Skip to main content
Archives of Disease in Childhood logoLink to Archives of Disease in Childhood
. 2006 Apr 19;91(7):554–563. doi: 10.1136/adc.2006.098319

Consensus statement on management of intersex disorders

I A Hughes, C Houk, S F Ahmed, P A Lee, LWPES1/ESPE2 Consensus Group
PMCID: PMC2082839  PMID: 16624884

Short abstract

Management of intersex disorders

Keywords: sex development disorders, gonads, intersex


The birth of an intersex child prompts a long term management strategy that involves a myriad of professionals working with the family. It is estimated that genital anomalies occur in 1 in 4500 births. There has been progress in diagnosis, surgical techniques, understanding psychosocial issues, and recognising and accepting the place of patient advocacy. The Lawson Wilkins Pediatric Endocrine Society (LWPES) and the European Society for Paediatric Endocrinology (ESPE) considered it timely to review the management of intersex disorders from a broad perspective, to review data on longer term outcome, and to formulate proposals for future studies. The methodology comprised establishing several working groups whose membership was drawn from 50 international experts in the field. The groups prepared prior written responses to a defined set of questions resulting from an evidence based review of published reports. At a subsequent gathering of participants, a framework for a consensus document was agreed. This paper constitutes its final form.

Advances in identification of molecular genetic causes of abnormal sex with heightened awareness of ethical issues and patient advocacy concerns necessitate a re‐examination of nomenclature.1 Terms such as intersex, pseudohermaphroditism, hermaphroditism, sex reversal, and gender based diagnostic labels are particularly controversial. These terms are perceived as potentially pejorative by patients,2 and can be confusing to practitioners and parents alike. The term “disorders of sex development” (DSD) is proposed, as defined by congenital conditions in which development of chromosomal, gonadal, or anatomical sex is atypical.

The proposed changes in terminology are summarised in table 1. A modern lexicon is needed to integrate progress in molecular genetic aspects of sex development. As outcome data in individuals with DSD are limited, it is essential to employ precision when applying definitions and diagnostic labels.3,4 It is also appropriate to use terminology that is sensitive to the concerns of patients. The ideal nomenclature should be sufficiently flexible to incorporate new information yet robust enough to maintain a consistent framework. Terms should be descriptive and reflect genetic aetiology when available, and accommodate the spectrum of phenotypical variation. Clinicians and scientists must value its use and it must be understandable to patients and their families. An example of how the proposed nomenclature could be applied in a classification of DSD is shown in table 2.

Table 1 Proposed revised nomenclature.

Previous Proposed
Intersex Disorders of sex development (DSD)
Male pseudohermaphrodite
Undervirilisation of an XY male 46,XY DSD
Undermasculinisation of an XY male
Female pseudohermaphrodite
Overvirilisation of an XX female 46,XX DSD
Masculinisation of an XX female
True hermaphrodite Ovotesticular DSD
XX male or XX sex reversal 46,XX testicular DSD
XY sex reversal 46,XY complete gonadal dysgenesis

Table 2 An example of a DSD classification.

Sex chromosome DSD 46,XY DSD 46,XX DSD
(A) 45,X (Turner syndrome and variants) (A) Disorders of gonadal (testicular) development (A) Disorders of gonadal (ovarian) development
1. Complete gonadal dysgenesis (Swyer syndrome) 1. Ovotesticular DSD
2. Testicular DSD (eg, SRY+, dup SOX9)
(B)47,XXY (Klinefelter syndrome and variants) 2. Partial gonadal dysgenesis 3. Gonadal dysgenesis
3. Gonadal regression
4. Ovotesticular DSD
(C)45,X/46,XY (mixed gonadal dysgenesis, ovotesticular DSD) (B)Disorders in androgen synthesis or action (B) Androgen excess
1. Androgen biosynthesis defect (eg, 17‐hydroxysteroid dehydrogenase deficiency, 5α reductase deficiency, StAR mutations 1. Fetal (eg, 21‐hydroxylase deficiency, 11‐hydroxylase deficiency)
2. Fetoplacental (aromatase deficiency, POR)
(D)46,XX/46,XY (chimeric, ovotesticular DSD) 2. Defect in androgen action (eg, CAIS, PAIS) 3. Maternal (luteoma, exogenous, etc)
 
3. LH receptor defects (eg, Leydig cell          
hypoplasia, aplasia)        
4. Disorders of AMH and AMH receptor (persistent mullerian duct syndrome)        
(C) Other        
(C)Other (eg, cloacal extrophy, vaginal atresia, MURCS, other        
(eg, severe hypospadias, cloacal extrophy) syndromes)        

While consideration of karyotype is useful for classification, unnecessary reference to karyotype should be avoided; ideally, a system based on descriptive terms (for example, androgen insensitivity syndrome) should be used wherever possible.

AMH, anti‐mullerian hormone; CAIS, complete androgen insensitivity syndrome; DSD, disorders of sex development; MURCS, mullerian, renal, cervicothoracic somite abnormalities; PAIS, partial androgen insensitivity syndrome; POR, cytochrome P450 oxidoreductase.

Psychosexual development is traditionally conceptualised as three components. Gender identity refers to a person's self representation as male or female (with the caveat that some individuals may not identify exclusively with either). Gender role (sex‐typical behaviours) describes the psychological characteristics that are sexually dimorphic within the general population, such as toy preferences and physical aggression. Sexual orientation refers to the direction(s) of erotic interest (heterosexual, bisexual, homosexual) and includes behaviour, fantasies, and attractions. Psychosexual development is influenced by multiple factors such as exposure to androgens, sex chromosome genes, and brain structure, as well as social circumstance and family dynamics.

Gender dissatisfaction denotes unhappiness with assigned sex. Causes of gender dissatisfaction are poorly understood, even among individuals without DSD. Gender dissatisfaction occurs more frequently in individuals with DSD than in the general population, but is difficult to predict from karyotype, prenatal androgen exposure, degree of genital virilisation, or assigned gender.5,6,7 Prenatal androgen exposure is clearly associated with other aspects of psychosexual development.8,9 There are dose related effects on childhood play behaviour in girls with congenital adrenal hyperplasia (CAH), whereby those with the more severe mutations and marked genital virilisation play more with boys' toys.10 Prenatal androgen exposure is also associated with other psychological characteristics such as maternal interest and sexual orientation. It is important to emphasise the separability of sex‐typical behaviour, sexual orientation, and gender identity. Thus homosexual orientation (relative to sex of rearing) or strong cross‐sex interest in an individual with DSD is not an indication of incorrect gender assignment. Understanding variations in psychosexual development in individuals with DSD requires reference to studies in non‐human species that show marked but complex effects of androgens on sex differentiation of the brain and on behaviour. Outcomes can be influenced by timing, dose, and type of androgen exposure, receptor availability, and modification by the social environment.11,12,13,14

Data from rodent studies suggest that sex chromosome genes may also influence brain structure and behaviour directly.15,16 However, studies in individuals with complete androgen insensitivity syndrome (CAIS) do not indicate a behavioural role for Y chromosome genes, although data are limited.17 Sex differences in brain structures have been identified across species, some of which coincide with pubertal onset, perhaps suggesting hormonal responsivity.18,19,20 The limbic system and hypothalamus, both playing a role in reproduction, show sex differences in specific nuclei but it is not clear when these differences emerge. Interpretation of sex differences is complicated by the effect of cell death and synaptic pruning on normal maturation and by effects of experience on the brain. Structure of the brain is not currently useful for gender assignment.

Investigation and management of DSD

General concepts of care

Optimal clinical management of individuals with DSD21 should comprise the following:

  • gender assignment must be avoided before expert evaluation in newborns;

  • evaluation and long term management must be carried out at a centre with an experienced multidisciplinary team;

  • all individuals should receive a gender assignment;

  • open communication with patients and families is essential and participation in decision making is encouraged;

  • patient and family concerns should be respected and addressed in strict confidence.

The initial contact with the parents of a child with a DSD is important, as first impressions from these encounters often persist. A key point to emphasise is that the DSD child has the potential to become a well adjusted, functional member of society. While privacy needs to be respected, DSD is not shameful. It should be explained to the parents that the best course of action may not initially be clear, but the health care team will work with the family to reach the best possible set of decisions in the circumstances. The health care team should discuss with the parents what information to share in the early stages with family members and friends. Parents need to be informed about sexual development, and web based information may be helpful, provided the content and focus of the information is balanced and sound (http://www.sickkids.ca/childphysiology/cpwp/genital/genitalintro.html).

Ample time and opportunity should be made for continued discussion with review of information previously provided.1

The multidisciplinary team

Optimal care for children with DSD requires an experienced multidisciplinary team which is generally found in tertiary care centres. Ideally, the team includes paediatric subspecialists in endocrinology, surgery or urology or both, psychology/psychiatry, gynaecology, genetics, neonatology, and, if available, social work, nursing, and medical ethics.22 Core composition will vary according to DSD type, local resources, developmental context, and location. Ongoing communication with the family primary care physician is essential.23

The team has a responsibility to educate other health care staff in the appropriate initial management of affected newborn infants and their families. For new DSD patients, the team should develop a plan for clinical management with respect to diagnosis, gender assignment, and treatment options before making any recommendations. Ideally, discussions with the family are conducted by one professional with appropriate communication skills.24 Transitional care should be organised with the multidisciplinary team operating in an environment comprising specialists with experience in both paediatric and adult practice. Support groups have an important role in the delivery of care to DSD patients and their families25 (see appendix 1).

Clinical evaluation

A family and prenatal history, a general physical examination with attention to any associated dysmorphic features, and an assessment of the genital anatomy in comparison with published norms needs to be recorded (table 3). Criteria that suggest DSD include:

Table 3 Anthropometric measurements of the external genitalia.

Sex Population Age Stretched penile length Penile width Mean testicular Reference
(PL) (cm) (cm) volume (ml)
M USA 30 wks GA 2.5 26
M USA Full term 3.5 (0.4) 1.1 (0.1) 0.52 (median) 26, 27
M Japan Term to 14 years 2.9 (0.4) to 8.3 (0.8) 28
M Australia 24–36 weeks GA PL  = 2.27 + (0.16 GA) 29
M China Term 3.1 (0.3) 1.07 (0.09) 30
M India Term 3.6 (0.4) 1.14 (0.07) 30
M N America Term 3.4 (0.3) 1.13 (0.08) 30
M Europe 10 years 6.4 (0.4) 0.95 to 1.20 27, 31
M Europe Adult 13.3 (1.6) 16.5 to 18.2 27, 31
Sex Population Age Clitoral length (mm) Clitoral width (mm) Perineal length* (mm) Reference
F USA Full term 4.0 (1.24) 3.32 (0.78) 32
F USA Adult nulliparous 15.4 (4.3) 33
F UK Adult 19.1 (8.7) 5.5 (1.7) 31.3 (8.5) 34

Values are mean (SD) unless specified.

*Distance from posterior fourchette to anterior anal margin.

GA, gestational age; PL, penile length.

  • overt genital ambiguity (for example, cloacal exstrophy);

  • apparent female genitalia with an enlarged clitoris, posterior labial fusion, or an inguinal/labial mass;

  • apparent male genitalia with bilateral undescended testes, micropenis, isolated perineal hypospadias, or mild hypospadias with undescended testis;

  • a family history of DSD such as CAIS;

  • a discordance between genital appearance and a prenatal karyotype.

Most causes of DSD are recognised in the neonatal period; later presentations in older children and young adults include: previously unrecognised genital ambiguity; inguinal hernia in a girl; delayed or incomplete puberty; virilisation in a girl; primary amenorrhoea; breast development in a boy; and gross and occasionally cyclical haematuria in a boy.

Diagnostic evaluation

Considerable progress has been made over understanding the genetic basis of human sexual development,35 yet a specific molecular diagnosis is identified in only about 20% of cases of DSD. The majority of virilised 46,XX infants will have CAH. In contrast, only 50% of 46,XY children with DSD will receive a definitive diagnosis.36,37 Diagnostic algorithms do exist, but with the spectrum of findings and diagnoses, no single evaluation protocol can be recommended in all circumstances. Some tests, such as imaging by ultrasound, are operator dependent. Hormone measurements need to be interpreted in relation to the specific assay characteristics and to normal values for gestational and chronological age. In some cases serial measurements may be needed.

First line testing in newborns includes: karyotyping with X and Y specific probe detection (even when prenatal karyotype is available), imaging (abdomino‐pelvic ultrasound), measurement of 17‐hydroxyprogesterone, testosterone, gonadotropins, anti‐mullerian hormone, serum electrolytes, and urinanalysis. The results of these investigations are generally available within 48 hours and will be sufficient for making a working diagnosis. Decision making algorithms are available to guide further investigation.38 These include hCG and ACTH stimulation tests to assess testicular and adrenal steroid biosynthesis, urinary steroid analysis by GC mass spectroscopy, imaging studies, and biopsies of gonadal material. Some gene analyses are carried out in clinical service laboratories. However, current molecular diagnosis is limited by cost, accessibility, and quality control.39 Research laboratories provide genetic testing, including functional analysis, but may face restrictions on communicating results.40

Gender assignment in newborn infants

Initial gender uncertainty is unsettling and stressful for families. Expediting a thorough assessment and decision is required. Factors that influence gender assignment include the diagnosis, genital appearance, surgical options, need for life long replacement therapy, the potential for fertility, views of the family, and sometimes the circumstances relating to cultural practices. More than 90% of 46,XX CAH patients41 and all 46,XY CAIS assigned females in infancy42 identify as females. Evidence supports the current recommendation to raise markedly virilised 46,XX infants with CAH as female.43 Approximately 60% of 5α‐reductase (5αRD2) deficient patients assigned female in infancy and virilising at puberty (and all assigned male) live as males.5 In 5αRD2 and possibly 17β‐hydroxysteroid dehydrogenase (17βHSD3) deficiencies, where the diagnosis is made in infancy, the combination of a male gender identity in the majority and the potential for fertility (documented in 5αRD2, but unknown in 17βHSD3) should be discussed when providing evidence for gender assignment.5,44,45 Among patients with PAIS, androgen biosynthetic defects, and incomplete gonadal dysgenesis, there is dissatisfaction with the sex of rearing in about 25% of individuals, whether raised male or female.46 Available data support male rearing in all patients with micropenis, taking into account equal satisfaction with assigned gender in those raised male or female, but no need for surgery, and the potential for fertility in patients reared male.42 The decision on sex of rearing in ovotesticular DSD should consider the potential for fertility based on gonadal differentiation and genital development, and assuming the genitalia are, or can be made, consistent with the chosen sex. In the case of mixed gonadal dysgenesis (MGD), factors to consider include prenatal androgen exposure, testicular function at and after puberty, phallic development, and gonadal location. Individuals with cloacal exstrophy reared female show variability in gender identity outcome, but more than 65% appear to live as female.6

Surgical management

The surgeon has a responsibility to outline the surgical sequence and subsequent consequences from infancy to adulthood. Only surgeons with expertise in the care of children and specific training in the surgery of DSD should undertake these procedures. Parents now appear to be less inclined to choose surgery for less severe clitoromegaly.47 Surgery should only be considered in cases of severe virilisation (Prader III, IV, and V) and should be carried out in conjunction, when appropriate, with repair of the common urogenital sinus. As orgasmic function and erectile sensation may be disturbed by clitoral surgery, the surgical procedure should be anatomically based to preserve erectile function and the innervation of the clitoris. Emphasis is on functional outcome rather than a strictly cosmetic appearance. It is generally felt that surgery that is carried out for cosmetic reasons in the first year of life relieves parental distress and improves attachment between the child and the parents.48,49,50,51 The systematic evidence for this belief is lacking.

There is inadequate evidence currently in relation to establishment of functional anatomy, to abandon the practice of early separation of the vagina and urethra.52 The rationale for early reconstruction is based on guidelines on the timing of genital surgery from the American Academy of Pediatrics (AAP),53 the beneficial effects of oestrogen on tissue in early infancy, and the avoidance of potential complications from the connection between the urinary tract and peritoneum through the Fallopian tubes. It is anticipated that surgical reconstruction in infancy will need to be refined at the time of puberty.54,55,56 Vaginal dilatation should not be undertaken before puberty. The surgeon must be familiar with several operative techniques in order to reconstruct the spectrum of urogenital sinus disorders. An absent or inadequate vagina (with rare exceptions) requires a vaginoplasty performed in adolescence when the patient is psychologically motivated and a full partner in the procedure. No one technique has been universally successful; self dilatation, skin substitution, and bowel vaginoplasty each have specific advantages and disadvantages.

In the case of a DSD associated with hypospadias,57 standard techniques for surgical repair such as chordee correction, urethral reconstruction, and the judicious use of testosterone supplementation apply. The magnitude and complexity of phalloplasty in adulthood should be taken into account during the initial counselling period if successful gender assignment is dependent on this procedure.58 At times this may affect the balance of gender assignment. Patients must not be given unrealistic expectations about penile reconstruction, including the use of tissue engineering. There is no evidence that prophylactic removal of asymptomatic discordant structures, such as a utriculus or mullerian remnants, is required although symptoms in future may indicate surgical removal. For the male who has a successful neophalloplasty in adulthood, an erectile prosthesis may be inserted but has a high morbidity.

The testes in patients with CAIS35 and those with PAIS, raised female, should be removed to prevent malignancy in adulthood. The availability of oestrogen replacement therapy allows for the option of early removal at the time of diagnosis which also takes care of the associated hernia, psychological problems with the presence of testes, and the malignancy risk. Parental choice allows deferment until adolescence, recognising that the earliest reported malignancy in CAIS is at 14 years of age.59 The streak gonad in a patient with MGD raised male should be removed laparoscopically (or by laparotomy) in early childhood.35 Bilateral gonadectomy is performed in early childhood in females (bilateral streak gonads) with gonadal dysgenesis and Y chromosome material. In patients with androgen biosynthetic defects raised female, gonadectomy should be undertaken before puberty. A scrotal testis in patients with gonadal dysgenesis is at risk for malignancy. Current recommendations are testicular biopsy at puberty seeking signs of the premalignant lesion termed carcinoma in situ or undifferentiated intratubular germ cell neoplasia. If positive, the option is sperm banking before treatment with local low dose radiotherapy which is curative.60

Surgical management in DSD should also consider options that will facilitate the chances of fertility. In patients with a symptomatic utriculus, removal is best undertaken laparoscopically to increase the chance of preserving continuity of the vasa deferentia. Patients with bilateral ovotestes are potentially fertile from functional ovarian tissue.35,61 Separation of ovarian and testicular tissue can be technically difficult and should be undertaken, if possible, in early life.

Sex steroid replacement

Hypogonadism is common in patients with dysgenetic gonads, defects in sex steroid biosynthesis, and resistance to androgens. The timing of initiation of puberty may vary but this is an occasion that provides an opportunity to discuss the condition and set a foundation for long term adherence to treatment. Hormonal induction of puberty should attempt to replicate normal pubertal maturation to induce secondary sexual characteristics, a pubertal growth spurt, and optimal bone mineral accumulation, together with psychosocial support for psychosexual maturation.62 Intramuscular depot injections of testosterone esters are commonly used in males; other options include oral testosterone undecanoate, and transdermal preparations are also available.63,64,65 Patients with PAIS may require supraphysiological doses of testosterone for optimal effect.66 Females with hypogonadism require oestrogen supplementation to induce pubertal changes and menses. A progestin is usually added after breakthrough bleeding develops or within one to two years of continuous oestrogen. There is no evidence that the addition of cyclic progesterone is beneficial in women without a uterus.

Psychosocial management

Psychosocial care provided by mental health staff with expertise in DSD should be an integral part of management in order to promote positive adaptation. This expertise can facilitate team decisions about gender assignment/reassignment, timing of surgery, and sex hormone replacement. Psychosocial screening tools that identify families at risk for maladaptive coping with a child's medical condition are available.67 Once the child is sufficiently developed for a psychological assessment of gender identity, such an evaluation must be included in discussions about gender reassignment. Gender identity development begins before the age of 3 years,68 but the earliest age at which it can be reliably assessed remains unclear. The generalisation that the age of 18 months is the upper limit of imposed gender reassignment should be treated with caution and viewed conservatively. Atypical gender role behaviour is more common in children with DSD than in the general population but should not be taken as an indicator for gender reassignment. In affected children and adolescents who report significant gender dysphoria, a comprehensive psychological evaluation69 and an opportunity to explore feelings about gender with a qualified clinician is required over a period of time. If the desire to change gender persists, the patient's wish should be supported and may require the input of a specialist skilled in the management of gender change.

The process of disclosure concerning facts about karyotype, gonadal status, and prospects for future fertility is a collaborative ongoing action which requires a flexible individual based approach. It should be planned with the parents from the time of diagnosis.70 Studies in other chronic medical disorders and of adoptees indicate that disclosure is associated with enhanced psychosocial adaptation.71 Medical education and counselling for children is a recurrent gradual process of increasing sophistication which is commensurate with changing cognitive and psychological development.72

Quality of life encompasses falling in love, dating, attraction, ability to develop intimate relationships, sexual functioning, and the opportunity to marry and to raise children, regardless of biological indicators of sex. The most frequent problems encountered in DSD patients are sexual aversion and lack of arousability, which are often misinterpreted as low libido.73 Health care staff should offer adolescent patients opportunities to talk confidentially without their parents and encourage the participation in condition specific support groups which enhance the ability of the patient to discuss their concerns comfortably. Some patients avoid intimate relationships and it is important to address fears of rejection and advise on the process of building a relationship with a partner. The focus should be on interpersonal relationships and not solely on sexual function and activity. Referral for sex therapy may be needed. Repeated examination of the genitalia, including medical photography, may be experienced as deeply shaming.74 Medical photography has its place for record keeping and education, but should be undertaken whenever possible if the patient is under anaesthesia for a procedure and with appropriate consent. Medical interventions and negative sexual experiences may have fostered symptoms of post‐traumatic stress disorder and referral to a qualified mental health professional may be indicated.75

Outcome in DSD

As a general statement, information across a range of assessments is insufficient in DSD. The following is based on those disorders where some evidence base is available. They include CAH, CAIS and PAIS, disorders of androgen biosynthesis, gonadal dysgenesis syndromes (complete and partial), and micropenis. Long term outcome in DSD should include the following: external and internal genital phenotype, physical health including fertility, sexual function, social and psychosexual adjustment, mental health, quality of life, and social participation. There are additional health problems in individuals with DSD. These include the consequences of associated problems such as other malformations, developmental delay and intellectual impairment, delayed growth and development, and unwanted effects of hormones on libido and body image.76

Surgical outcome

Some studies suggest satisfactory outcomes from early surgery.43,46,47,77 Nevertheless, outcomes from clitoroplasty identify problems related to decreased sexual sensitivity, loss of clitoral tissue, and cosmetic issues.78 Techniques for vaginoplasty carry the potential for scarring at the introitus, necessitating repeated modification before sexual function can be reliable. Surgery to construct a neo‐vagina carries a risk of neoplasia.79 The risks from vaginoplasty are different for high and low confluence of the urethra and vagina. Analysis of long term outcomes is complicated by a mixture of surgical techniques and diagnostic categories.80 Few women with CAIS need surgery to lengthen the vagina.81

The outcome in undermasculinised males with a phallus is dependent on the degree of hypospadias and the amount of erectile tissue. Feminising as opposed to masculinising genitoplasty requires less surgery to achieve an acceptable outcome and results in fewer urological difficulties.46 Long term data on sexual function and quality of life among those assigned female as well as male show great variability. There are no controlled clinical trials of the efficacy of early (less than 12 months of age) versus late surgery (in adolescence and adulthood), or of the efficacy of different techniques.

Risk of gonadal tumours

Interpretation of published reports is hampered by unclear terminology and by effects of normal cell maturation delay.82,83,84 The highest tumour risk is found in TSPY (testis‐specific protein Y encoded) positive gonadal dysgenesis and PAIS with intra‐abdominal gonads, while the lowest risk (<5%) is found in ovotestis85 and CAIS.83,86 Table 4 provides a summary of the risk of tumour development according to diagnosis and recommendations for management.

Table 4 Risk of germ cell malignancy according to diagnosis.

Risk group Disorder Malignancy risk Recommended action Studies (n) Patients (n)
 
High GD* (+Y)† intra‐abdominal 15–35 Gonadectomy‡ 12 >350
PAIS non‐scrotal 50 Gonadectomy‡ 2 24
Frasier 60 Gonadectomy‡ 1 15
Denys‐Drash (+Y) 40 Gonadectomy‡ 1 5
Intermediate Turner (+Y) 12 Gonadectomy‡ 11 43
17β‐HSD 28 Monitor 2 7
GD (+Y)‡ scrotal Unknown Biopsy§ and irradiation? 0 0
PAIS scrotal gonad Unknown Biopsy§ and irradiation? 0 0
Low CAIS 2 Biopsy§ and ??? 2 55
Ovotestis DSD 3 Testis tissue removal? 3 426
Turner (–Y) 1 None 11 557
No (?) 5α‐reductase 0 Unresolved 1 3
Leydig cell hypoplasia 0 Unresolved 2

*Gonadal dysgenesis (including not further specified, 46XY, 46X/46XY, mixed, partial, complete).

†GBY region positive, including the TSPY gene.

‡At time of diagnosis.

§At puberty, allowing investigation of at least 30 seminiferous tubules, with diagnosis preferably based on OCT3/4 immunohistochemistry.

CAIS, complete androgen insensitivity syndrome; DSD, disorders of sex development; HSD, hydroxysteroid dehydrogenase deficiency; PAIS, partial androgen insensitivity syndrome.

Cultural and social factors

DSD may carry a stigma. Social and cultural factors, as well as hormonal effects, appear to influence gender role in 5α‐reductase deficiency. Gender role change occurs at different rates in different societies, suggesting that social factors may also be important modifiers of gender role change.

In some societies, female infertility precludes marriage, which also affects employment prospects and creates economic dependence. Religious and philosophical views may influence how parents respond to the birth of an infant with a medical condition. Fatalism and guilt feelings in relation to congenital malformations or genetic conditions have an influence, while poverty and illiteracy negatively affect access to health care.87

Future studies

Establishing a precise diagnosis in DSD is just as important as in other chronic medical conditions with lifelong consequences. Considerable progress has been achieved with molecular studies, as illustrated in tables 5 and 6, which summarise the genes known to be involved in DSD. Use of tissue specific animal knock out models, comparative genomic hybridisation, and microarray screens of the mouse urogenital ridge will provide benefits in identifying new genes causing DSD.89 It is essential that the momentum for an international collaborative approach to this task is maintained.

Table 5 Genes known to be involved in disorders of sex development: 46,XY.

Gene Protein OMIM Locus Inheritance Gonad Mullerian structures External genitalia Associated features/variant phenotypes
46,XY DSD
Disorders of gonadal (testicular) development: single gene disorders
WT1 TF 607102 11p13 AD Dysgenetic testis +/− Female or ambiguous Wilms tumour, renal abnormalities, gonadal tumours (WAGR, Denys‐Drash, and Frasier syndromes)
SF1 (NR5A1) Nuclear receptor TF 184757 9q33 AD/AR Dysgenetic testis +/− Female or ambiguous More severe phenotypes include primary adrenal failure; milder phenotypes have isolated partial gonadal dysgenesis
SRY TF 480000 Yp11.3 Y Dysgenetic testis or ovotestis +/− Female or ambiguous
SOX9 TF 608160 17q24‐25 AD Dysgenetic testis or ovotestis +/− Female or ambiguous Campomelic dysplasia (17q24 rearrangements milder phenotype than point mutations)
DHH Signalling molecule 605423 12q13.1 AR Dysgenetic testis + Female The severe phenotype of one patient included minifascicular neuropathy, other patients have isolated gonadal dysgenesis
ATRX Helicase (?chromatin remodelling) 300032 Xq13.3 X Dysgenetic testis Female, ambiguous or male α‐Thalassaemia, mental retardation
ARX TF 300382 Xp22.13 X Dysgenetic testis Ambiguous X‐linked lissencephaly, epilepsy, temperature instability
Disorders of gonadal (testicular) development: chromosomal changes involving key candidate genes
DMRT1 TF 602424 9p24.3 Monosomic deletion Dysgenetic testis +/− Female or ambiguous Mental retardation
DAX1 (NR0B1) Nuclear receptor TF 300018 Xp21.3 dupXp21 Dysgenetic testis or ovary +/− Female or ambiguous
WNT4 Signalling molecule 603490 1p35 dup1p35 Dysgenetic testis + Ambiguous Mental retardation
Disorders of hormone synthesis or action
LHGCR G protein receptor 152790 2p21 AR Testis Female, ambiguous or micropenis Leydig cell hypoplasia
DHCR7 Enzyme 602858 11q12‐13 AR Testis Variable Smith‐Lemli‐Opitz syndrome: coarse facies, second‐third toe syndactyly, failure to thrive, developmental delay, cardiac and visceral abnormalities
STAR Mitochondrial membrane protein 600617 8p11.2 AR Testis Female Congenital lipoid adrenal hyperplasia (primary adrenal failure), pubertal failure
CYP11A1 Enzyme 118485 15q23‐24 AR Testis Female or Ambiguous Congenital adrenal hyperplasia (primary adrenal failure), pubertal failure
HSD3B2 Enzyme 201810 1p13.1 AR Testis Ambiguous CAH, primary adrenal failure, partial androgenisation due to ↑DHEA
CYP17 Enzyme 202110 10q24.3 AR Testis Female, ambiguous or micropenis CAH, hypertension due to ↑corticosterone & 11‐deoxycorticosterone (except in isolated 17,20‐lyase deficiency)
POR (P450 oxido‐reductase) CYP enzyme electron donor 124015 7q11.2 AR Testis Male or ambiguous Mixed features of 21‐hydroxylase deficiency, 17α‐hydroxylase/17,20‐lyase deficiency, and aromatase deficiency; sometimes associated with Antley Bixler craniosynostosis
HSD17B3 Enzyme 605573 9q22 AR Testis Female or ambiguous Partial androgenisation at puberty, ↑androstenedione:testosterone ratio
SRD5A2 Enzyme 607306 2p23 AR Testis Ambiguous or micropenis Partial androgenisation at puberty, ↑testosterone:DHT ratio,
AMH Signalling molecule 600957 19p13.3‐13.2 AR Testis + Normal male Persistent mullerian duct syndrome (PMDS). Male external genitalia, bilateral cryptorchidism
AMH‐Receptor Serine‐threonine kinase transmembrane receptor 600956 12q13 AR Testis + Normal male
Androgen receptor Nuclear receptor TF 313700 Xq11‐12 X Testis Female, ambiguous, micropenis or normal male Phenotypic spectrum from complete androgen insensitivity syndrome (female external genitalia) and partial androgen insensitivity (ambiguous) to normal male genitalia/infertility

AD, autosomal dominant; AR, autosomal recessive; CAH, congenital adrenal hyperplasia; TF, transcription factor; X, X‐linked recessive; Y, Y‐linked recessive.

Copyright 2002, The Endocrine Society.

Table 6 Genes known to be involved in disorders of sex development: 46,XX.

Gene Protein OMIM Locus Inheritance Gonad Mullerian structures External genitalia Associated features/variant phenotypes
Disorders of gonadal (ovarian) development
SRY TF 480000 Yp11.3 Translocation Testis or ovotestis Male or ambiguous
SOX9 TF 608160 17q24 dup17q24 ND Male or ambiguous
Androgen excess
HSD3B2 Enzyme 201810 1p13 AR Ovary + Clitoromegaly CAH, primary adrenal failure, partial androgenisation due to ↑DHEA
CYP21A2 Enzyme 201910 6p21‐23 AR Ovary + Ambiguous CAH, phenotypic spectrum from severe salt losing forms associated with adrenal failure to simple virilising forms with compensated adrenal function, ↑17‐hydroxyprogesterone
CYP11B1 Enzyme 202010 8q21‐22 AR Ovary + Ambiguous CAH, hypertension due to ↑11‐deoxycortisol and 11‐deoxycorticosterone
POR (P450 oxido‐reductase) CYP enzyme electron donor 124015 7q11.2 AR Ovary + Ambiguous Mixed features of 21‐hydroxylase deficiency, 17α‐hydroxylase/17,20‐lyase deficiency and aromatase deficiency; associated with Antley Bixler craniosynostosis
CYP19 Enzyme 107910 15q21 AR Ovary + Ambiguous Maternal androgenisation during pregnancy, absent breast development at puberty, except in partial cases
Glucocorticoid receptor Nuclear receptor TF 138040 5q31 AR Ovary + Ambiguous ↑ACTH, 17‐hydroxyprogesterone and cortisol; failure of dexamethasone suppression(NB patient heterozygous for a mutation in CYP21)

ACTH, adrenocorticotropin; AD, autosomal dominant; AR, autosomal recessive; CAH, congenital adrenal hyperplasia; ND, not determined; TF, transcription factor; X, X chrosomal; Y, Y chromosomal.

Chromosomal rearrangements likely to include key genes are included. Modified from Achermann et al88, with permission from the Endocrine Society.

Much remains to be clarified about the determinants of gender identity in DSD. Future studies require representative sampling to carefully conceptualise and measure gender identity, recognising that there are multiple determinants to consider and gender identity may change into adulthood. In terms of psychological management, studies are needed to evaluate the effectiveness of information management with regard to timing and content. The pattern of surgical practice in DSD is changing with respect to the timing of surgery and the techniques employed. It is essential to evaluate the effects of early versus later surgery in an holistic manner, recognising the difficulties posed by an ever evolving clinical practice.

The consensus has clearly identified a major shortfall in information about long term outcome. Future studies should use appropriate instruments that assess outcomes in a standard manner68,69 and take cognisance of guidelines relevant to all chronic conditions (http://www.who.int/classifications/icf/en/). These should preferably be prospective in nature and designed to avoid selection bias. Several countries already have registers of DSD cases but there could be added benefit from pooling such resources to enable prospective multicentre studies to be undertaken on a larger number of cases that are clearly defined. Allied to this should be an educational programme to ensure that professionals tasked with providing care for DSD families are suitably trained to discharge their responsibilities.

Acknowledgements

The LWPES and ESPE gratefully acknowledge unrestricted educational grant support for the consensus meeting from Pfizer Endocrine Care, Novo Nordisk, Ferring, and Organon. The work of Alan Rogol, Joanne Rogol, Pauline Bertrand, and Pam Stockham in organising the meeting is gratefully appreciated.

Consensus Group

The following participants contributed to the production of the Consensus document: John Achermann (London, UK), Faisal Ahmed (Glasgow, UK), Laurence Baskin (San Francisco, USA), Sheri Berenbaum (University Park, USA), Sylvano Bertelloni (Pisa, Italy), John Brock (Nashville, USA), Polly Carmichael (London, UK), Cheryl Chase (Rohnert Park, USA), Peggy Cohen‐Kettenis (Amsterdam, Netherlands), Felix Conte (San Francisco, USA), Patricia Donohoue (Iowa City, USA), Chris Driver (Aberdeen, UK), Stenvert Drop (Rotterdam, Netherlands), Erica Eugster (Indianapolis, USA), Kenji Fujieda (Asahikawa, Japan), Jay Giedd (Bethesda, USA), Richard Green (London, UK), Melvin Grumbach (San Francisco, USA), Vincent Harley (Victoria, Australia), Melissa Hines (London, UK), Olaf Hiort (Lübeck, Germany), Ieuan Hughes (Cambridge, UK), Peter Lee (Hershey, USA), Leendert Looijenga (Rotterdam, Netherlands), Berenice Mendonça (Sao Paulo, Brazil), Heino Meyer‐Bahlburg (New York, USA), Claude Migeon (Baltimore, USA), Yves Morel (Lyon, France), Pierre Mouriquand (Lyon, France), Anna Nordenström (Stockholm, Sweden), Phillip Ransley (London, UK), Robert Rapaport (New York, USA), William Reiner (Oklahoma City, USA), Hertha Richter‐Appelt (Hamburg, Germany), Richard Rink (Indianapolis, USA), Emilie Rissman (Charlottesville, USA), Paul Saenger (New York, USA), David Sandberg (Buffalo, USA), Justine Schober (Erie, USA), Norman Spack (Boston, USA), Barbara Thomas (Rottenburg am Neckar, Germany), Ute Thyen (Lübeck, Germany), Eric Vilain (Los Angeles, USA), Garry Warne (Melbourne, Australia), Amy Wisniewski (Des Moines, USA), Jean Wilson (Dallas, USA), Christopher Woodhouse (London, UK), Kenneth Zucker (Toronto, Canada).

Abbreviations

CAH - congenital adrenal hyperplasia

CAIS - complete androgen insensitivity syndrome

DSD - disorders of sex development

ESPE - European Society for Paediatric Endocrinology

LWPES - Lawson Wilkins Pediatric Endocrine Society

MGD - mixed gonadal dysgenesis

PAIS - partial androgen insensitivity syndrome

Appendix 1

ROLE OF SUPPORT GROUPS

The value of peer and parent support for many chronic medical conditions is widely accepted, and DSDs, being lifelong conditions which affect developmental tasks at many stages of life, are no exception.

Those affected by DSDs and parent members value the following:

  • Peer support ends isolation and stigma, providing a context in which conditions are put into perspective, and where intimate issues of concern can be discussed safely with someone who has “been there.”

  • Children who form relationships with peers and affected adults early in their lives benefit from a feeling of normalcy early on, with support in place well before adolescence. Adolescents often resist attempts to introduce them to peer support.

  • Support groups can help families and consumers find the best quality care.

While clinical practice may focus on gender and genital appearance as key outcomes, stigma and experiences associated with having a DSD (both within and outside the medical environment) are more salient issues for many affected people.

Support groups complement the work of the health care team and, together, can help improve services. Initiatives by support groups have led to improvements in management of DSD and research directed towards clinically relevant issues. Dialogue between health care professionals and support groups, and collaboration as partners is to be encouraged.

APPENDIX 2

LEGAL ISSUES

Basic principles of medical law will remain, even as research and clinical experience evolve in aetiology, diagnosis, and treatment. This appendix draws on practice in three countries on standards of medical negligence and patient informed consent. In the USA, the medical profession sets standards of care based on prevailing medical custom.90 However, a treatment may also be that used by a respected minority of practitioners.

USA

Informed consent in the USA was founded on the principle of battery, whereby it is an offence to violate another person's bodily integrity without consent. Nowadays, most states are concerned with negligent non‐disclosure to the patient. The standard of adequate disclosure may be physician based, requiring conduct of a reasonable practitioner, or it may be patient based, asking what a reasonable patient would find material. Physician based disclosure must include information about risks, alternatives, outcomes, and prognosis, with or without treatment.

US courts assume that parents know what is best for their child when parental authority applies to consent for the child (substituted judgement). Parental decisions are deferred to except in situations where potentially life saving treatment is withheld. Consent to treatment by a child is dependent on an understanding of its nature and consequences.

United Kingdom

Medical negligence in the United Kingdom defines treatment that falls below the standard expected of a reasonably competent practitioner. The standard of proof in court is whether negligence is demonstrated on the balance of probabilities. It is incumbent on the practitioner to demonstrate that treatment was consistent with a rationally defensible body of medical opinion. A shift in parental prerogative to consent to treatment was reflected in the Children Act 1989, in which parental rights were replaced by parental responsibilities.91 UK courts can intervene with orders made requiring or preventing a specific action related to the child. Age is not a barrier to informed consent, providing that a minor demonstrates an understanding of the issues sufficient to have the capacity to consent.

Colombia

Colombian law is noted for a reasoned set of guidelines advanced by the highest court in cases of DSD.92 A protocol was formulated for parental and physician intervention. The process of consent requires “qualified and persistent informed consent” over an extended period of time. Authorisation is given in stages to allow time for the parents to come to terms with their child's condition. The court aimed to strike a balance between parental autonomy for those who did and those who did not want early surgery for their child, until there was clear evidence of harm in deferring surgery until the child was competent to decide. Parents cannot consent for children over 5 years of age, as by then children are deemed to have identified with a gender and so are considered to be autonomous.

Footnotes

Competing interests: none declared

1Lawson Wilkins Pediatric Endocrine Society

2European Society for Paediatric Endocrinology

References

  • 1.Frader J, Alderson P, Asch A.et al Health care professionals and intersex conditions. Arch Pediatr Adolesc Med 2004158426–429. [DOI] [PubMed] [Google Scholar]
  • 2.Conn J, Gillam L, Conway G. Revealing the diagnosis of androgen insensitivity syndrome in adulthood. BMJ 2005331628–630. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Dreger A D, Chase C, Sousa A.et al Changing the nomenclature/taxonomy for intersex: A scientific and clinical rationale. J Pediatr Endocrinol Metab 200518729–733. [DOI] [PubMed] [Google Scholar]
  • 4.Brown J, Warne G. Practical management of the intersex infant. J Pediatr Endocrinol Metab 2005183–23. [DOI] [PubMed] [Google Scholar]
  • 5.Cohen‐Kettenis P T. Gender change in 46,XY persons with 5‐alpha‐reductase‐2 deficiency and 17‐beta‐hydroxysteroid dehydrogenase‐3 deficiency. Arch Sex Behav 200534399–410. [DOI] [PubMed] [Google Scholar]
  • 6.Meyer‐Bahlburg H F. Gender identity outcome in female‐raised 46,XY persons with penile agenesis, cloacal exstrophy of the bladder, or penile ablation. Arch Sex Behav 200534423–438. [DOI] [PubMed] [Google Scholar]
  • 7.Zucker K J. Intersexuality and gender identity differentiation. Annu Rev Sex Res 1999101–69. [PubMed] [Google Scholar]
  • 8.Cohen‐Bendahan C C C, van de Beek C, Berenbaum S A. Prenatal sex hormone effects on child and adult sex‐typed behavior: methods and findings. Neurosci Biobehav Rev 200529353–384. [DOI] [PubMed] [Google Scholar]
  • 9.Meyer‐Bahlburg H F. Gender and sexuality in congenital adrenal hyperplasia. Endocrinol Metab Clin North Am 200130155–171. [DOI] [PubMed] [Google Scholar]
  • 10.Nordenström A, Servin A, Bohlin G.et al Sex‐typed toy play behavior correlates with the degree of prenatal androgen exposure assessed by CYP21 genotype in girls with congenital adrenal hyperplasia. J Clin Endo Metab 2002875119–5124. [DOI] [PubMed] [Google Scholar]
  • 11.Goy R W, Bercovitch F B, McBrair M C. Behavioral masculinization is independent of genital masculinization in prenatally androgenized female rhesus macaques. Horm Behav 198822552–571. [DOI] [PubMed] [Google Scholar]
  • 12.Wallen K. Hormonal influences on sexually differentiated behavior in nonhuman primates. Front Neuroendocrinol 2005267–26. [DOI] [PubMed] [Google Scholar]
  • 13.Moore C L. The role of maternal stimulation in the development of sexual behavior and its neural basis. Ann NY Acad Sci 1992662160–177. [DOI] [PubMed] [Google Scholar]
  • 14.Wallen K. Nature needs nurture: the interaction of hormonal and social influences on the development of behavioral sex differences in rhesus monkeys. Horm Behav 199630364–378. [DOI] [PubMed] [Google Scholar]
  • 15.De Vries G J, Rissman E F, Simerly R B.et al A model system for study of sex chromosome effects on sexually dimorphic neural and behavioral traits. J Neurosci 2002229005–9014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Skuse D H, James R S, Bishop D V M.et al Evidence from Turner's syndrome of an imprinted X‐linked locus affecting cognitive function. Nature 1997387705–708. [DOI] [PubMed] [Google Scholar]
  • 17.Hines M, Ahmed F, Hughes I A. Psychological outcomes and gender‐related development in complete androgen insensitivity syndrome. Arch Sex Behav 20033293–101. [DOI] [PubMed] [Google Scholar]
  • 18.Arnold A P, Rissman E F, De Vries G J. Two perspectives on the origin of sex differences in the brain. Ann NY Acad Sci 20031007176–188. [DOI] [PubMed] [Google Scholar]
  • 19.Luders E, Narr K, Thompson P M.et al Gender differences in cortical complexity. Nature Neurosci 20047799–800. [DOI] [PubMed] [Google Scholar]
  • 20.Paus T. Mapping brain maturation and cognitive development during adolescence. Trends Cogn Sci 2005960–68. [DOI] [PubMed] [Google Scholar]
  • 21.Consortium on the Management of Disorders of Sex Differentiation Clinical guidelines: for the management of disorders of sex development in childhood. 2006 ( www.dsdguidelines.org )
  • 22.Lee P A. A perspective on the approach to the intersex child born with genital ambiguity. J Pediatr Endocrinol Metab 200417133–140. [DOI] [PubMed] [Google Scholar]
  • 23.American Academy of Pediatrics Council on Children with Disabilities. Care coordination in the medical home: integrating health and related systems of care for children with special health care needs, Pediatrics 20051161238–1244. [DOI] [PubMed] [Google Scholar]
  • 24.Cashman S, Reidy P, Cody K.et al Developing and measuring progress toward collaborative, integrated, interdisciplinary health teams. J Interprof Care 200418183–196. [DOI] [PubMed] [Google Scholar]
  • 25.Warne G. Support groups for CAH and AIS. Endocrinologist 200313175–178. [Google Scholar]
  • 26.Feldman K W, Smith D W. Fetal phallic growth and penile standards for newborn male infants. J Pediatr 197586395–398. [DOI] [PubMed] [Google Scholar]
  • 27.Schonfield W A, Beebe G W. Normal growth and variation in the male genitalia from birth to maturity. J Urol 194248759–777. [Google Scholar]
  • 28.Fujieda K, Matsuura N. Growth and maturation in the male genitalia from birth to adolescence. II Change of penile length. Acta Paediatr Japan 198729220–223. [DOI] [PubMed] [Google Scholar]
  • 29.Tuladhar R, Davis P G, Batch J.et al Establishment of a normal range of penile length in preterm infants. J Paediatr Child Health 199834471–473. [DOI] [PubMed] [Google Scholar]
  • 30.Cheng P K, Chanoine J P. Should the definition of micropenis vary according to ethnicity? Horm Res 200155278–281. [DOI] [PubMed] [Google Scholar]
  • 31.Zachmann M, Prader A, Kind H P.et al Testicular volume during adolescence: cross‐sectional and longitudinal studies. Helv Paediatr Acta 19742961–72. [PubMed] [Google Scholar]
  • 32.Oberfield S E, Mondok A, Shahrivar F.et al Clitoral size in full‐term infants. Am J Perinatol 19896453–454. [DOI] [PubMed] [Google Scholar]
  • 33.Verkauf B S, Von Thron J, O'Brien W F. Clitoral size in normal women. Obstet Gynecol 19928041–44. [PubMed] [Google Scholar]
  • 34.Lloyd J, Crouch N S, Minto C L.et al Female genital appearance: “normality” unfolds. BJOG 2005112643–646. [DOI] [PubMed] [Google Scholar]
  • 35.Grumbach M M, Hughes I A, Conte F A. Disorders of sex differentiation. In: Larsen PR, Kronenberg HM, Melmed S, et al, editors. Williams textbook of endocrinology, 10th edition. Philadelphia: WB Saunders, 2003842–1002.
  • 36.Ahmed S F, Cheng A, Dovey L.et al Phenotypic features, androgen receptor binding, and mutational analysis in 278 clinical cases reported as androgen insensitivity syndrome. J Clin Endocrinol Metab 200085658–665. [DOI] [PubMed] [Google Scholar]
  • 37.Morel Y, Rey R, Teinturier C.et al Aetiological diagnosis of male sex ambiguity: a collaborative study. Eur J Pediatr 200216149–59. [DOI] [PubMed] [Google Scholar]
  • 38.Ogilvy‐Stuart A L, Brain C E. Early assessment of ambiguous genitalia. Arch Dis Child 200489401–407. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Quillin J M, Jackson‐Cook C, Bodurtha J. The link between providers and patients: how laboratories can ensure quality results with genetic testing. Clin Leadership Manag Rev 200317351–357. [PubMed] [Google Scholar]
  • 40.Pagon R A, Tarczy‐Hornoch P, Baskin P K.et al GeneTests‐Gene Clinics: genetic testing information for a growing audience. Hum Mut 200219501–509. [DOI] [PubMed] [Google Scholar]
  • 41.Dessens A B, Slijper F M, Drop S L. Gender dysphoria and gender change in chromosomal females with congenital adrenal hyperplasia. Arch Sex Behav 200532389–397. [DOI] [PubMed] [Google Scholar]
  • 42.Mazur T. Gender dysphoria and gender change in androgen insensitivity or micropenis. ArchSex Behav200534411–421. [DOI] [PubMed] [Google Scholar]
  • 43.Clayton P E, Miller W L, Oberfield S E, Ritzen E M, Speisser P W, ESPE/LWPES CAH Working Group Consensus statement on 21‐hydroxylase deficiency from the Lawson Wilkins Pediatric Endocrine Society and the European Society for Paediatric Endocrinology. Horm Res 200258188–195. [DOI] [PubMed] [Google Scholar]
  • 44.Nicolino M, Bendelac N, Jay N.et al Clinical and biological assessments of the undervirilized male. BJU Int 200493(suppl 3)20–25. [DOI] [PubMed] [Google Scholar]
  • 45.Mendonca B B, Inacio M, Costa E M F.et al Male pseudohermaphroditism due to 5 alpha‐reductase 2 deficiency: outcome of a Brazilian Cohort. Endocrinologist 200313202–204. [Google Scholar]
  • 46.Migeon C J, Wisniewski A B, Gearhart J P.et al Ambiguous genitalia with perineoscrotal hypospadias in 46,XY individuals: long‐term medical, surgical, and psychosexual outcome. Pediatrics 2002110e31. [DOI] [PubMed] [Google Scholar]
  • 47.Lee P A, Witchel S F. Genital surgery among females with congenital adrenal hyperplasias: Changes over the past five decades. J Pediatr Endocrinol Metab 2002151473–1477. [DOI] [PubMed] [Google Scholar]
  • 48.Rink R C, Adams M C. Feminizing genitoplasty: state of the art. World J Urol 199816212–218. [DOI] [PubMed] [Google Scholar]
  • 49.Farkas A, Chertin B, Hadas‐Halpren I. 1‐Stage feminizing genitoplasty: 8 years of experience with 49 cases. J Urol 20011652341–2346. [DOI] [PubMed] [Google Scholar]
  • 50.Baskin L S. Anatomical studies of the female genitalia: surgical reconstructive implications. J Pediatr Endocrinol Metab 200417581–587. [DOI] [PubMed] [Google Scholar]
  • 51.Crouch N S, Minto C L, Laio L M.et al Genital sensation after feminizing genitoplasty for congenital adrenal hyperplasia: a pilot study. BJU Int 200493135–138. [DOI] [PubMed] [Google Scholar]
  • 52.Meyer‐Bahlburg H F, Migeon C J, Berkovitz G D.et al Attitudes of adult 46,XY intersex persons to clinical management policies. J Urol 20041711615–1619. [DOI] [PubMed] [Google Scholar]
  • 53.American Academy of Pediatrics Timing of elective surgery on the genitalia of male children with particular reference to the risks, benefits, and psychological effects of surgery and anaesthesia. Pediatrics 199697590–594. [PubMed] [Google Scholar]
  • 54.Eroglu E, Tekant G, Gundogdu G.et al Feminizing surgical management of intersex patients. Pediatr Surg Int 200420543–547. [DOI] [PubMed] [Google Scholar]
  • 55.Alizai N, Thomas D F M, Lilford R J.et al Feminizing genitoplasty for congenital adrenal hyperplasia: what happens at puberty? J Urol 19991611588–1591. [PubMed] [Google Scholar]
  • 56.Bailez M M, Gearhart J P, Migeon C G.et al Vaginal reconstruction after initial construction of the external genitalia in girls with salt wasting adrenal hyperplasia. J Urol 1992148680–684. [DOI] [PubMed] [Google Scholar]
  • 57.Mouriquand P D, Mure P Y. Current concepts in hypospadiology. BJU Int 200493(suppl 3)26–34. [DOI] [PubMed] [Google Scholar]
  • 58.Bettocchi C, Ralph D J, Pryor J P. Pedicled phalloplasty in females with gender dysphoria. BJU Int 200595120–124. [DOI] [PubMed] [Google Scholar]
  • 59.Hurt W G, Bodurtha J N, McCall J B.et al Seminoma in pubertal patient with androgen insensitivity syndrome. Am J Obstet Gynecol 1989161530–531. [DOI] [PubMed] [Google Scholar]
  • 60.Rorth M, Rajpert‐De Meyts E, Andersson L.et al Carcinoma in situ in the testis. Scand J Urol Nephrol Suppl 2000205166–186. [DOI] [PubMed] [Google Scholar]
  • 61.Nihoul‐Fékété C. The Isabel Forshall Lecture. Surgical management of the intersex patient: an overview in 2003, J Pediatr Surg 200439144–145. [DOI] [PubMed] [Google Scholar]
  • 62.Warne G L, Grover S, Zajac J D. Hormonal therapies for individuals with intersex conditions: protocol for use. Treat Endocrinol 2005419–29. [DOI] [PubMed] [Google Scholar]
  • 63.Rogol A D. New facets of androgen replacement therapy during childhood and adolescence. Expert Opin Pharmacother 200561319–1336. [DOI] [PubMed] [Google Scholar]
  • 64.Ahmed S F, Tucker P, Mayo A.et al Randomized, crossover comparison study of the short‐term effect of oral testosterone undecanoate and intramuscular testosterone depot on linear growth and serum bone alkaline phosphatase. J Pediatr Endocrinol Metab 200417941–950. [DOI] [PubMed] [Google Scholar]
  • 65.Mayo A, Macintyre H, Wallace A M.et al Transdermal testosterone application: pharmacokinetics and effects on pubertal status, short‐term growth, and bone turnover. J Clin Endocrinol Metab 200489681–687. [DOI] [PubMed] [Google Scholar]
  • 66.Weidemann W, Peters B, Romalo G.et al Response to androgen treatment in a patient with partial androgen insensitivity and a mutation in the deoxyribonucleic acid binding domain of the androgen receptor. J Clin Endocrinol Metab 2000831173–1181. [DOI] [PubMed] [Google Scholar]
  • 67.Kazak A E, Cant M C, Jensen M M.et al Identifying psychosocial risk indicative of subsequent resource use in families of newly diagnosed pediatric oncology patients. J Clin Oncol 2003213220–3225. [DOI] [PubMed] [Google Scholar]
  • 68.Martin C L, Ruble D N, Szkrybalo J. Cognitive theories of early gender development. Psychol Bull 2002128903–933. [DOI] [PubMed] [Google Scholar]
  • 69.Zucker K J. Measurement of psychosexual differentiation. Arch Sex Behav 200534375–388. [DOI] [PubMed] [Google Scholar]
  • 70.Carmichael P, Ransley P. Telling children about a physical intersex condition. Dialogues Pediatr Urol 2002257–8. [Google Scholar]
  • 71.Committee on Paediatric A I D S, American Academy of Pediatrics Disclosure of illness status to children and adolescents with HIV infection. Pediatrics 1999103164–166. [DOI] [PubMed] [Google Scholar]
  • 72.Money J.Sex errors of the body and related syndromes: a guide to counselling children, adolescents, and their families, 2nd edition. Baltimore: Paul H Brookes Publishing Co, 1994
  • 73.Basson R, Leiblum S, Brotto L.et al Definitions of women's sexual dysfunction reconsidered: advocating expansion and revision. J Psychosom Obstet Gynecol 200324221–229. [DOI] [PubMed] [Google Scholar]
  • 74.Creighton S, Alderson J, Brown S, Minto C L. Medical photography: ethics, consent and the intersex patient. BJU Int 20028967–71. [DOI] [PubMed] [Google Scholar]
  • 75.Ursano R J, Bell C, Eth S.et al Work Group on ASD and PTSD; Steering Committee on Practice Guidelines. Practice guidelines for the treatment of patients with acute stress disorder and posttraumatic stress disorder. Am J Psychiatry 2004161(11 suppl)3–31. [PubMed] [Google Scholar]
  • 76.Kuhnle U, Bullinger M Outcome of congenital adrenal hyperplasia Pediatr Surg Int. 1997;12:511–515. doi: 10.1007/BF01258714. [DOI] [PubMed] [Google Scholar]
  • 77.Warne G, Grover S, Hutson J.et al Murdoch Childrens Research Institute Sex Study Group. A long‐term outcome study of intersex conditions. J Pediatr Endocrinol Metab 200518555–567. [DOI] [PubMed] [Google Scholar]
  • 78.Creighton S M. Long‐term outcome of feminization surgery: the London experience. BJU Int 200493(suppl 3)44–46. [DOI] [PubMed] [Google Scholar]
  • 79.Steiner E, Woernie F. Carcinoma of the neovagina: case report and review of the literature. Gynecol Oncol 200284171–175. [DOI] [PubMed] [Google Scholar]
  • 80.Schober J M. Long‐term outcomes of feminizing genitoplasty for intersex. In: Pediatric surgery and urology: long‐term outcomes. London: WB Saunders Co (in press),
  • 81.Wisniewski A B, Migeon C J, Meyer‐Bahlburg H F.et al Complete androgen insensitivity syndrome: long‐term medical, surgical, and psychosexual outcome. J Clin Endocrinol Metab 2000852664–2669. [DOI] [PubMed] [Google Scholar]
  • 82.Honecker F, Stoop H, de Krijger R R.et al Pathobiological implications of the expression of markers of testicular carcinoma in situ by fetal germ cells. J Pathol 2004203849–857. [DOI] [PubMed] [Google Scholar]
  • 83.Cools M, Van Aerde K, Kersemaekers A M.et al Morphological and immunohistochemical differences between gonadal maturation delay and early germ cell neoplasia in patients with undervirilization syndromes. J Clin Endocrinol Metab 2005905295–5303. [DOI] [PubMed] [Google Scholar]
  • 84.Cools M, Honecker F, Stoop H.et al Maturation delay of germ cells in trisomy 21 fetuses results in increased risk for the development of testicular germ cell tumors. Hum Pathol 200637101–111. [DOI] [PubMed] [Google Scholar]
  • 85.Ramani P, Yeung C K, Habeebu S S. Testicular intratubular germ cell neoplasia in children and adults with intersex. Am J Surg Pathol 1993171124–1133. [DOI] [PubMed] [Google Scholar]
  • 86.Hannema S E, Scott I S, Rajperts‐De Meyts E.et al Testicular development in the complete androgen insensitivity syndrome. J Pathol 2006208518–527. [DOI] [PubMed] [Google Scholar]
  • 87.Warne G L, Bhatia V. Intersex, East and West. In: Sytsma S, editor. Ethics and intersex. Berlin: Springer, 2006183–205.
  • 88.Achermann J C, Ozisik G, Meeks J J.et al Genetic causes of human reproductive disease. J Clin Endocrinol Metab 2002872447–2454. [DOI] [PubMed] [Google Scholar]
  • 89.Small C L, Shima J E, Uzumcu M.et al Profiling gene expression during the differentiation and development of the murine embryonic gonad. Biol Reprod 200572492–501. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Martin P L. Moving toward an international standard in informed consent: the impact of intersexuality and the internet on the standard of care. Duke J Gend Law Policy 20029135–169. [PubMed] [Google Scholar]
  • 91.Department of Health Reference guide to consent for examination or treatment. London, DoH, 2001 ( www.doh.gov.uk/consent )
  • 92.Sentencia SU‐337/99 May 12, 1999; T‐551/99, Aug 2 1999

Articles from Archives of Disease in Childhood are provided here courtesy of BMJ Publishing Group

RESOURCES