
,
Hoppe
U
,
Bassler
J
,
Korsch
E
,
Ranke
MB
,
Dörr
HG
,
Hiort
O.
Molecular features and clinical phenotypes in androgen insensitivity syndrome in the absence and presence of androgen receptor gene mutations
.
J Mol Med (Berl)
2005
;
83
:
1005
1013
.

Hughes
IA.
Disorders of sex development: a new definition and classification
.
Best Pract Res Clin Endocrinol Metab
2008
;
22
:
119
134
.

Kathrins
M
,
Kolon
TF.
Malignancy in disorders of sex development
.
Transl Androl Urol
2016
;
5
:
794
798
.

Kim
DE
,
Chivian
D
,
Baker
D.
Protein structure prediction and analysis using the Robetta server
.
Nucleic Acids Res
2004
;
32
:
W526
531
.

Köhler
B
,
Achermann
JC.
Update–steroidogenic factor 1 (SF-1, NR5A1)
.
Minerva Endocrinol
2010
;
35
:
73
86
.

Laskowski
RA
,
MacArthur
MW
,
Moss
DS
,
Thornton
JM.
PROCHECK: a program to check the stereochemical quality of protein structures
.
J Appl Crystallogr
1993
;
26
:
283
291
.

Laskowski
RA
,
Swindells
MB.
LigPlot+: multiple ligand-protein interaction diagrams for drug discovery
.
J Chem Inf Model
2011
;
51
:
2778
2786
.

Lek
N
,
Tadokoro-Cuccaro
R
,
Whitchurch
JB
,
Mazumder
B
,
Miles
H
,
Prentice
P
,
Bunch
T
,
Zielińska
K
,
Metzler
V
,
Mongan
NP
et al.
Predicting puberty in partial androgen insensitivity syndrome: use of clinical and functional androgen receptor indices
.
EBioMedicine
2018
;
36
:
401
409
.

MacLaughlin
DT
,
Donahoe
PK.
Sex determination and differentiation
.
N Engl J Med
2004
;
350
:
367
378
.

Maimoun
L
,
Philibert
P
,
Bouchard
P
,
Öcal
G
,
Leheup
B
,
Fenichel
P
,
Servant
N
,
Paris
F
,
Sultan
C.
Primary amenorrhea in four adolescents revealed 5α-reductase deficiency confirmed by molecular analysis
.
Fertil Steril
2011
;
95
:
804.e1-804
e5
.

Mazen
I
,
Gad
YZ
,
Hafez
M
,
Sultan
C
,
Lumbroso
S.
Molecular analysis of 5α‐reductase type 2 gene in eight unrelated Egyptian children with suspected 5α‐reductase deficiency: prevalence of the G34R mutation
.
Clin Endocrinol (Oxf)
2003
;
58
:
627
631
.

McElreavey
K
,
Jorgensen
A
,
Eozenou
C
,
Merel
T
,
Bignon-Topalovic
J
,
Tan
DS
,
Houzelstein
D
,
Buonocore
F
,
Warr
N
,
Kay
RG
et al.
Pathogenic variants in the DEAH-box RNA helicase DHX37 are a frequent cause of 46, XY gonadal dysgenesis and 46, XY testicular regression syndrome
.
Genet Med
2020
;
22
:
150
159
.

Mendonca
BB
,
Domenice
S
,
Arnhold
IJ
,
Costa
EM.
46,XY disorders of sex development (DSD)
.
Clin Endocrinol (Oxf)
2009
;
70
:
173
187
.

Mendonca
BB
,
Gomes
NL
,
Costa
EM
,
Inacio
M
,
Martin
RM
,
Nishi
MY
,
Carvalho
FM
,
Tibor
FD
,
Domenice
S.
46,XY disorder of sex development (DSD) due to 17β-hydroxysteroid dehydrogenase type 3 deficiency
.
J Steroid Biochem Mol Biol
2017
;
165
:
79
85
.

Minto
CL
,
Liao
KL
,
Conway
GS
,
Creighton
SM.
Sexual function in women with complete androgen insensitivity syndrome
.
Fertil Steril
2003
;
80
:
157
164
.

Moorhead
PS
,
Nowell
PC
,
Mellman
WJ
,
Battips
DT
,
Hungerford
DA.
Chromosome preparations of leukocytes cultured from human peripheral blood
.
Exp Cell Res
1960
;
20
:
613
616
.

Nagaraja
MR
,
Rastogi
A
,
Raman
R
,
Gupta
DK
,
Singh
SK.
Molecular diagnosis of 46, XY DSD and identification of a novel 8 nucleotide deletion in exon 1 of the SRD5A2 gene
.
J Pediatric Endocrinol Metab
2010
;
23
:
379
385
.

Nagaraja
MR
,
Gubbala
SP
,
Delphine Silvia
CW
,
Amanchy
R.
Molecular diagnostics of disorders of sexual development: an Indian survey and systems biology perspective
.
Syst Biol Reprod Med
2019
;
65
:
105
120
.

Nie
M
,
Zhou
Q
,
Mao
J
,
Lu
S
,
Wu
X.
Five novel mutations of SRD5A2 found in eight Chinese patients with 46, XY disorders of sex development
.
Mol Hum Reprod
2011
;
17
:
57
62
.

Oba
K
,
Yanase
T
,
Nomura
M
,
Morohashi
KI
,
Takayanagi
R
,
Nawata
H.
Structural characterization of HumanAd4bp (SF-1) gene
.
Biochem Biophys Res Commun
1996
;
226
:
261
267
.

Pakula
AA
,
Young
VB
,
Sauer
RT.
Bacteriophage lambda cro mutations: effects on activity and intracellular degradation
.
Proc Natl Acad Sci U S A
1986
;
83
:
8829
8833
.

Paliwal
P
,
Sharma
A
,
Birla
S
,
Kriplani
A
,
Khadgawat
R
,
Sharma
A.
Identification of novel SRY mutations and SF1 (NR5A1) changes in patients with pure gonadal dysgenesis and 46, XY karyotype
.
Mol Hum Reprod
2011
;
17
:
372
378
.

Pettersen
EF
,
Goddard
TD
,
Huang
CC
,
Couch
GS
,
Greenblatt
DM
,
Meng
EC
,
Ferrin
TE.
UCSF Chimera—a visualization system for exploratory research and analysis
.
J Comput Chem
2004
;
25
:
1605
1612
.

Philibert
P
,
Audran
F
,
Pienkowski
C
,
Morange
I
,
Kohler
B
,
Flori
E
,
Heinrich
C
,
Dacou-Voutetakis
C
,
Joseph
MG
,
Guedj
AM
et al.
Complete androgen insensitivity syndrome is frequently due to premature stop codons in exon 1 of the androgen receptor gene: an international collaborative report of 13 new mutations
.
Fertil Steril
2010
;
94
:
472
476
.

Pirgon
Ö
,
Dündar
BN.
Vanishing testes: a literature review
.
J Clin Res Pediatr Endocrinol
2012
;
4
:
116
120
.

Practice Committee of the American Society for Reproductive Medicine
.
Current evaluation of amenorrhea
.
Fertil Steril
2008
;
90
:
219
225
.

Richards
S
,
Aziz
N
,
Bale
S
,
Bick
D
,
Das
S
,
Gastier-Foster
J
,
Grody
WW
,
Hegde
M
,
Lyon
E
,
Spector
E
et al. ;
ACMG Laboratory Quality Assurance Committee
.
Standards and guidelines for the interpretation of sequence variants: a joint consensus recommendation of the American College of Medical Genetics and Genomics and the Association for Molecular Pathology
.
Genet Med
2015
;
17
:
405
424
.

Sahu
R
,
Boddula
R
,
Sharma
P
,
Bhatia
V
,
Greaves
R
,
Rao
S
,
Desai
M
,
Wakhlu
A
,
Phadke
S
,
Shukla
M
et al.
Genetic analysis of the SRD5A2 gene in Indian patients with 5α-reductase deficiency
.
J Pediatr Endocrinol Metab
2009
;
22
:
247
254
.

Saranya
B
,
Bhavani
G
,
Arumugam
B
,
Jayashankar
M
,
Santhiya
ST.
Three novel and two known androgen receptor gene mutations associated with androgen insensitivity syndrome in sex-reversed XY female patients
.
J Genet
2016
;
95
:
911
921
.

Seabright
M.
A rapid banding technique for human chromosomes
.
Lancet
1971
;
2
:
971
972
.

Shabir
I
,
Khurana
ML
,
Joseph
AA
,
Eunice
M
,
Mehta
M
,
Ammini
AC.
Phenotype, genotype and gender identity in a large cohort of patients from India with 5α‐reductase 2 deficiency
.
Andrology
2015
;
3
:
1132
1139
.

Speroff
L
,
Fritz
MA.
Clinical Gynecologic Endocrinology and Infertility
.
Philadelphia, PA
:
Lippincott Williams & Wilkins
,
2005
.

Stenson
PD
,
Mort
M
,
Ball
EV
,
Evans
K
,
Hayden
M
,
Heywood
S
,
Hussain
M
,
Phillips
AD
,
Cooper
DN.
The Human Gene Mutation Database: towards a comprehensive repository of inherited mutation data for medical research, genetic diagnosis and next-generation sequencing studies
.
Hum Genet
2017
;
136
:
665
677
.

Strauss
JF
,
Barbieri
RL
,
Gargiulo
AR
. Yen & Jaffe’s Reproductive Endocrinology: Physiology, Pathophysiology, and Clinical Management: 8th Edition. Amsterdam, The Netherlands: Elsevier Inc., 2017.

Thigpen
AE
,
Davis
DL
,
Milatovich
A
,
Mendonca
BB
,
Imperato-McGinley
J
,
Griffin
JE
,
Francke
U
,
Wilson
JD
,
Russell
DW.
Molecular genetics of steroid 5 alpha-reductase 2 deficiency
.
J Clin Invest
1992
;
90
:
799
809
.

Vilchis
F
,
Valdez
E
,
Ramos
L
,
García
R
,
Gómez
R
,
Chávez
B.
Novel compound heterozygous mutations in the SRD5A2 gene from 46, XY infants with ambiguous external genitalia
.
J Hum Genet
2008
;
53
:
401
406
.

Wada
Y
,
Okada
M
,
Hasegawa
T
,
Ogata
T.
Association of severe micropenis with Gly146Ala polymorphism in the gene for steroidogenic factor-1
.
Endocr J
2005
;
52
:
445
448
.

Wang
H
,
Zhang
L
,
Wang
N
,
Zhu
H
,
Han
B
,
Sun
F
,
Yao
H
,
Zhang
Q
,
Zhu
W
,
Cheng
T
et al.
Next-generation sequencing reveals genetic landscape in 46, XY disorders of sexual development patients with variable phenotypes
.
Hum Genet
2018
;
137
:
265
277
.

Wang
Y
,
Gong
C
,
Wang
X
,
Qin
M.
AR mutations in 28 patients with androgen insensitivity syndrome (Prader grade 0–3)
.
Sci China Life Sci
2017
;
60
:
700
706
.

Wiederstein
M
,
Sippl
MJ.
ProSA-web: interactive web service for the recognition of errors in three-dimensional structures of proteins
.
Nucleic Acids Res
2007
;
35
:
W407
410
.

Wisniewski
AB
,
Batista
RL
,
Costa
EM
,
Finlayson
C
,
Sircili
MH
,
Dénes
FT
,
Domenice
S
,
Mendonca
BB.
Management of 46, XY differences/disorders of sex development (DSD) throughout life
.
Endocr Rev
2019
;
40
:
1547
1572
.

Wong
M
,
Ramayya
MS
,
Chrousos
GP
,
Driggers
PH
,
Parker
KL.
Cloning and sequence analysis of the human gene encoding steroidogenic factor 1
.
J Mol Endocrinol
1996
;
17
:
139
147
.

Xiao
Q
,
Wang
L
,
Supekar
S
,
Shen
T
,
Liu
H
,
Ye
F
,
Huang
J
,
Fan
H
,
Wei
Z
,
Zhang
C.
Structure of human steroid 5α-reductase 2 with the anti-androgen drug finasteride
.
Nat Commun
2020
;
11
:
1
1
.

Yang
YF
,
Cheng
KC
,
Tsai
PH
,
Liu
CC
,
Lee
TR
,
Lyu
PC.
Alanine substitutions of noncysteine residues in the cysteine‐stabilized αβ motif
.
Protein Sci
2009
;
18
:
1498
1506
.

Yang
Y
,
Wang
BA
,
Guo
QH
,
Dou
JT
,
Lv
ZH
,
Ba
JM
,
Lu
JM
,
Pan
CY
,
Mu
YM.
Clinical and genetic analysis of three Chinese patients with steroid 5α-reductase type 2 deficiency
.
J Pediatr Endocrinol Metab
2012
;
25
:
1077
1082
.

Yu
BQ
,
Liu
ZX
,
Gao
YJ
,
Wang
X
,
Mao
JF
,
Nie
M
,
Wu
XY.
Prevalence of gene mutations in a Chinese 46, XY disorders of sex development cohort detected by targeted next-generation sequencing
.
Asian J Androl
2021
;
23
:
69
.

Zoppi
S
,
Wilson
CM
,
Harbison
MD
,
Griffin
JE
,
Wilson
JD
,
McPhaul
MJ
,
Marcelli
M.
Complete testicular feminization caused by an amino-terminal truncation of the androgen receptor with downstream initiation
.
J Clin Invest
1993
;
91
:
1105
1112
.

This article is published and distributed under the terms of the Oxford University Press, Standard Journals Publication Model (https://dbpia.nl.go.kr/pages/standard-publication-reuse-rights)

Supplementary data