Find information on thousands of medical conditions and prescription drugs.

Pseudohermaphroditism

An intersexual or intersex person (or animal of any unisexual species) is one who is born with genitalia and/or secondary sex characteristics determined as neither exclusively male nor female, or which combine features of the male and female sexes. more...

Home
Diseases
A
B
C
D
E
F
G
H
I
J
K
L
M
N
O
P
Arthritis
Arthritis
Bubonic plague
Hypokalemia
Pachydermoperiostosis
Pachygyria
Pacman syndrome
Paget's disease of bone
Paget's disease of the...
Palmoplantar Keratoderma
Pancreas divisum
Pancreatic cancer
Panhypopituitarism
Panic disorder
Panniculitis
Panophobia
Panthophobia
Papilledema
Paraganglioma
Paramyotonia congenita
Paraphilia
Paraplegia
Parapsoriasis
Parasitophobia
Parkinson's disease
Parkinson's disease
Parkinsonism
Paroxysmal nocturnal...
Patau syndrome
Patent ductus arteriosus
Pathophobia
Patterson...
Pediculosis
Pelizaeus-Merzbacher disease
Pelvic inflammatory disease
Pelvic lipomatosis
Pemphigus
Pemphigus
Pemphigus
Pendred syndrome
Periarteritis nodosa
Perinatal infections
Periodontal disease
Peripartum cardiomyopathy
Peripheral neuropathy
Peritonitis
Periventricular leukomalacia
Pernicious anemia
Perniosis
Persistent sexual arousal...
Pertussis
Pes planus
Peutz-Jeghers syndrome
Peyronie disease
Pfeiffer syndrome
Pharmacophobia
Phenylketonuria
Pheochromocytoma
Photosensitive epilepsy
Pica (disorder)
Pickardt syndrome
Pili multigemini
Pilonidal cyst
Pinta
PIRA
Pityriasis lichenoides...
Pityriasis lichenoides et...
Pityriasis rubra pilaris
Placental abruption
Pleural effusion
Pleurisy
Pleuritis
Plummer-Vinson syndrome
Pneumoconiosis
Pneumocystis jiroveci...
Pneumocystosis
Pneumonia, eosinophilic
Pneumothorax
POEMS syndrome
Poland syndrome
Poliomyelitis
Polyarteritis nodosa
Polyarthritis
Polychondritis
Polycystic kidney disease
Polycystic ovarian syndrome
Polycythemia vera
Polydactyly
Polymyalgia rheumatica
Polymyositis
Polyostotic fibrous...
Pompe's disease
Popliteal pterygium syndrome
Porencephaly
Porphyria
Porphyria cutanea tarda
Portal hypertension
Portal vein thrombosis
Post Polio syndrome
Post-traumatic stress...
Postural hypotension
Potophobia
Poxviridae disease
Prader-Willi syndrome
Precocious puberty
Preeclampsia
Premature aging
Premenstrual dysphoric...
Presbycusis
Primary biliary cirrhosis
Primary ciliary dyskinesia
Primary hyperparathyroidism
Primary lateral sclerosis
Primary progressive aphasia
Primary pulmonary...
Primary sclerosing...
Prinzmetal's variant angina
Proconvertin deficiency,...
Proctitis
Progeria
Progressive external...
Progressive multifocal...
Progressive supranuclear...
Prostatitis
Protein S deficiency
Protein-energy malnutrition
Proteus syndrome
Prune belly syndrome
Pseudocholinesterase...
Pseudogout
Pseudohermaphroditism
Pseudohypoparathyroidism
Pseudomyxoma peritonei
Pseudotumor cerebri
Pseudovaginal...
Pseudoxanthoma elasticum
Psittacosis
Psoriasis
Psychogenic polydipsia
Psychophysiologic Disorders
Pterygium
Ptosis
Pubic lice
Puerperal fever
Pulmonary alveolar...
Pulmonary hypertension
Pulmonary sequestration
Pulmonary valve stenosis
Pulmonic stenosis
Pure red cell aplasia
Purpura
Purpura, Schoenlein-Henoch
Purpura, thrombotic...
Pyelonephritis
Pyoderma gangrenosum
Pyomyositis
Pyrexiophobia
Pyrophobia
Pyropoikilocytosis
Pyrosis
Pyruvate kinase deficiency
Uveitis
Q
R
S
T
U
V
W
X
Y
Z
Medicines

(The terms hermaphrodite and pseudohermaphrodite, which have been used in the past, are now considered pejorative and inaccurate and are no longer used to refer to an intersexual person.) Sometimes the phrase "ambiguous genitalia" is used.

Overview

According to the highest estimates (Fausto-Sterling et. al., 2000) perhaps 1 percent of live births exhibit some degree of sexual ambiguity , and that between 0.1% and 0.2% of live births are ambiguous enough to become the subject of specialist medical attention, including surgery to disguise their sexual ambiguity. Other sources (Leonard Sax, 2002) estimate the incidence of true intersexual conditions as far lower, at approximately 0.018%.

In typical fetal development, the presence of the SRY gene causes the fetal gonads to become testes; the absence of it allows the gonads to continue to develop into ovaries. Thereafter, the development of the internal reproductive organs and the external genitalia is determined by hormones produced by certain fetal gonads (ovaries or testes) and the cells' response to them. The initial appearance of the fetal genitalia (a few weeks after conception) is basically feminine: a pair of "urogenital folds" with a small protuberance in the middle, and the urethra behind the protuberance. If the fetus has testes, and if the testes produce testosterone, and if the cells of the genitals respond to the testosterone, the outer urogenital folds swell and fuse in the midline to produce the scrotum; the protuberance grows larger and straighter to form the penis; the inner urogenital swellings swell, wrap around the penis, and fuse in the midline to form the penile urethra.

Because there is variation in all of these processes, a child can be born with a sexual anatomy that is typically female, or feminine in appearance with a larger than average clitoris; or typically male, masculine in appearance with a smaller than average penis that is open along the underside. The appearance may be quite ambiguous, describable as female genitals with a very large clitoris and partially fused labia, or as male genitals with a very small penis, completely open along the midline ("hypospadic"), and empty scrotum.

There are dozens of named medical conditions that may lead to intersex anatomy. Fertility is variable. The distinctions "male pseudohermaphrodite", "female pseudohermaphrodite" and especially "true hermaphrodite" are vestiges of 19th century thinking that placed "true sex" in the histology (microscopic appearance) of the gonads.

The common habit in the 21st century of elevating the role of the sex chromosomes above all other factors when determining gender may be analogous to the older habit of finding "true" sex in the gonads. Though high school biology teaches that men have XY and women XX chromosomes, in fact there are quite a few other possible combinations such as Turner_syndrome XO, Triple-X syndrome XXX, Klinefelter's Syndrome XXY, XYY, XO/XY, XX male, Swyer syndrome XY female, and there are many individuals who do not follow the typical patterns (such as cases with four or even more sex chromosomes).

Read more at Wikipedia.org


[List your site here Free!]


Intersexuality and gender identity differentiation
From Annual Review of Sex Research, 1/1/99 by Zucker, Kenneth J

People born with physical intersex conditions (often better known as hermaphroditism) remind us that the ordinary or "normal" process of physical sex differentiation is by no means automatic.1 Throughout the course of human history, all cultural groups have had to make multiple decisions about newborns with physical hermaphroditism (at least for those who did not die from medical complications inherent to, or associated with, their condition) and their subsequent place within the fabric of social life. At times, this matter has been expressed in an almost urgent manner: Tuffier and Lapointe (1911), for example, wrote that "For hermaphrodites as well as for normal subjects, the possession of a [single] sex is a necessity of our social order" (p. 256, cited and translated in Dreger, 1995a).

Over the past 2 centuries, there have been considerable advances in developing a more accurate scientific taxonomy of physical intersex conditions, although Shearman (1982) quipped that one way to tax a taxonomist is to "ask him to classify intersexes" (p. 325). Over the past 50-60 years, there has also been a great deal of interest in human hermaphroditism from a behavioral and psychological point of view, particularly with regard to what it can teach us about normative or typical psychosexual development and differentiation. Along parallel lines, many researchers in the field of animal sexology have relied on the experimental induction of "pseudohermaphroditism" in order to understand better the mechanisms and processes underlying normative sexdimorphic behavioral differentiation (e.g., Baum, 1979; Dixson, 1998; Goy & McEwen, 1980; Wallen, 1996). Indeed, the study of hermaphroditism has been generally guided by the assumption that understanding atypical sexual differentiation (from both a physical and behavioral point of view) will enhance understanding of more typical sexual differentiation. In this respect, research on hermaphroditism is a prototype example of the interface between the study of normal and abnormal development (cf. Cicchetti, 1993; Sroufe, 1990).

At present, the clinical and scientific literature on physical intersex conditions is in a state of great debate and controversy. One might even go so far as to say that the field is in a sort of crisis. There are, for example, considerable concerns and objections about the use of surgical interventions to "normalize" the ambiguous genitalia of people with physical intersex conditions (see, e.g., Chase, 1998a; Diamond & Sigmundson, 1997a; Schober, 1998a, 1998b, 1999; Wilson & Reiner, 1998), and some critics have even gone so far as to call for a "moratorium" on surgical interventions until the status of previously treated patients are followed-up with greater precision (e.g., Diamond, 1999; Kipnis & Diamond, 1998; cf. American Academy of Pediatrics, 1996; Glassberg, 1998). Indeed, in May and August 1999, the Constitutional Court of Colombia in South America provided a potentially landmark ruling on what constitutes informed consent for surgical interventions for infants and children with physical intersex conditions (Intersex Society of North America, 1999; see also Greenberg, 1999). Although guidelines for psychological counselling have been available for some time (e.g., Money, 1965, 1994; see also Green, Stoller, & MacAndrew, 1966; MeyerBahlburg, 1982, 1993; Stoller, Garfinkel, & Rosen, 1962), there has been a resurgence of discussion about the uneven quality of psychological counselling that is available to people with physical intersex conditions and their families (e.g., Preves, 1998). Some of the current controversy has also been fueled by adults with physical intersex conditions who are quite critical about the care that they received as children and adolescents (see, e.g., Chase, 1998b; Holmes, 1994). Other aspects of the current debate have been stimulated by critics of the "medical model," who utilize a "social constructionist" approach in appraising the extant literature (e.g., Dreger, 1995a, 1995b, 1998a, 1998b, 1998c; Fausto-Sterling, 1993; Findlay, 1995; Kessler, 1990, 1998; Lee, 1994). Indeed, many of these issues were recently discussed and debated at a 1999 conference in Dallas, Texas entitled "Pediatric Gender Reassignment: A Critical Reappraisal" (see Zderic, Canning, Snyder, & Carr, in press).

In this article, I will begin by providing a brief overview of relevant terminology. I will then review one specific aspect of the contemporary literature on intersexuality, namely what is known about gender identity formation and differentiation. I have chosen to focus the review on this specific topic for a couple of reasons. First, there has been, over the years, periodic debate regarding the question of the "appropriate" sex assignment and subsequent gender of rearing of infants and children with physical intersex conditions (see, e.g., Diamond, 1965; ImperatoMcGinley, Guerrero, Gautier, & Peterson, 1974; Zuger, 1970), primarily in relation to the formulation originally introduced and then later amplified by the work of John Money and his colleagues at Johns Hopkins Hospital beginning in the 1950s (e.g., Money, 1952; Money, Hampson, & Hampson, 1955). Second, over the last few years, this debate has re-surfaced, with even greater force and intensity, following the publication of long-term psychosexual follow-up data pertaining to a specific case claimed to have especially important relevance to this issue (Diamond & Sigmundson, 1997b).

As it has in the past, much of this debate continues to revolve around the relative roles of nature and nurture in gender identity formation in particular and psychosexual differentiation in general. Although many scholars in the field have long rejected the nature-nurture axis as a false dichotomy (see, e.g., Bancroft, 1991; Money, 1985; Wallen, 1996), the fact remains that in practice the debate carries on. It is timely, therefore, to provide a critical summary of the empirical database on gender identity formation in people with physical intersex conditions and then to evaluate the candidate explanations that best account for the pattern of results.

Terminology

Webster's Seventh New Collegiate Dictionary (1963) defines an intersexual as a person who is "intermediate in sexual characters between a typical male and a typical female" (p. 443). A hermaphrodite is defined a bit more narrowly as "an animal or plant having both male and female reproductive organs" (p. 389). According to Dreger (1995a), it was Goldschmidt (1923) who first used the term intersexuality as part of an effort to describe a range of physical sex ambiguities. Over time in the medical literature, intersexuality has become the preferred term used to encompass the diverse class of syndromes characterized by some abnormality or anomaly in the process of physical sex differentiation (Haqq & Donahoe, 1998; Shearman, 1982; Sizonenko, 1993). Thus, the term intersexuality can be understood to represent a broader class of syndromes than those that have been traditionally subsumed under terms like true hermaphroditism, female pseudohermaphroditism, and male pseudohermaphroditism, which were used in one of the first taxonomic efforts to classify physical intersex conditions (Klebs, 1876).

Efforts to devise accurate taxonomic systems are, of course, dependent on our knowledge of the multiple parameters that constitute biological sex (Table 1). Thus, for example, in Klebs' time, sex chromosome abnormalities were not part of any system of classification as it was only in the 1950s that reliable techniques were developed to karyotype the sex chromosomes (Moore & Barr, 1955). Klebs' proposed system of classification placed great emphasis on the importance of certain diagnostic techniques, including clinical microscopy, laparotomy, and biopsy (Dreger, 1995a). The insistence on histologic analysis appeared to have resulted in the nature of the gonads becoming the final arbiter in deciding upon the "true sex" of the hermaphrodite, a development that had substantial implications. As noted by Meyer-Bahlburg (1998), among others, the notion of a "true sex" in intersexuality is problematic. Given that a person's physical sex is multidimensional in nature, there is no reason to insist that one parameter should necessarily hold precedence over another. Thus, from a descriptive point of view all that is required is an accurate delineation of the physical sex parameters that are affected in particular intersex syndromes.

Which Sex? Which Gender?

A common aspect of several physical intersex conditions involves the differentiation of ambiguous external genitalia. When this occurs, there is often uncertainty whether the neonate's sex assignment should be that of a male or a female and the gender assignment that of a boy or a girl. Not surprisingly, such uncertainty causes anxiety in parents and in the professional involved in determining a newborn's sex, whether that professional is a physician or nurse working in the modern hospital delivery room or a midwife working in some remote "third world" community far removed from the postmodern Western scene. Perhaps reminiscent of Tuffier and Lapointe's (1911) remark quoted above, many contemporary physicians characterize this uncertainty as a "medical" and "psychosocial" emergency that requires immediate attention and resolution (e.g., Izquierdo & Glassberg, 1993; Pagon, 1987).

That the visible physical markers of biological sex are psychologically salient has even been documented by some empirically minded scientists. In one study of parental behavior shortly after the birth of a newborn, Woollett, White, and Lyon (1982) observed that the majority of the verbal comments pertained to the infant's sex. In another study, IntonsPeterson and Reddel (1984) had "Parent-collaborators" call their friends following the birth of their babies. Overall, 80% of the initial questions were about the baby's sex. The single most frequently asked question was "Is it a boy or a girl?"

In an interesting analogue study, Delk, Madden, Livingston, and Ryan (1986) showed that perceptions of a toddler's behavior were affected by labeling it as a male, a female, or a hermaphrodite. Several hundred health professional trainees (e.g., medical and psychology students) viewed an 8-minute videotape of a 22-month-old infant engaged in various activities. Every 15 seconds, the trainees were asked to rate the toddler's last activity as masculine, feminine, or neutral. Activities were more likely to be rated as masculine than feminine when the toddler was labeled a male, whereas the converse occurred when the toddler was labeled a female. When the toddler was labeled a hermaphrodite, a similar proportion of the activities was rated as masculine and feminine.

Sex and gender assignment at birth are believed to be the first of a cascade of events that fall under the rubric of gender socialization (Ruble & Martin, 1998); nowadays, with the development of techniques such as amniocentesis and ultrasound, parents can acquire information about fetal sex, which likely generates a variety of specific feelings and thoughts about their future child (e.g., Birnholz, 1983; Fletcher & Evans, 1983; Winestine, 1989). Following these first events, whether they occur prenatally or after parturition, parents often select a name for their newborn that has a stereotypical masculine or feminine connotation. Many books are available to aid parents in these selections, the popular press routinely reports on the most common given names of boys and girls (Hartocollis, 1999; Roberts, 1996), and there are scholars who actually study the psychology and sociology of naming (Lieberson & Bell, 1992).

It is also common for parents to dress male and female infants in sexstereotypical ways, including the North American tradition of sexdimorphic "color coding" in pink or blue that began in the 1920s (Paoletti & Thompson, 1987). Shakin, Shakin, and Sternglanz (1985) observed infants at a shopping mall in Long Island, New York, and found that about 75% of the females had at least some pink in their clothing, compared to 0% of the males, and that 79% of the males had at least some blue in their clothing compared to only 8% of females.

Sex assignment and subsequent "rearing" as a boy or girl have long been viewed as powerful socialization influences that account for sex differences in psychosexual differentiation. Thorne (1993), for example, writes that

While many still see gender as the expression of natural differences, the women's movement of the 1970s and 1980s launched a powerful alternative perspective: notions of femininity and masculinity, the gender divisions one sees on school playgrounds ... the idea of gender itself-all are social constructions.... Parents dress infant girls in pink and boys in blue, give them gender-differentiated names and toys, and expect them to act differently... peer groups . also perpetuate gender-typed play and interaction. In short, if boys and girls are different, they are not born, but made that way. (p. 2, italics in original)

In stark contrast to this view, theorists with a biological bent also emphasize single-factor influences. For example, Swaab, Gooren, and Hofman (1992) asserted that gender identity is very difficult to change, "probably because... [it is] fixed in the brain" (p. 52).

But because the rearing of an infant as a boy or a girl is usually perfectly confounded with biological sex, researchers have long made the point that it is actually difficult to disentangle the relative contribution of biological and psychosocial influences. For some researchers, it was this methodological and interpretive dilemma that led to the study of children with physical intersex conditions in the hope of providing at least a partial resolution to this problem.

Initial Empirical Studies: The Work of Money and Colleagues

Beginning in the 1950s, Money and colleagues began to report data on the psychosexual development of children born with physical intersex conditions. Money, Hampson, and Hampson (1957) noted that since hermaphrodites are "neither exclusively male or female, [they] are likely to grow up with contradictions existing between the sex of assignment and rearing, on the one hand, and various physical sexual variables, singly or in combination, on the other" (p. 333). Thus, Money et al. asked "whether the gender role and orientation2 that a hermaphrodite establishes during the course of growing up is concordant with the sex of assignment and rearing, or whether it is predominantly concordant with one or another of the ... physical sexual variables" (p. 333).

In their study of 105 hermaphrodites, Money et al. (1957) found that only 5 of 105 patients had a "gender role and orientation [that] was ambiguous and deviant from the sex of assignment and rearing" (p. 333). Thus, Money et al. concluded that "the sex of assignment and rearing is consistently and conspicuously a more reliable prognosticator of a hermaphrodite's gender role and orientation than is the chromosomal sex, the gonadal sex, the hormonal sex, the accessory internal reproductive morphology, or the ambiguous morphology of the external genitalia" (p. 333). This conclusion echoed that of Ellis (1945), who had previously reviewed the available literature on 84 cases of hermaphroditism:

The hermaphrodite assumes a heterosexual libido and sex role that accords primarily not with his or her internal and external somatic characteristics, but rather with his or her masculine or feminine upbringing .... heterosexuality and homosexuality in hermaphrodites are primarily caused not by directly hormonal or other physiological factors but by environmental ones. (p. 120)

Money et al. (1957) offered one additional finding for the relative importance of socialization factors in determining gender identity differentiation among children with physical intersex conditions:

The clinching piece of evidence concerning the psychologic importance of the sex of assignment and rearing is provided when, among persons of identical physical diagnosis, some are reared as boys, some as girls. It is indeed startling to see, for example, two children with female [congenital adrenal hyperplasial in the company of one another in a hospital playroom, one of them entirely feminine in behavior and conduct, the other entirely masculine, each according to upbringing. (p. 334)

Given such evidence for the apparent malleability and plasticity in gender identity differentiation (see also Money, 1991), it became necessary to replace the reliance on identifying the patient's "true sex" with a different model for guiding decisions about gender assignment.

As summarized by Meyer-Bahlburg (1998), the model developed by Money and the Johns Hopkins school of pediatric endocrinology can be characterized as the optimal gender policy of psychosocial and medical management. This policy aimed to result in the best possible prognosis with regard to six variables, which are listed in Table 2.

Is There A Critical Period for Gender Identity Formation?

Money et al.'s (1957) conclusion about the relative salience of sex of rearing3 had one important caveat, namely the advisability of an early decision about sex assignment. They recommended that when there was uncertainty about the appropriate sex of assignment, the final decision about it should certainly be made no later than 18-24 months, and argued that "uncompromising adherence to the decision is desirable" (p. 334). Otherwise, it was claimed that the child would be vulnerable to "psychologic nonhealthiness" (p. 334), which presumably was related to a more conflicted or ambiguous gender identity, which Stoller (1964) later referred to as a "hermaphroditic" gender identity. In part, Money et al.'s (1957) recommendation of an early sex assignment was based on the observation that among intersexed children who experienced a sex reassignment after the neonatal period, 11 of 14 children adjusted to the change without complications if the reassignment occurred prior to 27 months, in contrast to only 1 of 4 children who adjusted to the change without complications if it occurred after 27 months (Fisher's exact test, p = .0379, one-tailed, my analysis).

Money et al. (1957) interpreted this age effect as evidence for a process akin to the phenomenon of imprinting, or a critical period, that had been described by ethologists with regard to other behavioral characteristics, such as attachment formation (eg., Bowlby, 1969). Thus, the first 2 years or so of life were deemed to represent a critical period for gender identity formation (see also Money & Annecillo, 1987).

Since the 1950s, the critical-period construct has been subject to a great deal of general empirical scrutiny, and the concept of a "sensitive" or an "optimal" period was introduced in order to expand the window of time in which certain environmental experiences might exert their greatest impact, but without implicating the irreversibility that was believed to occur in the case of critical periods (e.g., Bornstein, 1987, 1989; Hess, 1973).

The existence of a sensitive period for gender identity formation suggests a certain malleability or plasticity, in which gender identity differentiation can more readily move in one direction or another but, after some imprecisely defined window of time, this becomes more difficult.

Even if we put the concept of a sensitive period aside, normative data that have accrued over the past several decades on gender identity formation certainly suggest that important developments occur in the first 2 or 3 years (e.g., Fagot, 1995; Ruble & Martin, 1998). Let us consider what is known about the development of the capacity to correctly self-label one's gender.

There is now some evidence that infants under the age of 12 months are able to perceptually discriminate males from females, relying on cues such as hair-style and voice quality. Although it is unlikely that infants have any conscious or reflective understanding of these distinctions, the data suggest that phenotypic markers that commonly distinguish males from females are salient to the infant. Between the second and third year of life, nonverbal techniques (e.g., sorting tasks) indicate that toddlers can correctly self-label their own gender and that of others and, by age 3, can answer correctly the question "Are you a boy or a girl?" Thus, the corpus of extant normative developmental data certainly support Money's original observations that signs of gender identity differentiation can be detected by age 2 years. Interestingly enough, however, it is not clear to what extent children rely on the configuration of the external genitalia in making judgments about their own gender. Indeed, some have argued that children pay more attention to markers such as hair-style and clothing style than they do to the configuration of the external genitalia, perhaps because they have more comparative cue exposure to the former than to the latter (de Marneffe, 1997; McConaghy, 1979), and it is only later in childhood that the external genitalia are used more systematically in the correct gender labeling of self and others (McConaghy, 1979).

In any case, the data suggest that a gender reassignment after the age of 2 or 3 years would require that the typical toddler unlearn his or her original gender self-label and learn another one. It is not clear how difficult such a task might be. For example, Kohlberg (1966) argued that although toddlers and preschoolers display a rudimentary cognitive understanding of gender, they do not truly appreciate its invariance: "the child age two to four is very uncertain of the constancy of his [gender] identity, and the label 'boy' is for him as arbitrary as the label 'Johnny"' (P. 87). In Kohlberg's (1966) now classical cognitive-developmental account of gender constancy development, the child's eventual understanding that gender is an invariant part of the self is attained only with the development of concrete-operational thought (around 5-7 years of age), which permits the child to appreciate the principle of invariance in the face of "superficial" or surface transformations in gender-related behavior, such as activity preferences or clothing style (for reviews of the gender constancy literature, see Ruble & Martin, 1998; Zucker et al., in press). It is unlikely that one could conduct an experiment in which, for example, parents of ordinary toddlers try to teach males that they are girls, not boys, and females that they are boys, not girls, after they have acquired an initial correct self-labeling. Perhaps one could try a re-labeling experiment around something more innocuous (e.g., to re-teach a toddler that his nose is really his ear and viceversa). To this, of course, one also has to add the cognitive complexity required of parents to change sets around their child's sex and gender. One would imagine that this would become a more difficult task after raising a child as one gender for a couple of years, rather than the issue being settled in the first few days or months after birth.

Appraisal of the Gender Identity Formation Data

To what extent have the original Money et al. (1957) data, as well as Ellis' (1945) earlier conclusion, been substantiated by subsequent research on gender identity formation in children with physical intersex conditions? To answer this question, one can take advantage of the increased precision in identifying physical intersex conditions on a syndrome by syndrome basis. Moreover, one can consider the nature of the syndromes themselves and which aspects of physical sex differentiation are affected in each.

As a point of departure, it should be noted that certain physical intersex conditions do not result in any substantive uncertainty about sex assignment and gender of rearing at the time of birth. Let us consider two such examples. In genetic males born with the complete form of the androgen insensitivity syndrome (cAIS) (Morris, 1953; Morris & Mahesh, 1963; Perez-Palacios, Chavez, Mendez, Imperato-McGinley, & UlloaAguirre, 1987; Rutgers & Scully, 1991), which is inherited as an X-linked recessive trait (for recent accounts of molecular genetic studies, see Brown et al., 1993; Griffin, 1992; Quigley et al., 1995; Radmayr et al., 1997), the external genitalia differentiate as female because of a defect in androgen receptors. Accordingly, the baby is assigned to the female sex and reared as a girl. At puberty, there is relatively normal breast development (induced by testicular estrogens), but there is a lack of certain secondary sex characteristics, such as axillary hair and pubic hair. The testes remain undescended and the vagina is blind-ended. There is no cervix, uterus, or fallopian tubes. Clinical presentation often occurs for the first time during adolescence because of the absence of menarche.4

In the case of cAIS, the available empirical studies and case reports indicate an unequivocal female gender identity (e.g., Costa et al., 1997; Hampson, Hampson, & Money, 1955; Imperato-McGinley, Pichardo, Gautier, Voyer, & Bryden, 1991; Masica, Money, & Ehrhardt, 1971; Slijper, Drop, Molenaar, & de Muinck Keizer-Schrama, 1998; see also Money, Ehrhardt, & Masica, 1968; Slob et al., 1993). Indeed, MeyerBahlburg (1999b) has pointed out that the literature does not appear to contain any report of a person with cAIS raised as a female who later voluntarily changed her gender to that of a male. Thus, a female gender identity differentiates in a person with XY chromosomes and testicular gonads.

From a psychosocial standpoint, one might infer that cAIS individuals who are assigned unequivocally to the female sex and reared as girls are subject to the same gender socialization influences as are unaffected girls. From a biological standpoint, it has been argued that the androgen resistance results in female-typical brain differentiation, which also plays a putative role in sex-dimorphic psychosexual differentiation (Collaer & Hines, 1995). Taken together, then, the putative psychosocial and biological mechanisms that produce "normal" or typical psychosexual differentiation act in concert among newborns with cAIS.5

The second example concerns males born with hypospadias, a genital anomaly. In this condition, the opening of the urethra appears at various levels on the undersurface of the penis and its position is used to classify the nature of the hypospadic anomaly (e.g., glandular, penile, penoscrotal, and perineal). The typical corrective surgery aims at reconstructing the distal urethra, which permits voiding in the standing position, facilitates sexual functioning in adolescence and adulthood, and "normalizes" the appearance of the penis. Unless the anomaly co-occurs with other abnormalities in genital differentiation (e.g., as in the micropenis syndromes), there is usually little uncertainty about sex or gender assignment at birth. In boys born with "pure forms" of hypospadias, the literature suggests an unequivocal male gender identity (Mureau, Slijper, Slob, & Verhulst, 1997; Sandberg, Meyer-Bahlburg, Aranoff, Sconzo, & Hensle, 1989; Sandberg et al., 1995).

From a psychosocial standpoint, one might infer that boys born with hypospadias are subject to the same gender socialization influences as are unaffected boys. From a biological standpoint, to the extent that some type of prenatal androgen insufficiency contributes to the condition, it would appear that such effects are relatively weak, in the sense that the penis is relatively well-differentiated, and there is no compelling evidence to believe that the putative androgen insufficiency has affected sex-dimorphic brain differentiation (Collaer & Hines, 1995). Taken together, the putative psychosocial and biological mechanisms that produce "normal" psychosexual differentiation appear to be relatively intact among boys born with hypospadias.

It is with regard to those physical intersex conditions in which there might be some uncertainty at birth regarding sex assignment that the relative importance of gender socialization can best be evaluated. In genetic females, the syndrome of congenital adrenal hyperplasia (CAH) is most relevant; in genetic males, the relevant syndromes include steroid 5alpha-reductase 2 deficiency (5-ARD), partial androgen insensitivity syndrome (pAIS), micropenis, penile agenesis (aphallia), and cloacal exstrophy. Regarding gender identity differentiation, these syndromes share two characteristic features: (a) there may be some uncertainty regarding sex assignment at birth, in part because the configuration of the external genitalia is severely affected and, as a result, there may be some uncertainty regarding the "optimal" gender in which the child should be reared; (b) either the prenatal hormonal milieu or the configuration of the external genitalia (and sometimes both) can be atypical in relation to the gender in which the child is reared.

Congenital Adrenal Hyperplasia in Genetic Females

In genetic females with CAH, the overproduction of androgenic steroids during fetal development, which has been documented from amniotic fluid assays in at-risk pregnancies (Carson et al., 1982; Forest, 1985), causes genital masculinization ranging from mild clitoral enlargement to complete fusion of the labioscrotal folds with a phallic urethra (New, Ghizzoni, & Speiser, 1996). It is this aspect of the syndrome that, at times, creates uncertainty with regard to sex assignment at birth. When the condition is properly diagnosed, several medical interventions typically ensue, including surgical "feminization" of the enlarged clitoris (Allen, Hardy, & Churchill, 1982; Donahoe & Schnitzer, 1996; Randolph, Hung, & Rathlev, 1981) and cortisolreplacement therapy to control or eliminate postnatal virilization (New & Josso, 1988; Wilkins, Lewis, Klein, & Rosemberg, 1950). Under these conditions, a female sex assignment is made and the infant is, invariably, raised as a girl. For further details on the syndrome itself, including recent understanding of its molecular genetic basis, see Pang (1997), Wedell (1998), and Wilson, Mercado, Cheng, and New (1995).

Gender Identity Differentiation in Childhood

What do we know about the gender identity development of girls with CAH raised under these conditions? Ehrhardt, Epstein, and Money (1968) compared 15 girls with CAH and 15 control girls with regard to a variety of sex-dimorphic behaviors (M age, 10.5 years; range, 5-16 years). Based on interview data regarding gender identity, 7 (47%) of the CAH girls were classified as "content or prefers to be a girl," 5 (33%) were classified as "ambivalent," and 3 (20%) were judged to "[desire] expressly to be a boy." The corresponding percentages for the controls were 93%, 0%, and 7%, respectively. Thus, there was some evidence that girls with CAH were less content with their gender identity than were the controls; however, Ehrhardt et al. ro-marked that only one of the CAH girls appeared to be severely gender dysphoric and whose general psychosocial functioning was markedly impaired. In a similar study, Ehrhardt and Baker (1974) asked their youngsters whether it was better to be a girl or a boy. Of 17 girls with CAH (age range, 4.3-19.9 years), 6 (35%) indicated that they were undecided or thought that they might have chosen to be a boy if such a choice had been possible. In contrast, only 1 (9%) of 11 sisters gave a similar response. Ehrhardt and Baker noted, however, that "none of the [CAH] girls had a conflict with her female gender identity or was unhappy about being a girl" (p. 43).

The gender identity of girls with CAH has also been evaluated in two more recent studies. Slijper et al. (1998; Slijper, personal communication, April 18, 1999) assessed 18 girls with CAH (M age, 13.5 years; range, 2-27 years). Of these, 10 were assigned to the female sex at birth, but 8 others were initially assigned to the male sex (and subsequently reassigned to the female sex no later than age 6 months). Of the 18 girls, 2 (11%) were judged to meet the DSM-IV criteria for gender identity disorder (GID) (American Psychiatric Association, 1994). The remaining 16 girls were deemed reasonably content with their female gender identity.

Berenbaum and Bailey (1998) studied 31 girls with CAH, an adrecruited group of 7 "tomboys" without known somatic intersexuality, and 22 unaffected sisters of both CAH girls and boys, and of the tomboys (M age, 10.9 yrs). A 9-item interview schedule assessed what was termed "continuous gender identity" (p. 9). On this measure, there was little evidence that the girls with CAH were uncomfortable being female. Item analysis indicated that the girls with CAH were more similar to the sister-control group, but different from that of the tomboys. For the continuous measure, the girls with CAH had a mean score that was in between that of the other two groups. Berenbaum and Bailey concluded that their data confirmed earlier reports that girls with CAH "have female-typical gender identity" (p. 12).

Gender Identity Differentiation in Adulthood

Adult follow-up of women with CAH provides a more definitive picture with regard to gender identity differentiation. Over the years, there have been several follow-up reports pertaining to the gender identity development of women with CAH; in addition, inferences about gender identity development in adulthood can be gleaned from reports in which there is cursory mention of gender identity (typically, in the context of medical or surgical aspects of CAH).

In one study, Zucker et al. (1996) assessed the gender identity of 31 women with CAH (M age, 24.4 years) and 15 sister/female cousin controls (M age, 25.6 years). Gender identity was assessed via a semistruetured interview and by a Gender Dysphoria/Identification self-report questionnaire.

At the time of assessment, all of the probands were living, in the broadest sense, as women (i.e., they were known to others as females and were registered as such on legal or other official documents). For the interview ratings of current and lifetime gender dysphoria, the proband-control comparisons were not significant. On the self-report questionnaire, the two groups did not differ on the factor labeled Gender Dysphoria.

Although these data did not provide any clear evidence for gender dysphoria or discontent among the CAH probands, it should be noted that there were 10 additional potential probands who refused to participate in the study and 13 others could not be traced (including one who had died, and two who were raised as boys from infancy by parental decision). To test for selective attrition factors, we examined three variables: current age, sex assignment at birth, and salt-wasting status.6 The participants did not differ from the nonparticipants with regard to current age and the percentage assigned to the female sex at birth; however, the refusers were all classified as salt-wasters compared to 61% of the participants (Fisher's exact test, p = .0412, two-tailed). The untraced probands did not differ from the participants on this variable (46% vs. 61%).

By virtue of an independent clinical referral, one of the refusers (age, 19 years) had been previously assessed by me (in another hospital setting) because of extreme gender dysphoria. This proband was diagnosed with Transsexualism (with a homosexual sexual orientation) using the criteria in the DSM-III-R (American Psychiatric Association, 1987). It will be recalled that two CAH females (siblings) had been raised as boys by their parents, who had declined medical treatment for both of them in infancy/early childhood. Thus, of the 53 potential probands (excluding the one who had died in infancy), 3 (6%) were currently living as men.

This percentage was compared to one prevalence estimate of femaleto-male transsexualism in genetic females, 1 in 30,400 (0.0000329%) (Bakker, van Kesteren, Gooren, & Bezemer, 1993). Using this baseline prevalence value, the odds ratio was 1823.70:1 that a genetic female with CAH in our sample was living, as an adult, in the male social role compared to genetic females in the general population living in the male social role (if we exclude the two CAH patients reared as boys from infancy, the odds ratio was 607.9:1).

Our group data appear to be comparable with other reports on the gender identity status of adult females with CAH. All of these studies indicate (or imply) that the vast majority differentiated a female gender identity (e.g., Dittmann, Kappes, & Kappes, 1992; Kuhnle & Bullinger, 1997). One early study of women with CAH is of particular interest. Ehrhardt, Evers, and Money (1968) studied 23 CAH women (M age, 33 years) who were "late-treated," (i.e., they did not receive early corticosteroid replacement therapy and thus had lived for many years with the "stigma of heavy virilization, sometimes uncorrected genital morphology and lack of feminine secondary sexual development" [p. 117]). The mean age of treatment with cortisone was 26 years (range, 8-47 years). Although Ehrhardt et al. did not directly assess the gender identity of these patients, all were living as women and none were judged to be severely gender dysphoric.

In some cohorts of patients, a percentage of genetic females with CAH were assigned to the male sex at birth (invariably due to the extreme masculinization of the external genitalia) and subsequently raised as boys without apparent complications. For example, in one large cohort, Mulaikal, Migeon, and Rock (1987) reported that 9 (6%) of 158 genetic females with CAH were assigned to the male sex and reared as boys. This cohort appeared to include at least some patients born prior to the availability of treatment with corticosteroids (see also Abdullah et al., 1991; Chan-Cua, Freidenberg, & Jones, 1989; Hinman, 1951a, 1951b; Kandemir & Yordam, 1997; Money & Dalery, 1976; Sripathi, Ahmed, Sakati, & Al-Ashwal, 1997; van Seters & Slob, 1988; but also see Hochberg, Gardos, & Benderly, 1987). Given the gradual improvement in the early diagnosis and detection of CAH, it is likely that, in more contemporary cohorts, the percentage of genetic females declared to be males will decrease (Frank, 1997).

Genetic females with CAH reared as boys is interesting in its own right because it tells us that a male gender identity can differentiate in a person who, for example, has female sex chromosomes and internal reproductive structures. It is likely that the masculinization of the external genitalia, which go "uncorrected," work in concert with masculine gender socialization. Moreover, socialization as boys may well augment the putative prenatal androgenization of the CNS that predisposes such youngsters to behavioral masculinity.

More interesting, however, are the cases of gender change from female to male that occur gradually over the life course at the instigation of the person with CAH, not others. Recently, Meyer-Bahlburg et al. (1996) reviewed this aspect of the CAH literature and presented data on four new patients in which this type of transformation occurred (see also Hinman, 1951b, Money, 1991, Chapter 9, Case 2). Meyer-Bahlburg et al. identified four factors that appeared contributory: (a) lack of surgical feminization or delay beyond infancy; (b) poor adherence to glucocorticoid replacement therapy, resulting in progressive physical virilization; (c) markedly masculine childhood gender role behavior; and (d) sexual attraction to females.

Summary

In summary, the data on gender identity differentiation among genetic females with CAH generally support the Money et al. (1957) argument that gender identity differentiates primarily in accordance with gender of rearing. Nonetheless, there appears to be variability in the extent to which females with CAH are satisfied or content with their gender identity, and this variability appears to be greater than what is observed among control females.

To some extent, the assessment of gender identity in CAH females has been less adequately appraised than other aspects of their psychosexual development, including assessment of gender role behavior and sexual orientation, for which better psychometric measures have been utilized (e.g., Berenbaum & Hines, 1992; Meyer-Bahlburg et al., 1999; Zucker et al., 1996). Future studies on both girls and women should attempt to rely on more rigorous assessment techniques, such as the Gender Identity Interview for Children (Zucker et al., 1993) and the DSM-IV criteria for the assessment of GID. Regarding the latter, although Slijper et al. (1998) reported relying on the DSM-IV criteria, it was not clear from their study the precise symptoms that were present, and it would be useful in subsequent studies to report on the actual manner in which the DSM-IV criteria were rated and the percentage of patients who met criteria for each of the symptoms that contribute to the overall diagnosis.

Recently, Diamond and Sigmundson (1997a) have suggested that genetic females with CAH with marked genital masculinization at birth be raised as males, in contrast to those with less marked genital masculinization (e.g., with a hypertrophied clitoris). It is not clear if the extant data provide direct support for this recommendation. On the one hand, it is likely the case that the CAH females who change gender later in life generally come from the subgroup with the most marked genital masculinization at birth. On the other hand, it is unlikely that this variable alone accounts for the gender change (see, e.g., Meyer-Bahlburg et al., 1996). To adopt the Diamond and Sigmundson (1997a) recommendation, one would like to see evidence that the quality of life of markedly masculinized genetic females with CAH who would be raised as males would be in any way superior to that of those reared as females, who receive the corresponding surgical feminization of the genitalia, regulation of the condition with cortisol-replacement therapy, and retain the potential for fertility (see, e.g., Meyer-Bahlburg, 1999c; Zacharin, 1999).

5-Alpha-Reductase Deficiency

In genetic males, this syndrome often, among other things, results in marked ambiguity of the external genitalia at birth. In the absence of a proper work-up, the genital configuration is often judged female (see, e.g., Opitz et al., 1972; Simpson, New, Peterson, & German, 1971). Prior to the discovery of the underlying metabolic defect (see below), the condition was characterized descriptively (Nowakowski & Lenz, 1961) and was subsequently called pseudovaginal perineoscrotal hypospadias (Opitz et al., 1972; Simpson et al., 1971).

During fetal development, an impairment of steroid 5alpha-reductase activity leads to an underproduction in plasma dihydrotestosterone (DHT), which causes the incomplete masculinization of the external genitalia (Russell & Wilson, 1994).7 Because testos-terone production is unaffected, masculinization of the internal reproductive structures is normal.

Moreover, at puberty, there is relatively normal physical masculinization-both primary and secondary sex characteristics develop along male lines (e.g., the phallus enlarges to become, in some instances, a functional penis; the testes descend, if they had not done so already; the voice deepens; and there is an increase in muscle mass). Although affected patients may have normal sperm counts, many are infertile or have azoospermia or oligospermia associated with undescended testes. Fertility is often impaired secondary to a rudimentary prostate and underdeveloped seminal vesicles, which results in highly viscous semen and a low volume of ejaculate (see, e.g., Cantu et al., 1976); however, a recent case report described a successful conception via intrauterine insemination (Katz et al., 1997). For further details on the syndrome itself, including recent understanding of its molecular genetic basis, see Fratianni and ImperatoMcGinley (1994), Imperato-McGinley et al. (1974), Thigpen et al. (1992), Walsh et al. (1974), and Wilson et al. (1993).

Gender Identity Differentiation

In behavioral sexology, 5-ARD began to receive a great deal of attention about 25 years ago. At that time, Imperato-McGinley et al. (1974) described a cohort of affected individuals from the Dominican Republic, in which the prevalence of the condition was unusually high because of in-breeding (Imperato-McGinley, Gautier, Peterson, & Shackleton, 1986), showing a gradual change in gender identity from female to male (i.e., they "switched" from living as girls/women to boys/men).8 ImperatoMcGinley et al. (1974) noted that newborns with 5-ARD were often assigned to the female sex and subsequently "raised as girls." At puberty, the girls' physical masculinization was so striking that they became known locally as "guevedoces"-penis at 12 (years of age). ImperatoMcGinley et al. (1974) commented that after puberty "psychosexual orientation" was male in all of the 18 individuals who had been reared as females (i.e., they perceived themselves as male, adopted a masculine gender role, and were sexually attracted to females). Imperato-McGinley et al. concluded that the

male sex drive appears to be testosterone related and not dihydrotestosterone related ... and the sex of rearing as females ... appears to have a lesser role in the presence of two masculinizing events-testosterone exposure in utero and again at puberty with the development of a male phenotype. (p. 1215)

In a subsequent article, Imperato-McGinley, Peterson, Gautier, and Sturla (1979) provided additional details about their subjects, including a description of the community's social life, the community's behavioral expectations for children, and the gender role tasks of the community's adults. They interviewed affected individuals and significant others (e.g., parents, siblings, neighbors) "to discern any sexual ambiguity in the rearing of subjects raised as girls and to determine in these subjects the validity of the change to a male ... gender identity and male ... gender role" (pp. 1233-1234). Of 18 subjects "unambiguously raised as girls ... 17 of 18 changed to a male ... gender identity and 16 of 18 to a male ... gender role [during or after puberty]" (p. 1233). They summarized their data as follows:

The 17 subjects who changed to a male ... gender identity began to realize that they were different from other girls in the village between 7 and 12 years of age, when they did not develop breasts, when their bodies began to change in a masculine direction and when masses were noted in the inguinal canal or scrotum. These subjects showed self-concern over their true gender. A male ... gender identity gradually evolved over several years as the subjects passed through stages of no longer feeling like girls, to feeling like men and, finally, to the conscious awareness that they were indeed men. (p. 1234)

Since these initial reports, similar accounts of gender change from female to male have been described in cohorts of subjects from Papua New Guinea (Imperato-McGinley, Miller, et al., 1991), Mexico (Mendez et al., 1995; cf. Perez-Palacios et al., 1984), Brazil (Mendonca et al., 1996), and the Middle East (Al-Attia, 1996; Elsayed, Al-Maghraby, Hafeiz, & Taha, 1988; Hochberg et al., 1996; Taha, 1994), as well as in single case or small series reports (e.g., Cantu et al., 1976; Deslypere, Coucke, Robbe, & Vermeulen, 1985; Imperato-McGinley et al., 1980; Kuttenn et al., 1979; Nordenskjold, Magnus, Aagenaes, & Kundtzon, 1998; Price et al., 1984; Savage et al., 1980). Moreover, there are two other syndromes-30 and 170-hydroxysteroid dehydrogenase deficiency-that have an endocrinological natural history very similar to that of 5-ARD, and several reports have described the same type of gender change from female to male (Imperato-McGinley, Peterson, Stoller, & Goodwin, 1979; Mendonca et al., 1987; Rosler & Kohn, 1983; see also Gross et al., 1986).

In some quarters, the reaction to the reports by Imperato-McGinley and colleagues was not to dispute the veracity of the gender change, but the explanation for it (see Zucker & Bradley, 1995, pp. 208-212). Here, I will summarize and provide an update on the empirical status of the literature and of the conceptual issues.

Regarding the Imperato-McGinley cohort, the initial commentaries contained a great deal of discourse about the nature of the gender assignment at birth, particularly among post first-generation probands. For these probands, the family and community had acquired some knowledge about the natural history of the condition, including the physical masculinization that occurred at puberty. Thus, for these probands, it was argued that they were not assigned unambiguously to the female sex and reared as girls, but that the culture created both "third sex" and "third gender" categories (cf. Herdt, 1994) that would guide postnatal socialization. Money (1976), for example, commented on the psychological significance of the "folk prognosis" of guevedoces. Money argued that since the parents rapidly became aware of the impending physical masculinization at puberty that they

could not confidently assign a newborn hermaphroditic baby as a girl. Even if the birth certificate was assigned female, the parents would know they were rearing a guevedoce who would not look feminine after childhood. Even in the first generation in which hermaphrodites appeared in two families in the family tree ... before they could be defined as guevedoces, parents would rear their child as one of ambiguous sex, not knowing what to expect at puberty. (p. 872)

Subsequently, Herdt (1990; Herdt & Davidson, 1988) made similar arguments with regard to the social environment of 5-ARD individuals from Papua New Guinea.

The rebuttal to Money by Imperato-McGinley, Peterson, and Gautier (1976) was, quite interestingly, equivocal. On the one hand, they argued that

Money obviously does not know how the parents raised these children. Our interviews with some of the affected males and their parents indicate that in the first generation, the affected subjects ... were raised as girls and there was no ambiguity on the part of the parents as to the sex of the child at birth or in early childhood. They believed they were raising a little girl. (p. 872)

On the other hand, Imperato-McGinley et al. (1979) also observed that because the condition had become better recognized, the

villagers now either raise the subjects as boys from birth, rear them as boys as soon as the problem is recognized in childhood or raise them ambiguously as girls. Now that the villagers are familiar with the condition, the affected children and adults are sometimes objects of ridicule and are referred to as guevedoce, guevote (penis at 12 years of age) or machihembra (first woman, then man). (p. 1235)

How are we to best interpret the phenomenon of gender change among individuals with 5-ARD? On the one hand, some have read the data as suggestive of a strong main effect (to use the language of the factorial design) of biological influences on gender identity differentiation. On the other hand, it has been argued that the evidence is more supportive of interaction effects: that is, genetic males with 5-ARD have, despite their ambiguous genitalia at birth, male-typical prenatal masculinization, including putative CNS effects, that predisposes to postnatal behavioral masculinity. The response in the social environment augments this biological predisposition, which is augmented even further by the spontaneous physical masculinization that occurs at puberty. Among 5-ARD individuals raised from infancy as boys, the social environment certainly appears to work in concert with the putative prenatal masculinizing CNS effects (e.g., Farkas & Rosler, 1993; Ivarsson, Neilsen, & Lindberg, 1988; see also Forti et al., 1996; Maes, Sultan, Zerhouni, Rothwell, & Migeon, 1979; Ng et al., 1990; Odame, Donaldson, Wallace, Cochran, & Smith, 1992; Opitz et al., 1972, Cases 6-8; Sinnecker et al., 1996, Cases 5-6, 8-9).

How best to resolve these competing explanations? Here, it would be extremely useful to have natural history data on individuals with 5ARD who are treated differently, both medically and psychosocially, than in the cultural groups noted above. For example, in the United States and Europe, it is possible that at least some 5-ARD individuals would be raised as girls on the grounds that the external genitalia will not masculinize sufficiently to permit comfortable (heterosexual) sexual functioning (but see Ivarsson et al., 1988). Thus, in accordance with the optimal gender policy (see Table 2), such individuals would be castrated in infancy, receive surgical feminization of the genitalia, and be placed on feminizing hormones at puberty (see, e.g., Opitz et al., 1972; Simpson et al., 1971; Walsh et al., 1974). In this respect, then, the situation would be comparable to the usual course of surgical, hormonal, and psychosocial events for girls with CAH.

Unfortunately, the psychosexual development of 5-ARD individuals treated in this manner has been described rather poorly. The relevant case reports are scattered widely throughout the biomedical literature and reference to gender identity differentiation is often described rather briefly within the context of more detailed accounts of the syndrome itself. To my knowledge, only one prior review has made an effort, albeit in a cursory manner, to describe such individuals systematically (Wilson et al., 1993, pp. 586, 589; see also Carpenter et al., 1990; McCauley, 1990). Because of the importance of such individuals to the debate about gender identity differentiation, I will summarize these reports in more detail here.

In two early reports, both Walsh et al. (1974) and Saenger et al. (1978) each described two patients with 5-ARD, born in the United States, who were all reared as girls (all were gonadectomized anywhere from infancy to early adolescence); however, no detailed information was provided about their psychosexual differentiation (see also Hodgins, Clayton, & London, 1977; Wieacker, Flecken, & Breckwoldt, 1992). Perhaps the absence of any commentary implies that their gender identity was female.

Donahoe, Crawford, and Hendren (1977, Case 2) described another case, in which the child was first evaluated at 10 days, reassigned to the female sex, and received orchiectomy and surgical feminization of the phallus at around 4 months. At age 6 years, the patient was said to have "no problem with social or psychologic adjustment" as a girl (p. 1049).

Fisher et al. (1978) reported on two sisters, ages 13 and 12 years. The "gender orientation" of the older patient was said to be "definitely female" although she was "very athletic and ... nicknamed 'superwoman' at school" (p. 654). It was not clear from the report if this patient's nickname derived from her sex-typed behavior, physical masculinization, or both. Information about the younger sibling's gender identity was not provided, but there was no apparent evidence of gender dysphoria. Both sisters were treated surgically and placed on feminizing hormones. Money (1979) anecdotally described another patient so treated who "in teenage ... is ... consistently female in gender identity/role" (p. 771). Dewhurst, Chapman, Muram, and Donnison (1983) reported on two sisters, ages 9 and 10 years, respectively, who were reared as girls and who were "well adjusted," presumably with regard to gender identity differentiation in particular and psychosocially in general. Both were gonadectomized at the time of evaluation.

Cantu et al. (1980) reported on an 18-year-old patient from Mexico, reared as a girl, who was referred because of primary amenorrhea, lack of breast development, and postpubertal masculinization. Even in the prior absence of feminizing hormone therapy and bilateral orchiectomy, the patient was reported to have differentiated a female gender identity.

Several other case reports have described patients in adolescence or young adulthood referred for similar reasons. Okon et al. (1980) described two sisters, ages 18 and 17 years, respectively, who were reared as "normal females." At the time of evaluation, both were gonadectomized and the younger sibling's enlarged clitoris was amputated. Presumably, both patients were placed on feminizing hormones, but this was not specifically mentioned in the report. Mauvais-Jarvis, Kuttenn, Mowszowicz, and Wright (1981) reported on two patients, ages 25 and 20 years, respectively. The first patient's "psychosexual orientation" was said to become "definitely male at puberty" (noted above in Kuttenn et al., 1979) whereas the second patient "remained female" (p. 460). Corrall et al. (1984) reported on a 21-year-old patient, reared as a girl, who "possessed an unequivocally female gender identity" (p. 538). Bartsch, Decristoforo, and Schweikert (1987) reported on a 16-year-old patient, reared as a girl, who was reported to have differentiated a female gender identity. Subsequent to feminizing hormonal therapy and surgery, the patient was reported to have married and to have "regular sexual intercourse" (p. 387). Hurtig (1992) described the psychosexual development of two sisters with 5ARD, but with scant detail, both of whom received urogenital surgery and estrogen therapy. One of the two girls made a "satisfactory adjustment to her status as a female," but the other girl "rejected her female identity and has taken on the identity of a male, including sexual attraction and orientation to women" (P. 24). Boudon et al. (1995) reported on four patients from three families. All were said to develop a "female sexual identity," of whom three were not treated medically prior to puberty. Lastly, Sinnecker et al. (1996, Cases 2-4) described three patients who also appeared to differentiate a female gender identity, of whom two had had no medical treatment prior to puberty.

Further evidence for variability in gender identity differentiation among patients with 5-ARD comes from a recent study of 16 patients from 10 families in Brazil (Mendonca et al., 1996). In this series, the age at diagnosis ranged from 4-33 years, the majority (12/16) being diagnosed in adolescence or adulthood. Of the 15 assigned to the female sex at birth, 12 requested a change to the male gender, but three did not (for similar variability, see also Al-Attia, 1996; Elsayed et al., 1988). Two of the three patients living as women did not receive feminizing hormonal and surgical treatments until late adolescence and transpired at the patients' request. Of the three living as women, follow-ups were anywhere between mid-adolescence and adulthood (range, 15-32 years). All of these latter patients also reported a heterosexual sexual orientation and had sexual relationships with men.

From an interpretive standpoint, one could argue that a female gender identity differentiated in these patients because the social environment vis-a-vis gender identity (along with the corresponding surgical feminization of the genitalia, etc.) overrode the prenatal hormonal masculinization and its putative CNS effects. But even this argument may not be strong enough because in the Fisher et al. (1978), Cantu et al. (1980), Okon et al. (1980), Mauvais-Jarvis et al. (1981, Case 2), Corrall et al. (1984), Bartsch et al. (1987), and Mendonca et al. (1996) cases, a female gender identity differentiated in the absence of any hormonal or surgical intervention prior to early adolescence or young adulthood. Thus, even under the circumstance of a masculinizing body habitus at puberty, if not earlier, these patients retained a female gender identity, a finding reminiscent of the female gender identity among late-treated women with CAH (Ehrhardt, Evers, & Money, 1968).

Summary

In summary, the data on gender identity differentiation among patients with 5-ARD are strikingly complex. On the one hand, the phenomenon of gender change first reported on by Imperato-McGinley et al. (1974) has been verified in several additional cohorts of patients, thus constituting an important replication of the original observation (Wilson et al., 1993). On the other hand, differentiation of a female gender identity also occurs in a minority of patients, even in cultures in which there is strong external pressure for the patient to change from living as a female to a male (see, e.g., Al-Attia, 1996; Mendonca et al., 1996; Perlmutter, 1994). To some extent, this latter finding appears to challenge the now prevailing perspective in the literature on the psychosexual natural history of patients with 5-ARD (i.e., that gender change is inevitable subsequent to the masculinizing events that occur around the time of puberty).

There are, however, some important caveats that must be borne in mind. First, the evidence for female gender identity differentiation in some patients with 5-ARD is presented in a rather informal manner, with little information provided about the clinical assessment, and there is a virtual corresponding absence of any formal psychometric assessment tools. Second, little is said about these patients' gender role behavior (which, one might predict, would be shifted in a male-typical direction). Third, the sexual orientation of patients living as women has been poorly described and none of the reports that described this parameter relied on formal interview techniques or psychometrically sound self-report questionnaires.

To account for this variability in gender identity outcome, what are the best candidate explanations? One possibility is that variation in gender identity differentiation is related to the degree of prenatal hormonal masculinization of the CNS and, perhaps, to the degree of postnatal physical masculinization at the time of puberty and beyond (see Sinnecker et al., 1996). A second possibility is that the variation is accounted for by the classical sex of rearing hypothesis originally advanced by Money et al. (1957): For those patients who are unequivocally reared as females, a female gender identity ensues; for those reared ambiguously (along with the corresponding pubertal physical masculinization), a gender change ensues; for those reared as males from infancy onwards, a male gender identity ensues. Unfortunately, as I have argued elsewhere (Zucker & Bradley, 1995), this interpretation has been hampered by the absence of good prospective data and the reliance on post hoc interpretations of the eventual outcome. As I also argued elsewhere (Zucker & Bradley, 1995), what is urgently needed is a psychometrically sound method of assessing the sex-of-rearing construct, in which observers can agree whether the affected individual is being raised unequivocally as a girl or as a boy or, alternatively, somewhere in between these two traditional modes of gender rearing.

Here, I would like to introduce a third interpretation of the data, based on Bem's (1996) recent theoretical analysis of the developmental factors involved in the differentiation of sexual orientation, which considers both between-sex and within-sex variation in sexual orientation development. In the model, Bem argues against a purely biological or "essentialist" account of sexual orientation development (i.e., in accounting for the fact that most men are attracted sexually to women and most women are attracted sexually to men-a between-sex difference-as well as for within-sex variation in sexual attraction). Rather, Bem argues for a biological account that predisposes individuals to show variation in aspects of temperament (e.g., activity level and aggression) which, in turn, is associated with some of the most common sex-dimorphic behaviors that distinguish boys from girls (e.g., playmate preferences, gender role interests, and so on).

Regarding within-sex variation in sex-dimorphic behavior, it is now well-established that, on average, homosexual men and women show more childhood cross-gender behavior than heterosexual men and women (Bailey & Zucker, 1995). Bem argued that for children who grow up in cultures that emphasize the differences in gender role behavior between boys and girls (most cultures? all cultures?), those with "deviant" gender role behaviors experience themselves as feeling different from same-sex peers. Similarly, children with more ordinary gender role behavior experience themselves as feeling different from opposite-sex peers. In Bem's analysis, by middle childhood, heightened nonspecific autonomic arousal occurs in relation to that class of peers from whom one feels different which, in turn, eventually becomes sexualized: "the exotic becomes erotic." Thus, in this model, it is the pattern of gender role behavior that is the mediating variable that accounts for whether a particular child will eroticize same-sex or opposite-sex conspecifics.

Can we extend this model to variations in gender identity differentiation in individuals with 5-ARD (and to other intersex conditions)? I believe that we can in the following manner. From a phenomenologic perspective, it is clear that many 5-ARD patients reared as girls show masculine role interests during childhood. Like the behavioral masculinity of girls with CAH, many authors attribute this gender role preference to the influence of prenatal androgens. Such gender role interests likely contribute to the 5-ARD patient's sense of being different from other "girls," which may be accentuated by any corresponding awareness of genital differences. It is also clear that this sense of being different is accentuated further by the physical masculinization that occurs with the onset of puberty.

The corresponding nascent awareness of sexual feelings towards other "girls" likely further augments the 5-ARD patient's sense of being different. In cultures, or subcultures, in which homoerotic feelings are taboo, then a reasonable solution is to change genders, which would then "normalize" the sexual attraction to same-sex peers: one is heterosexual, not homosexual. My reading of the literature on 5-ARD suggests that there has been an underappreciation of the role of what is commonly referred to in other literatures as internalized homophobia. In other words, there may well be a complex transactional interaction between the 5-ARD patient's sense of difference with regard to both "deviant" gender role behaviors and sexual feelings, and both contribute to the gradual change in gender identity from female (or an ambiguous gender identity) to male. One prediction that follows from this analysis is that, for 5-ARD patients who do not grow up in a culture in which homoerotic feelings are subject to strong sanctions, they will not necessarily feel the "need" to change their gender, but rather will adopt a homoerotic sexual orientation. In this regard, it would be important to better ascertain the sexual orientation of 5-ARD patients who appear to have differentiated a female gender identity. Unfortunately, these data are sorely lacking in the case reports described earlier, and we do not have a clear sense if the proportion of 5-ARD patients living as women are disproportionately homoerotic in their sexual orientation. If they are, it would suggest that the relatively normal prenatal masculinization in such patients has a stronger influence on sexual orientation development, perhaps mediated by the mechanisms proposed by Bem (1996), than it does on gender identity development.

Partial Androgen Insensitivity Syndrome

According to Perez-Palacios et al. (1987), partial androgen insensitivity syndrome (pAIS) is apparently the least common of the androgen resistance syndromes. As the name implies, there is only partial resistance to androgens at the cellular (or postcellular) level; hence, there is a partial masculinization of the external genitalia (see, e.g., Assael, Lancet, & Shani, 1976; Mark et al., 1983). Because there is variation in the degree of androgen resistance, the appearance of the genitalia varies and, therefore, so does gender assignment. Even in so-called familial pAIS (Reifenstein's syndrome), the severity of the condition varies, so some newborns are assigned as boys, others as girls (Money & Ogunro, 1974). Because of familiality in pAIS, the condition is sometimes anticipated in subsequent births and decisions about gender assignment may thus be influenced by the course of events in affected older siblings (see, e.g., Beheshti, Hardy, Churchill, & Daneman, 1983).

Like cAlS, pAIS shows an X-linked pattern of inheritance, although it has been noted that the two conditions never occur in the same pedigree (Morris & Mahesh, 1963; Perez-Palacios et al., 1987), which is one reason that the two conditions have been considered to be distinct entities. Other studies have reported that cAIS and pAIS have distinct underlying molecular abnormalities, which likely account for the variation in androgen action at the cellular level (see, e.g., Brown et al., 1993; Medina, Chavez, & Perez-Palacios, 1981; Quigley et al., 1995) although Warne and Zajac (1998) have argued that this is not always the case. Regarding differential diagnosis, it has been further established that pAIS is, in fact, distinct from 5-ARD. For example, in patients with pAIS, administration of human chorionic gonadotropin does not induce phallic growth despite normal testosterone synthesis, normal 5-alpha reductase activity in the genital skin fibroblasts, and normal DHT levels in blood (e.g., Mark et al., 1983).

Gender Identity Differentiation

What do we know about the gender identity development of patients with pAIS? Morris and Mahesh (1963) reported on one patient, age 19 years, reared as a girl, and noted to have a "female ... psychological orientation" (p. 732; presumably, meaning a female gender identity and a heterosexual sexual orientation), but no details were provided. Teter and Boczkowski (1966) reported on another patient, age 21 years, with a hypertrophied clitoris who was reared as a girl. Gender identity was female and sexual orientation was apparently heterosexual. Madden, Walsh, MacDonald, and Wilson (1975) reported on a third patient, age 26 years, also reared as a girl. Details about her gender identity development were not provided and no information was given about her sexual orientation. Assael et al. (1976) reported that their patient was reared as a girl and differentiated a female gender identity; in adulthood, however, the patient became psychotic, which apparently remitted following orchiectomy and after the physical signs of marked virilization (e.g., facial hair) were reduced. Lastly, Quattrin, Aronica, and Mazur (1990) reported on a 21-year follow-up of one patient, initially assigned male but changed to female at age 13 days, with subsequent surgical feminization and gonadectomy at 4 months. The patient and parents had fairly regular counselling throughout childhood and adolescence, Although the patient apparently displayed elements of girlhood tomboyism (p. 706), a female gender identity differentiated. No information was provided about the patient's sexual orientation.

In contrast to these five case reports, Gooren and Cohen-Kettenis (1991) reported on a patient assigned at birth to the male sex because of an enlarged clitoris, but reassigned five days later and "reared as a girl." At age 30, their patient requested sex reassignment following a long history of masculine gender role interests and behavior and a "heterosexual" sexual orientation (i.e., an attraction to biological females).

To my knowledge, there are only two group studies of patients with pAIS. Money and Ogunro (1974) reported on a series of 10 patientseight were reared as boys, one as a "hermaphroditic girl," and one as a girl. In this series, the medical diagnosis was not made until adolescence or adulthood. At the time of last follow-up, the median age was 24 years (range, 13-39 years). Based on a variety of descriptive and qualitative data, gender identity was judged to differentiate in accordance with the gender of rearing.

Slijper et al. (1998) assessed eight patients with pAIS, with a mean age of 11.5 years (range, 6-23 years). In contrast to the sex assignment pattern in Money and Ogunro's (1974) series, seven were assigned to the female sex at birth. This appears largely a result of the Slijper et al. clinic team following the optimal gender policy described earlier (Table 2). Of the seven patients assigned to the female sex, three were reported to display "deviant" gender role behavior (cf. Crawford, 1970), at least by their parents, including one who met criteria for GID. Details on the assessment of the girls' gender role behavior was not provided; however, Slijper et al. commented that the girls' "boyish conduct [cross-gender behavior] was perceived as an indication that the decision to assign the female sex had been wrong. In particular, the wild, rough play of these [girls] was difficult for their parents to regulate" (p. 137).

Summary

Compared to the psychosexual literature on genetic females with CAH, the psychosexual data on pAIS are surprisingly sparse. The case report literature is patchy at best, with minimal information provided about the patients' psychosexual differentiation. The two group studies had very different distribution patterns regarding gender assignment, which was likely due to cohort effects in terms of the policy of psychosocial and medical management (Money & Ogunro, 1974; Slijper et al., 1998).

For patients reared as girls, one can glean from the available cases evidence for a common pattern of girlhood tomboyism, which appears comparable to the data on genetic females with CAR Unfortunately, the behavioral data are much poorer in quality for patients with pAIS than with CAH. Nonetheless, in both syndromes, prenatal androgenization may well be the active biological mechanism that accounts for the behavioral masculinity (Collaer & Hines, 1995). In CAH, prenatal androgenization is clearly elevated compared to unaffected females. In pAIS, it is not clear if prenatal androgenization would be comparable to unaffected males, but it is, no doubt, greater than in biologically normal females.

The data on gender identity suggest a mixed pattern, with some patients differentiating an uncomplicated female gender identity, but others developing a conflicted gender identity, including a change to the male gender later in life. If the Slijper et al. (1998) data are valid, the presence of a formal GID in one of seven patients in childhood is obviously well above the likely prevalence of GID in the general population.

In my own clinical experience, I have assessed three youngsters with the diagnosis of pAIS, all of whom were assigned to the female sex in infancy-in two cases, the assignment was made after a brief trial of exogenous testosterone treatment, which did not result in adequate phallic growth. These youngsters likely represent a biased sample because they were referred, in part, because of concerns about their gender identity development. Moreover, in two of the three cases, the parents remained uncertain about the "correctness" of the gender assignment.

On standardized assessment measures, these youngsters generally showed evidence for masculine gender role preferences, illustrated in Table 3 with results from a sex-typed behavioral free play task. These youngsters also expressed some confusion about their gender identity. For example, one 8-year-old girl (Verbal IQ, 100; Performance IQ, 126) provided the following responses on a structured gender identity interview schedule (Zucker et al., 1993):

Interviewer (I): Are you a boy or a girl?

Child (C): Girl.

I: Are you a boy?

C: No.

I: When you grow up, will you be a mommy or a daddy?

C: Mommy.

I: Could you ever grow up to be a daddy?

C: No.

I: Are there any good things about being a girl?

C: Yes.

I: Tell me some of the good things about being a girl.

C: Don't be shy, wear dress, cut your hair long.

I: Are there any things that you don't like about being a girl?

C: Yes.

I: Tell me some of the things that you don't like about being a girl.

C: I can't play with Nintendo. I can't play with boys.

I: Do you think it is better to be a boy or a girl?

C: Boy.

I: Why?

C: Because you get to play with Nintendo, get to play with Sega and get to play with guns.

I: In your mind, do you ever think that you would like to be a boy?

C: Yes.

I: Can you tell me why?

C: Because you could play lots more things than girls, you can buy whatever you want, when you get married you don't need to do anything.

I: In your mind, do you ever get mixed up and you're not really sure if you are a boy or a girl?

C: Sometimes.

I: Tell me more about that.

C: No thanks.

I: Do you ever feel more like a boy than like a girl?

C: Yes ... a little bit more and ... because I could drink Coke [CocaCola(TM)] and more, I could eat lots of chips.

I: You know what dreams are, right? Well, when you dream at night, are you ever in the dream?

C: Yes.

I: In your dreams, are you a boy, a girl, or sometimes a boy and sometimes a girl?

C: A boy... Well, I dream about the Power Rangers and the Mystery Knights. I dream about when I was in the Power Rangers, I was Silver Ranger, I could beat everyone up.

I: Do you ever think that you really are a boy?

C: A little bit, not too much.

I: Tell me more about that.

C: No thanks.

Thus, on this interview, this youngster could correctly identify herself as a girl, not a boy, but showed evidence of some desire to be a boy, which, at least on the surface, was closely intertwined with her masculine gender role preferences and the belief that boys could do more things. There was also evidence, both from the interview schedule and additional information, that being a boy was associated with more power. Unfortunately, her parents' continued ambivalence and resistance to understanding how their daughter felt about herself precluded further clinical evaluation.

The Money, Devore, and Norman (1986) Study

A study by Money et al. (1986), which likely included cases of 5-ARD and pAIS, needs to be described separately because the sample was initially generated "prior to the era of diagnostically subtyping male hermaphroditism" (pp. 179-180) on the basis of specific endocrine profiles or other parameters. Money et al. studied 32 genetic male patients assigned to the female sex in infancy and then followed-up at a mean age of no younger than 18 years (range not specified). Of these, 26 patients were deemed "regular" referrals (i.e., they were seen through the Johns Hopkins pediatric endocrine clinic for routine long-term care and follow-up). The remaining six patients were deemed "special" referrals, as they came from other clinics or hospitals because the patient was requesting a sex reassignment (age of request not specified).

Money et al. (1986) examined several variables as potential correlates of the request for sex reassignment and/or a homosexual/bisexual sexual orientation. Across both types of referrals, nine patients were requesting sex reassignment and had a homosexual or bisexual sexual orientation (in relation to the patient's genetic sex) and six others were homosexual or bisexual but were not gender dysphoric. These two types of "gender transpositions" were associated with three variables: presence of a childhood history of stigmatization (both at home and in the community) pertaining to the physical intersex condition and age at which feminizing surgery and gonadectomy occurred. In the gender transposed group, feminizing surgery of the genitalia occurred at 3 years or later. The presence of gender transpositions was not associated with other variables, such as initial uncertainty about the infant's sex and the presence of gross family behavioral pathology.

In interpreting the pattern of findings, Money et al. (1986) considered in particular the delay of surgical feminization, both with regard to its potential impact on gender self-representation and on the reaction of significant others, such as parents and peers (cf. Meyer-Bahlburg et al., 1996): Stigmatization is more closely related to postponement of surgical feminization ... so that the child looks genitally abnormal. She becomes subject not only to stigmatization by others, but also to self-stigmatization. In a world dichotomized by sex, if she does not look completely like a girl, then the only other alternative is that she must look like a boy, or a halfboy, half-girl. Self-stigmatization leads the way to a self-generated development of a gender transposition, masculinizing all aspects of the girl's gender-identity/role. (pp. 178-179)

Micropenis Syndromes

According to Feldman and Smith (1975), the mean stretched length of a newborn male's penis is 3.5 cm. (SD = .35). The diameter is 1. 1 cm (SD = . 10). How does one decide that a penis should be called a micropenis? Money, Mazur, Abrams, and Norman (1981) used the criterion of a stretched length as no longer than 2 cm (i.e., 4 SD below the mean, which corresponds to a percentile value at or below the 3rd percentile).

As noted by Smith (1977) and Lee et al. (1980), a micropenis (or, at least an underdeveloped one) occurs in several known syndromes (e.g., pAIS), but it can also be idiopathic. Apart from etiological considerations, the birth of a newborn infant with a tiny penis, particularly when it does not enlarge in response to exogenous testosterone treatment, results in debate about whether the infant should be raised as a boy or a girl (Van Wyk & Calikoglu, 1999).

In this section, I will review data on gender identity differentiation among XY patients with micropenis that is not associated with other forms of intersexuality, such as pAIS or 5-ARD. The extant literature indicates marked inconsistency in psychosexual management: Some of these babies have been raised as boys, others as girls. For those reared as girls, the management plan includes surgical feminization of the genitalia and administration of feminizing hormones at puberty.

Gender Identity in Patients Reared as Boys

Hinman (1972) reported on 20 cases, noting that "in all cases directly under the author's control the decision was for repair rather than sex conversion" (p. 501); however, details on psychosexual development, including gender identity differentiation, were not provided.

Money and Mazur (1977) reported on one patient, who was referred at age 9 months for micropenis (1.5 cm) and undescended testicles. The patient was then lost to follow-up until the age of 23 months, at which time the stretched penis remained at 1.5 cm. During this period, the child was adopted, and there had been some responsiveness to exogenous testosterone cream. It appeared that the infant had been raised as a boy, as the adoptive parents "were habituated to treating the child as a boy, and his gender identity/role was already well advanced in its differentiation, as illustrated in the application of gender dimorphic nouns and pronouns in self-reference" (p. 192). Continued exposure to testosterone application resulted, after 2 months, in a stretched penis length of 2.4 cm, but it did not enlarge thereafter. At age 9, the boy was receptive to wearing a prosthetic phallus made of nontoxic, flexible plastic, which was judged to enhance the youngster's "self-esteem and self-confidence as a male" (p. 193).

Money (1984b) provided a particularly detailed case report through young adulthood on another patient reared as a boy. In childhood, the patient tended to affiliate with girls and avoided competitive sports although, in other respects, there were no gross signs of behavioral femininity. This patient struggled for years with feelings of inferiority related to his small penis, which was associated for a considerable period of time, beginning in late childhood, with fantasies of sex-reassignment and then later on with severe depression and suicidal urges. As this patient lucidly described, "If you ever let parents who have a baby with a microphallus raise their kid as a male, you're a damn fool" (p. 363). After a period of heterosexual behavior in late adolescence, the patient reported stronger homoerotic attractions and, by young adulthood, had differentiated a homosexual sexual identity and greater sexual satisfaction in having sex with other men than with women.

Money, Lehne, and Pierre-Jerome (1985) reported adult follow-up (M age, 25 years; range, 22-31) of nine youngsters with micropenis (of diverse etiology) raised as boys. Money et al. noted that topical application of testosterone propionate resulted in a "partial and localized pubertal-type of growth spurt," but, by adolescence and adulthood, "the penis permanently resume[d] a disproportionately small dimension relative to the rest of the body" (p. 29). Thus, boys "with a micropenis [grow] up to be confronted with the teenaged and young-adult challenges of coping with a micropenis erotosexually and with respect to overall behavioral health" (P. 29).

Five patients appeared to differentiate an uncomplicated male gender identity, with a corresponding childhood history of male-typical behavioral masculinity, and all had a heterosexual sexual orientation. The remaining four patients, including the one described above (Money, 1984b), were less conventionally masculine and/or feminine during childhood, but appeared to have a male gender identity. Three of these four patients differentiated a homosexual sexual orientation; the one with a heterosexual sexual orientation had a co-occurring sadomasochistic paraphilia.

Reilly and Woodhouse (1989) provided follow-up data on 20 patients with micropenis of diverse etiology, including 6 who also had hypospadias: 8 were still prepubertal (2 of whom were severely mentally retarded) and 12 were between 17 and 43 years old. An additional 30 patients were unavailable for follow-up.

Based on Schonfeld's (1943) age-graded norms, the penile length of the prepubertal group at follow-up was at or below the 10th percentile (range, 2-5 cm), except for one with a stretched length of 7.2 cm. It was reported that "all parents in this group considered their children as normal boys .... but they expressed concern about [the] size of the penis, wondering if sexual function in adulthood would be a problem" (p. 569). One additional patient, lost to follow-up, was reported to have a lot of psychosocial difficulties secondary to the micropenis and nonpalpable testes.

Of the 12 postpubertal patients, stretched penile length was below the 10th percentile (range, 4-10 cm) although "the 3 largest penises look normal for body size and only measurement revealed the deficiency" (Reilly & Woodhouse, 1989, p. 571). Half of these patients recalled childhood experiences of being teased; the other half did not. All patients "felt male" although one was reported to lack "total confidence about it" (p. 571). Nine of the patients had had sexual intercourse and seven were married or co-habitating. It was noted that "the group was characterized by an experimental attitude to positions and methods" (p. 571) in order to accommodate the constraint imposed by the size of the penis.

Reilly and Woodhouse (1989) argued that parental attitudes were associated with how well this subgroup adapted to having grown up with a small penis although no formal statistics were used to demonstrate this putative relation (for a similar study, see Bin-Abbas, Conte, Grumbach, & Kaplan, 1999).

Gender Identity in Patients Reared as Girls

Money et al. (1981) provided descriptive data on the initial decisionmaking process in assigning 14 newborns with micropenis to the female sex. Follow-up data (range, 10-25 years) on 3 patients showed that a female gender identity had differentiated without known complications. Subsequently, Money (1984a) provided a particularly detailed case report through young adulthood on another patient reared as a girl. Gender identity was female and sexual orientation was heterosexual.

Money and Norman (1988) reported on four patients with micropenis reared as girls (age at last follow-up ranged from 10-18 years). In this series, there was a co-occurring history of either learning disability or the CHARGE syndrome (coloboma, heart disease, atresia choanae, retarded growth, genital hypoplasia, and ear anomalies). Eight other patients reared as girls were lost to follow-up. All four patients appeared to differentiate a female gender identity, although their general psychosocial adaptation was complicated by their co-occurring developmental problems.

Summary

The data on gender identity differentiation among patients with micropenis suggest a great deal of variability and, if these reports are accurate, largely a function of the gender of rearing. There are, however, several limitations to these reports. First, it should be recognized that the number of reports is relatively few in number if one compares them to some of the other syndromes, such as CAH and 5-ARD (Meyer-Bahlburg, 1999a). Second, cases of "pure" micropenis are relatively rare, as the condition often occurs in concert with other signs of physical intersexuality, rather than in isolation. Third, the rigor of assessment is rather poor in these studies, with reports of outcome largely confined to clinical description, rather than reliance on standardized psychometric assessment instruments or structured interview schedules. Lastly, several of these reports had a very high selection bias, with many patients unavailable or lost to follow-up (e.g., 60% in Reilly & Woodhouse [19891 and 67% in Money & Norman [1988]). Thus, generalization about outcome must be made with great caution.

Penile Agenesis

Penile agenesis (aphallia) is an intersex condition in which the penis fails to differentiate, although the scrotum is normally formed and contains the testicles (Kessler & McLaughlin, 1973; Richart & Benirschke, 1960). Prenatal androgenization, including putative CNS effects, is apparently normal. The condition is quite rare although conjectures about prevalence have varied wildly, from 1 in 50,000 (Young, Cockett, Stoller, Ashley, & Goodwin, 1971) to 1 in 10-30 million live births (see Kessler & McLaughlin, 1973)! Mortality is high because of associated urinary and gastrointestinal tract abnormalities (see, e.g., Farah & Reno, 1972; Kirshbaum, 1950). Accordingly, it is not surprising that there are very few studies of affected patients with regard to psychosexual differentiation.

Gender Identity Differentiation

The literature suggests that, when reared as a boy, a male gender identity differentiates, albeit at times with complications. Stoller (1965), for example, reported on two cases in some detail. At age 4, the first patient was judged to have a male gender identity, with typically masculine gender role interests. The patient's mother described his behavior as follows:

He likes to wrestle and box. He likes all kinds of sports ... and he told me that he wants to be a wrestler ... when he is big. He plays with dolls, but when he does, he is the father and his sister is the mother.... You can give him a little stick and send him out to play, and he can make everything out of that stick you can imagine.... [Other children] know he has a catheter on. They have seen it and they accepted it and treat him like he was a boy... He dislikes anything that looks girlish to him ... he wants everything boy's.... Sometimes he is Superman. (pp. 209-210)

The boy's step-father noted that his favorite game was "Gas Station" (mimicking the father's occupation as a manager of a gas station), using a "cat's tail as a gasoline pump." The youngster was also quite attached to his indwelling catheter, which he often displayed proudly to others.

In the second case, first seen at age 15 years, the patient was also reared as a boy, but in a family environment in which his parents were gauged to be distant and indifferent. Since the age of 1.5 years, he had numerous surgical procedures, including a phalloplasty described as a "monstrosity of unearthly appearance." Although the patient had differentiated a male gender identity, he was severely paranoid, particularly when under stress, describing himself as "the grandson of God and maybe I am the Messiah." Since age 7, he was reported to have engaged in sadistic homosexual activity, including a game called "the Pull," in which each of the two partners pulled forward on the other's penis in order to produce pain (the patient himself, however, did not experience pain and was known by his friends to be faking it). The patient was also reported to be preoccupied with knives.

Drury and Schwarzell (1935) reported on a 13-year-old boy who appeared to have a male-typical gender identity. Kessler and McLaughlin (1973) reported on two cases also reared as boys. By late childhood, the first patient was judged to have "no problem with gender identity" (p. 228). In the second case, the patient at age 22 years was judged to have a male gender identity, but with severe psychiatric problems and was reported to be homosexual. Lisa et al. (1973) reported on an 11-yearold boy, whose "role as that of a boy was clearly defined" (p. 328). Gautier, Salient, Pena, Imperato-McGinley, and Peterson (1981) reported on two cases, but only one was old enough to assess gender identity which, at age 6 years, was judged to be male. Oesch, Pinter, and Ransley (1987) reported on six cases, one of whom was reared as a boy. At age 9 years, there was no mention of gender identity problems, but the result of numerous surgical attempts at phalloplasty was described as "not very satisfactory cosmetically and functionally" (p. 172). In a remarkable case, Rosenblum and Turner (1973) reported on a 45-year-old male who grew up on a farm in rural South Carolina. He was seen medically for the first time at this age complaining of dysuria. Upon physical examination, the patient was noted to lack a penis. The patient was married and appeared to be exclusively heterosexual.

The optimal gender management policy has been prescribed in other cases, in which a female gender reassignment was made. Pohlandt, Kuhn, Teller, and Thoma (1974) reported on one patient, first seen at 2.5 years, at which time the sex assignment was changed from male to female. Follow-up at age 5.5 years indicated that "the sex identification was female" (P. 2166). Skoog and Belman (1989) reported on another patient reared as a girl from the age of 14 days. At age 8, the patient was described as "an apparently well adjusted ... girl" (p. 589). Two other patients were too young to report on their psychosexual differentiation. Young et al. (1971) reported on two cases in which the decision was made to raise the infants as girls (ages not specified), but follow-up data were not reported. Oesch et al. (1987) reported on four patients reared as girls. Cases 3 and 6 appeared to be too young to report information on gender identity development; Case 2 at age 9 years was apparently normal (nothing in the report indicated otherwise), whereas Case 4 was described as having "severe behavior problems" (p. 172) at the time of puberty, but details were not provided. Stolar et al. (1987, Case 2) reported on a 13year-old patient, with an apparent female gender identity. Evans, Erdile, Greenberg, and Chudley (1999) reported on four patients, only one of whom survived. At age 8 years, there was no mention of any problems in the patient's gender identity development although the youngster was noted to be of "borderline low-normal intelligence" and had been recently diagnosed with attention deficit disorder. Several other case reports noted that there was a female gender assignment, but there was no information on follow-up (e.g., Azpiroz, 1971; Bruch, Meuli, & Harrison, 1996; Gluer, Fuchs, & Mildenberger, 1998; Johnston, Yeatman, & Weigel, 1977; Stolar et al., 1987, Case 1; Talwar & Kapoor, 1988).

Dittmann (1998) provided a particularly detailed case report in which the patient was given a female name at the age of 2 months. At age 4 months, however, the patient was given a boy's name because of his genetic sex and, at age 16 months, was legally registered as a male. At age 3, it was suggested by referring physicians that the patient might better be raised as a girl, and the patient was admitted to hospital for further evaluation. During this evaluation, a gender re-assignment was recommended, which occurred shortly before the patient turned 4, at which time gonadectomy was also performed. Prior to age 7, the patient was judged to be "content with her female role" (p. 259). Shortly thereafter, however, the patient was noted to have "serious doubts" (p. 260) about her female gender identity. Yet, as late as age 15 years, the patient was said to be "convinced ... that the decision to rear her as a girl had been correct. . . [and] she ... does not wish to be a boy" (p. 260). By late adolescence, however, the situation changed, and the patient began to live as a man, which maintained itself through the last follow-up at 27 years. The patient was attracted sexually to females and married a woman at the age of 24 years.

Summary

It is obvious that the literature on penile agenesis and psychosexual differentiation is quite patchy, both with regard to standardized assessment during childhood and long-term follow-up. The extant data suggest the following provisional conclusions: (a) Like the other intersex syndromes reviewed so far in which there are either male-typical levels of prenatal androgen exposure (e.g., as in 5-ARD) or levels of prenatal androgen exposure that, at minimum, are greater than in normal females (e.g., as in CAH in genetic females), childhood gender role behavior is shifted in the direction usually observed in unaffected males; (b) A male gender identity can differentiate in the absence of a penis, particularly when the parents appear to have no uncertainty that the decision to raise the child as a boy was the correct one (e.g., Stoller, 1965, Case 1). Regarding this observation, the conclusion is similar to what was reached in cases of micropenis in which the gender assignment was that of a boy; (c) The long-term implications of growing up without a penis remain unclear. It would not be surprising that selfesteem issues related to one's sexual competence during adolescence and adulthood will be variable, as occurs in other conditions, such as hypospadias (e.g., Mureau, Slijper, Nijman, et al., 1995; Mureau, Slijper, Slob, & Verhulst, 1995; Mureau, Slijper, Slob, Verhulst, & Nijman, 1996). In this regard, assessment of general psychological functioning and social support networks will likely be of great importance in accounting, at least in part, for variations in sexual adjustment.

Unlike the situation with 5-ARD and in the micropenis syndromes, in which there are some long-term follow-up data regarding "successful" gender identity differentiation as a female, none of the case reports of patients with penile agenesis raised as girls have provided information on gender identity status in adulthood, with the exception of Dittmann's (1998) case report, in which there was a patient-initiated change from female to male.

Cloacal Exstrophy

Cloacal exstrophy, which occurs in both genetic males and females, is a complex disorder of embryogenesis involving the genitourinary and intestinal tracts (Howell, Caldamone, Snyder, Ziegler, & Duckett, 1983). It includes exstrophic bowel of the ileocecal region, imperforate anus, two exstrophic hemibladders, omphalocele, vertebral anomalies, and anomalies of the external genitalia. Myelomeningocele and scoliosis are common associated anomalies. Prior to 1960, surgical interventions were usually not attempted and the affected newborn was left to die (Rickham, 1960). Over the past 4 decades, advances in pediatric surgery have resulted in a more optimistic approach to management and survival rates now exceed 80% (Howell et al., 1983).

In genetic males with cloacal exstrophy, the configuration of the external genitalia is often poorly differentiated or grossly anomalous (e.g., the penis is split and the two halves are widely separated), if not entirely absent. In bladder exstrophy, the external genitalia are less severely affected, and penile reconstruction is the usual course of medical management, although not always (see, e.g., Mesrobian, Kelalis, & Kramer, 1986; Reiner, Gearhart, & Jeffs, 1999). Although such youngsters appear to differentiate a male gender identity without notable complications, there is some evidence, not particularly surprising, that during adolescence and adulthood, such individuals experience a great deal of uncertainty about their sexual functioning and in initiating intimate sexual relationships (e.g., Reiner et al., 1999).

When the reconstruction of the genitalia is impossible, particularly in the case of cloacal exstrophy, the most common management policy has been to recommend a female sex assignment (Tank & Lindenauer, 1970; Zwiren & Patterson, 1965). Thus, from a psychosexual perspective, this strategy is of interest because, like in penile agenesis, prenatal androgenization, including putative CNS effects, is apparently normal.

Gender Identity Differentiation

What do we know about the gender identity development of genetic males with cloacal exstrophy or bladder exstrophy who have been raised as girls? Unfortunately, not very much; for example, several reports in the literature make mention of sex and gender reassignment, but no information on psychosexual outcome was provided (e.g., Howell et al., 1983).

Hayden, Chapman, and Stevenson (1973, Case 1) reported on a genetic male sex-reassigned right after birth, with removal of the left gonad at 7 months. At around age 3, the right gonad descended, which was removed just prior to age 5 years, a time when the youngster was physically well enough to begin kindergarten. At age 5, Hayden et al. commented on the parents' apparent ambivalence in adapting to the gender assignment:

The psychosocial problem causing the most difficulty at the present time is the parents' adjustment to the arbitrary sex assignment for their child. When the mother called to describe the descent of the testis, her confusion was evident. The multiple determinants of sex, including chromosomal, gonadal, hormonal, and psychological had been discussed with the parents, and they seemed to intellectually understand the plan to override the cellular and gonadal maleness of this child by rearing her as a female with plastic surgery and hormonal intervention as she grew older. That emotionally they have not been able to do this is evident from an increasing tendency to dress the child in boyish clothes and to call her by a masculine nickname. At the time of the removal of the remaining gonad this whole problem was reviewed with the parents, and hopefully with continued counseling they will be able to adjust to a pattern of rearing this child as a girl. Retrospectively, it may have been an error to inform these parents about the chromosomal and gonadal findings. The difficulties that have arisen in this area are in part iatrogenic and point to the hazard that the multiple disciplinary approach may compound confusion unless one member of the team is responsible for relaying and interpreting the mass of information gained. (p. 883)

Although Hayden et al. provided a nice illustration of parental uncertainty, it would have been helpful to know even more, including whether or not the child's gender role behavior itself was contributing to parental uncertainty. On this point, no information was provided.

In another report, Stein et al. (1994) noted that one male (out of a total of 31) with bladder exstrophy, with multiple other malformations and a "lower IQ," had been castrated at age 4 years and "raised as a girl." After puberty, "the patient decided to change his sexual identity and name legally but no further operation was performed" (p. 1414). Feitz, Van Grunsven, Froeling, and de Vries (1994) reported on another male (out of a total of 11) with bladder exstrophy who

was kept home since birth and ... raised as a girl. At the age of 52 years he presented for treatment after the death of both parents. After extensive psychological examinations he underwent urinary diversion ... and reconstruction as a man. He presently demonstrates male behavior. (p. 1418)

In both reports, no further details were provided about the life histories of the two patients, including information relevant to their psychosexual rearing.

Reiner (1997c) reported preliminary impressions on a cohort of 15 genetic males castrated at birth "who are reared unequivocally as females" (p. 225; see also Reiner, 1997a, 1997b). Prior to the age of 12 years, two patients "declared themselves to be males" and three others "spontaneously described themselves as the most masculine girl they know" (1997c, p. 225). No other details were provided.

In another report, Montagnino et al. (1998) reported on the general adjustment of 29 exstrophy patients (M age, 7.8 years; range, 3-18), including 10 genetic males, 4 of whom were raised as girls. Although these four patients did not differ in their general adjustment from the other patients, no specific information was provided regarding their psychosexual adaptation. Perhaps it is reasonable to conclude that if any of these four patients were severely gender dysphoric, it would have been commented on. To my knowledge, there are two additional series of cloacal exstrophy patients, in which a subgroup of genetic males were sex-reassigned but, in both cases, detailed information about psychosexual differentiation is lacking (Kodman-Jones & Zderic, 1999; Mitchell, 1999) although Mitchell alluded to the fact that some of the patients reared as females were unhappy living as girls/women.

Lastly, Meyer-Bahlburg, Ehrhardt, Pinel, and Gruen (1989) reported on two genetic males (ages 8 and 12 years) with cloacal exstrophy who were sex-reassigned shortly after birth. Based on standardized psychometric tests, both patients were reported to have "marked shifts in gender-role behavior" (i.e., in the masculine direction), but "current gender identity appeared female."

I have also assessed one 12-year-old genetic male with cloacal exstrophy (IQ = 100), who was sex-reassigned in the newborn period. Because of co-occurring physical anomalies, the patient was confined to a wheelchair. Her various physical problems had necessitated about 18 different surgical procedures during her growing-up years. Although my patient knew a great deal about her various physical problems, it should be noted that she did not know that she was a genetic male and that a decision had to be made about her gender assignment in infancy. I was asked to see this youngster for a couple of reasons. First, she had recently had a dream in which she was a boy, and her mother wondered if this had any prognostic implications. Second, this youngster was ambivalent about starting feminizing hormone therapy to induce puberty. Clinically, it was unclear if this reluctance was related specifically to gender identity issues or if it was related to more general anxieties about growing up.

Like the patients provisionally described by Meyer-Bahlburg et al. (1989) and Reiner (1997c), my patient had a childhood history of girlhood masculinity. Her current phenotypic appearance (e.g., very short hair and a gender neutral clothing style) often resulted in naive observers perceiving her as a boy, in a manner very similar to that which occurs in girls with GID without somatic intersexuality (see Fridell, Zucker, Bradley, & Maing, 1996; McDermid, Zucker, Bradley, & Maing, 1998). In her local community, however, she was known to her peer group as a girl, and she never made explicit efforts to pass as a boy or to intentionally deceive anyone.

Like Meyer-Bahlburg et al.'s (1989) two patients, my patient did not, at age 12, show any signs of gross gender dysphoria, at least as could be inferred from clinical interview data and a structured interview schedule (Zucker et al., 1993):

Interviewer (I): Are you a boy or a girl?

Child (C): Girl.

I: Are you a boy?

C: No.

I: When you grow up, will you be a mom or a dad?

C: Mom.

I: Could you ever grow up to be a dad?

C: No.

I: Are there any good things about being a girl?

C: Yes.

I: Tell me some of the good things about being a girl.

C: Everything, I don't know, stuff ... I don't know.

I: Are there any things that you don't like about being a girl?

C: No.

I: Do you think it is better to be a boy or a girl?

C: I don't really care .... Because it's basically the same thing.

I: In your mind, do you ever think that you would like to be a boy?

C: No.

I: In your mind, do you ever get mixed up and you're not really sure if you are a boy or a girl?

C: No.

I: Do you ever feel more like a boy than like a girl?

C: No.

I: You know what dreams are, right? Well, when you dream at night, are you ever in the dream?

C: Yes.

L: In dreams, are you a boy, a girl, or sometimes a boy and sometime a girl?

C: Usually a boy.

I: Tell me about the dreams in which you're a boy.

C: I forget .... don't remember any of my dreams.

I: Do you ever think that you really are a boy?

C: No.

Summary

In many respects, the extant data on psychosexual differentiation for cloacal exstrophy patients suggest a pattern very similar to those with penile agenesis, so I will not repeat these observations. I will only emphasize here that it is obvious that more rigorous psychosexual assessments of such patients are required, including long-term follow-up.

Ablatio Penis

There is a final group of genetic male patients that has substantial relevance to the debate about gender identity differentiation, even though in these cases there is no presence of a physical intersex condition. These are patients who have experienced accidental or traumatic loss of the penis, as in cases of circumcision mishaps, amputation by a mentally ill parent, animal attacks, and motor vehicular accidents (e.g., Gearhart & Rock, 1989; Money, 1998; Ochoa, 1998). In one other rare situation involving conjoined male twins who share one set of genitalia, surgical separation results in penile and testicular loss for one of the twins (O'Neill et al., 1988). In these cases, there is, of course, no uncertainty about the patient's biological sex at the time of birth. These patients presumably experienced normal physical masculinization in utero, including putative CNS effects; in this regard, then, these patients are quite similar to those with penile agenesis and cloacal exstrophy.

As noted by Gearhart and Rock (1989; see also Jones, Park, & Rock, 1978), when there is total loss of the phallus, there are two main management options: (a) a gender re-assignment with immediate surgical feminization and subsequent hormonal feminization at puberty or (b) continued rearing of the infant or child as a boy, with eventual surgical reconstruction of the penis or phalloplasty. From a surgical point of view, the latter option is very difficult and complicated to execute during infancy or childhood. Such procedures become more viable by adulthood, and the procedures used would be along the lines that are performed in phalloplasty in female-to-male transsexuals (Dickey & Steiner, 1990).

Regarding the gender reassignment option, the theoretical and practical rationales (vis-a-vis the optimal gender policy) should by now be apparent: (a) the apparent malleability in gender identity differentiation that Money and colleagues had observed in children with physical intersex conditions; (b) the timing of the accident and gender re-assignment (within the window of the putative sensitive period for gender identity formation); (c) the medical and psychosocial difficulties that a young boy growing up without a penis would experience; and (d) the relative ease of surgical feminization (see Table 2).

Regarding gender identity differentiation, recall Money's emphasis on the importance of rearing; for example, Money et al. (1955) had argued that "in place of a theory of instinctive masculinity or femininity which is innate, the evidence of hermaphroditism lends support to a conception that, psychologically, [gender identity]9 is undifferentiated at birth and that it becomes differentiated as masculine or feminine in the course of the various experiences of growing up" (p. 308), which Diamond (1965) later characterized as a psychosexual "neutrality-at-birth" theory.

If this model is taken to its logical conclusion, then one could posit that a child with perfectly "normal" biological attributes could be successfully assigned and reared as a member of the opposite sex. Of course, an ideal test of this hypothesis would be to conduct a randomized control trial with a series of biologically normal newborn males, the particulars of which I have facetiously described elsewhere (Zucker, 1996). Of course, it is doubtful that such an experiment would be able to recruit volunteers or to pass an institutional ethics review committee. For obvious reasons, then, it is unlikely that such an "experiment of nurture" will ever be conducted. Thus, tests of the hypothesis have had to rely on special cases, and it is likely that cases of ablatio penis are the prototype in this regard.

In the early 1970s, Money and Ehrhardt (1972) (see also Money, 1975) reported on a case of ablatio penis that received widespread attention. This case involved a pair of monozygotic male twins, in which the penis of one twin was accidentally ablated (flush with the abdominal wall) during a circumcision by electrocautery at the age of 7 months. It then necrosed and sloughed off. The decision about gender reassignment was made at 17 months, with surgical castration and initial genital reconstruction occurring at 21 months.

Money (1975) reported follow-up data on this child through 9 years of age, at which time the patient was described as having many "tomboyish traits, such as abundant physical energy, a high level of activity ... and being often the dominant one in a girl's group" (p. 70). Money reported, however, that a female gender identity had apparently differentiated: "Her behavior is so normally that of an active little girl, and so clearly different by contrast from the boyish ways of her twin brother, that it offers nothing to stimulate one's conjectures" (p. 71). Thus, Money concluded that "gender identity is sufficiently incompletely differentiated at birth as to permit successful assignment of a genetic male as a girl ... and differentiates in keeping with the experiences of rearing" (p. 66). When Money first reported the case, it received widespread media attention (see "Biological Imperatives," 1973) and was noted in many pediatric, psychology, and sociology textbooks as a powerful proof of the importance of socialization influences on gender identity formation (see Diamond, 1982).

Subsequently, however, Diamond (1982) reported the further course of events for this patient. By early adolescence, the patient had rejected the female identity and began to live as a male at the age of 14 years (Diamond & Sigmundson, 1997b). Indeed, when interviewed in his early 30s, the patient's recall of his childhood gender development was that he had never felt comfortable as a girl, and his mother, father, and cotwin reported similar recollections (Colapinto, 1997, in press). At age 14, the patient received a mastectomy and began testosterone replacement therapy and surgical procedures for phallus construction were at ages 15 and 16. At age 25, the patient married a woman several years his senior and adopted her children. The patient reported an exclusive sexual attraction to females (Diamond & Sigmundson, 1997b).

The long-term psychosexual outcome of this patient, which also received recent widespread media attention ("A Tragedy Yields Insight Into Gender," 1997; Angier, 1997; Colapinto, 1997; Gorman, 1997; King, 1998), has been used as evidence against the importance of sex of rearing for gender identity formation and also as a general critique of the guidelines of psychosexual management that have been used in the care of infants with physical intersex conditions (e.g., Diamond, 1996a, 1996b, 1997, 1999; Diamond & Sigmundson, 1997a; cf. Benjamin, 1997; Glassberg, 1999; Schwarz, 1997).

There are several additional cases of ablatio penis with which to compare this case. O'Neill et al. (1988, Case 8) reported on a pair of male twins joined from the midthorax to the common perineum and who shared a single set of genitalia. The separated twin who retained the genitalia died from sequelae related to tight thoracic wall closure, but the twin who was sex-reassigned was reported, at age 10 years, to be "functioning well as a female" (p. 303). No other details were provided. However, Diamond (1999) has subsequently reported that this individual (at an unspecified age) now wishes to "transition as the male she believes she should be." Unfortunately, no further information on the course of events was described.

Several years ago, I evaluated another set of conjoined male twins who also shared one set of genitalia, one of whom was sex-reassigned at the time of surgical separation (5.5 years) (see Filler, 1988). Because of the twins' living conditions prior to surgery, they were both somewhat developmentally delayed at the time of surgery, but appeared to catchup subsequently. At age 12, the sex-reassigned twin shared many "boyish" interests with her brother, but, based on interview data with the youngster and her father, appeared to have differentiated a female gender identity. Given the apparent course of events in the O'Neill et al. (1988) case, it would be premature to draw any firm conclusions about the ultimate gender identity of this patient.

Gearhart and Rock (1989) reported on four cases of ablatio penis secondary to circumcision accidents: at birth, two at 2 days postpartum, and one at 2 months of age (this last case will be described below). In the first three cases, surgical removal of the testis and the creation of a neovagina occurred at 6 months (in two cases) and at 23 months in the third.

The first case began feminizing hormone treatment at 12 years and received vaginoplasty at 17 years (the two other cases were not old enough to start feminizing hormone therapy or to receive vaginoplasty). Regarding the first patient, Gearhart and Rock (1989) reported that she was "well adjusted" and sexually active (presumably with males, but this was not specified). Details about the gender identity development of the other two cases were not provided, but there was no indication in the report for any apparent difficulties in this regard.

Ochoa (1998) reported on a 6-month-old boy who was apparently emasculated by a dog; however, it was noted that the wound edges "suggested that a knife had been used" (p. 1116). The family agreed to a gender reassignment. At 5 years of age, the patient was said to have a "normal feminine identity" (details not provided) and, as a result, surgical feminization was performed. In adolescence, however, the patient refused to continue taking feminizing hormones and requested re-reassignment as a boy. Unfortunately, Ochoa did not provide further details other than to note that "it was not difficult to identify psychosocial factors that explained the gender disorder" (p. 1119), that is, the patient's request for gender re-reassignment and, on this point, details were also not provided.

Ochoa (1998) described four other cases of penile loss during infancy (at 4, 6, 8, and 9 months, respectively): In one case, the testicles and scrotum were also "missing ... as if made with a cutting instrument" (p. 1116); in the second case, the penis had been amputated by the infant's "mentally disturbed" mother; in the third case, the genitalia had been eaten by a dog; and, in the fourth case, the penis had been "destroyed" by a pig.

In these cases, the recommendation of gender reassignment was either refused by the parents or deferred for legal reasons: one case was subsequently lost to follow-up, but in the other three, various methods of surgical repair, including corpora cavernosa phalloplasty, were carried out. Time of last reported follow-up was at ages 12, 17, and 8 years, respectively. Although no details on the patients' gender identity were provided, one might presume that a male gender identity had differentiated.

Bradley, Oliver, Chernick, and Zucker (1998) reported on a final case of ablatio penis, originally alluded to by Money (1975, p. 65), but for whom there had been no previously reported follow-up (this was the fourth case in Gearhart & Rock, 1989). During an electrocautery circumcision at the age of 2 months, the patient sustained a burn of the skin of the entire penile shaft, and the penis eventually sloughed off. Consequently, the patient was unable to void through the urethra, resulting in a suprapubic cystostoma. The patient was subsequently hospitalized for care of the surgical complications and at age 7 months was referred to Johns Hopkins Hospital, where the remainder of the penis and the testes were removed. The suprapubic cystostoma was eventually closed, and the patient was then able to void through the urethra. Sometime between the circumcision accident and the hospital admission, the decision was made to reassign the patient as a female and to raise the baby as a girl. This was formally recognized at the time the patient was admitted to Johns Hopkins, as the infant was designated on hospital records to be a female.

The patient was interviewed at 16 years and 26 years. On both occasions, the patient appeared to have differentiated a female gender identity. Although tomboyish as a child, the patient also reported her closest friends to be girls. In adulthood, the patient reported a predominant sexual attraction to biological females, but had been sexually active with both men and women, had lived with both men and women, and characterized her sexual identity as "bisexual."

The ablatio penis case reported by Money (1975) and then by Diamond (1982; Diamond & Sigmundson, 1997b) and Colapinto (in press) contains the greatest details in which to make direct comparisons with the Bradley et al. (1998) case vis-a-vis psychosexual differentiation. Obviously, the two cases differed with regard to long-term gender identity differentiation, but there were similarities with regard to childhood gender role behavior and adult sexual orientation. Both patients had strong elements of behavioral masculinity in childhood. The Money/Diamond-Sigmundson patient had an exclusive heterosexual sexual orientation (attraction to biological females); our patient reported a bisexual sexual orientation, but a stronger attraction to biological females than to males. Lastly, the Money/Diamond-Sigmundson patient reported a heterosexual sexual identity whereas our patient reported a bisexual sexual identity (see also Meyer-Bahlburg, 1999a).

Summary

Because these cases are few and far between, it is difficult to reach any kind of definitive conclusion. In the Money/Diamond-Sigmundson case and in the case reported by Ochoa (1998), it would appear that the "experiment of nurture" vis-a-vis gender identity differentiation was unsuccessful. As in all experiments in which the null hypothesis cannot be rejected, it is difficult to ascertain why the experiment failed. Was it the result of the underlying normal prenatal masculine biology eventually overriding the efforts of gender socialization as a girl? Was it because the gender re-assignment occurred too late (i.e., because it was at the upper end of the putative age range for the sensitive period for gender identity formation)? Was it because of parental uncertainty regarding the appropriateness of the decision, thus resulting in the child receiving mixed messages in the arena of gender socialization?

These are some of the more common explanations that have been proffered in the extant literature. To these, we can add some additional possibilities. In the Money/Diamond-Sigmundson case, the patient was reported to have been relatively "tomboyish" in her childhood gender role behavior. Did such behavior cause the patient to see herself as more like a boy than like a girl, thus creating additional uncertainty regarding gender identity, along the lines that I have described earlier with regard to 5ARD patients? As this patient approached adolescence, there was also evidence that the patient became aware of sexual attraction towards biological females. Is it possible that such attractions were experienced by the patient as intolerable (a variant of "internalized homophobia") and, in an effort to "normalize" such feelings, the patient began to consider even more strongly the idea of changing genders, again as in the case of some 5-ARD patients? In this regard, Colapinto (in press) provides some supportive evidence.

In the Bradley et al. (1998) case, the experiment of nurture vis-a-vis gender identity differentiation was successful, at least as judged by the follow-up data at the age of 26 years. The only plausible explanation of our patient's differentiation of a female gender identity is that sex of rearing as a female, beginning at around age 7 months, overrode any putative influences of a normal prenatal masculine sexual biology. Although it is not possible to state with precision what constituted our patient's sex of rearing as a girl, it clearly included her parents agreeing to the sex reassignment decision (although this was easier for the mother than for the father), the provision of a stereotypical girl's name, and the patient being perceived as a girl by significant others in her social environment. In this case, then, the experiment of nurture was successful regarding female gender identity differentiation.

There are two likely reasons, perhaps related, why the gender identity development of our patient differentiated as a female, whereas it did not (at least in the long run) in the previous case. First, in our case, the decision to reassign the patient to the female sex occurred somewhere between ages 2 and 7 months; in the other case, it occurred only at 17 months, and surgical castration and vaginoplasty were done at 21 months. Second, it is possible that the parents of our patient, particularly the mother, had less ambivalence about the decision than the parents of the other patient, perhaps because the decision occurred earlier in life. Third, it is possible that our patient's awareness of sexual feelings towards biological females was not experienced as ego-dystonic (i.e., internalized homophobia), whereas such feelings were experienced in this manner by the Money/Diamond-Sigmundson patient.

General Summary

The aim of this article was to review what is known about gender identity differentiation among patients with selected physical intersex conditions. In particular, this review was focused on physical intersex conditions characterized by at least two common features: (a) the presence of ambiguous genitalia at birth, which might result in initial uncertainty as to the "appropriate" sex and gender assignment and (b) a prenatal hormonal milieu or external genital configuration (and sometimes both) that are atypical in relation to the gender in which the child is reared. In summarizing the review, I will first make some general comments pertaining to methodological matters and limitations of the extant database.

Methodological Considerations

1. Knowledge about gender identity differentiation varies considerably across syndromes. Thus, for example, there is a much larger database about the "natural history" of CAH and 5-ARD than for pAIS, micropenis, penile agenesis, and cloacal exstrophy. Thus, it is critical that the database itself be increased for these more poorly studied syndromes.

2. Within the extant databases for each syndrome, closer attention must be given to selection biases at the time of follow-up. In particular, greater efforts need to be made to reduce the number of cases lost to follow-up, as it is possible that participants differ in important respects from nonparticipants (see, e.g., Zucker et al., 1996). For example, in two of the more important studies of patients born with a micropenis, 60% (Reilly & Woodhouse, 1989) and 67% (Money & Norman, 1988) of the patients were unavailable for follow-up.

3. The reliance on standardized assessment protocols varies considerably across syndromes, with some accounts of gender identity differentiation quite poorly described. Thus, it is important that future researchers of gender identity differentiation (and other aspects of psychosexual differentiation), particularly in the less wellstudied syndromes, rely on more rigorous assessment methods and techniques.

4. The availability of long-term follow-up data also varies considerably across the different syndromes. Even among the better studied syndromes (e.g., CAH), adult follow-ups have generally been when the patients were in their early 20s (cf. Ehrhardt, Evers, & Money, 1968). Among gender dysphoric adults without known somatic intersexuality (and who have a homosexual sexual orientation), the decision to formally seek out sex-reassignment procedures often occurs in the mid-20s to early 30s (e.g., Blanchard, Clemmensen, & Steiner, 1987). Thus, among young intersex adults who may be struggling with gender identity issues, the "period of risk" may not have passed, so even longer term follow-up is likely required in order to gain a more accurate picture about gender identity differentiation. Overall, then, definitive conclusions about the natural history of gender identity differentiation are severely hampered by the variability in the quantity and quality of follow-up information.

5. For some of the syndromes reviewed in this article, it was noted that the gender assignment in infancy varies considerably. For example, among patients with 5-ARD, a minority of such individuals, particularly those living in Western countries, are assigned to the female sex and then proceed with surgical feminization of the genitalia, gonadectomy, and administration of feminizing hormones at puberty. Much less is known about the natural history of gender identity differentiation (and other aspects of psychosexual differentiation) among patients treated in this manner than among those from non-Western countries in which the treatment history is very different. Comparative data for these differing treatment histories are urgently needed and would greatly facilitate a more accurate understanding of the factors involved in gender identity differentiation.

Conclusions About Gender Identity Differentiation

The empirical database suggests the following broad conclusions:

1. Among the physical intersex conditions reviewed in this article, gender identity differentiation appears to be more variable than among individuals with apparently "normal" physical sex differentiation (cf. Meyer-Bahlburg, 1994, 1999b). In the latter population, the most relevant comparison condition is that of GID in children, adolescents, and adults (American Psychiatric Association, 1994). In GID, there is no ambiguity of the external genitalia at birth, and there is no evidence for a gross hormonal abnormality, as in the case of some of the physical intersex conditions, such as CAH in genetic females. Thus, GID occurs in individuals who are unambiguously assigned at birth to the male or female sex. The prevalence of GID in children is not known, but in adults it has been estimated to occur in 1:11,000-30,000 genetic males and 1:30,400-100,000 genetic females (American Psychiatric Association, 1994; Bakker et al., 1993). These figures are probably the most relevant since, among children with GID, prospective follow-up studies suggest that the majority of youngsters resolve their gender dysphoria and do not grow up to request sex-reassignment (Green, 1987; Zucker & Bradley, 1995).

2. Although gender identity differentiation appears to be more variable among individuals with the physical intersex conditions reviewed in this article, it is also apparent that there is significant and meaningful variability across syndromes. For example, although instances of gender change in genetic females with CAH appear to be more common than occurs in the general population (Meyer-Bahlburg et al., 1996; Zucker et al., 1996), it is noticeably less common than which appears to occur in 5-ARD, where the majority of cases reported so far appear to change genders. Thus, we need to account for two facts: (a) Why is gender identity differentiation more variable in general among people with physical intersex conditions than it is among people without such conditions? (b) How does one account for the cross-syndrome variability in gender identity differentiation?

Regarding the first question, there are two candidate explanations. The first explanation is that elements of the physical intersex condition itself serve as a predisposing factor for greater lability in gender identity differentiation (Hoenig, 1985). Thus, in genetic females with CAH assigned to the female sex and reared as girls, for example, the sexatypical exposure to prenatal androgens increases the potential for cross-gender behavior, including gender identity differentiation (Collaer & Hines, 1995).

The second explanation is that elements of the gender specific rearing environment also serve as a predisposing factor for greater lability in gender identity differentiation. It is this second point for which the literature is most divided. For example, among patients with 5-ARD who are assigned to the female sex and "raised" as girls, there has been a great deal of debate regarding the extent to which the gender of rearing has been truly unambiguous or for which there is, in fact, some uncertainty. On this point, the matter is far from settled. Here, prospective studies would be particularly informative, in order to track better the developmental course of gender socialization and the entire literature on physical intersex conditions is particularly poor in this regard.

Regarding the second question, there are also two candidate explanations. The first explanation is that elements of the physical intersex conditions themselves vary in their predisposing "power" to increase the potential for cross-gender behavior, including gender identity differentiation. Appraisal of this hypothesis is, however, very difficult because of the manner in which physical intersex conditions are treated, both medically and socially. For example, in the case of CAH in genetic females, the typical medical treatment regimen is to block the excess production of adrenal androgens with corticosteroids and to surgically "normalize" the masculinized genitalia. In contrast, among patients with 5-ARD raised as girls, at least in some cohorts, there is no medical intervention at all: Thus, the girl grows up with ambiguous genitalia, and there is progressive physical masculinization at the time of puberty. Because of this kind of confound, it cannot be concluded with certainty that the influence of prenatal androgen exposure itself is more powerful among 5-ARD patients than it is among genetic females with CAH in accounting for the higher rate of gender change in the former group.

The second explanation is that the consistency (or lack thereof) of the gender-rearing environment varies across syndromes, and it is this variability that accounts for the varying rates of gender change or gender dysphoria that occur in the different syndromes. Apart from the inherent difficulty in specifying the nature of the gender-rearing environment, it is likely that the variation, if it actually occurs, interacts in complex ways in which the physical intersex condition itself is treated both medically and surgically.

3. Perhaps the most important conclusion from this review is to reiterate the importance of respecting the complexity of gender identity differentiation in particular and psychosexual differentiation in general. The gender specific social environment clearly matters, because there are so many instances of gender identity differentiating in accordance with the gender assignment. Perhaps the most striking illustration of this comes from the Bradley et al. (1998) adult follow-up data on one genetic male with ablatio penis raised as a girl. It is hard to account for this patient's female gender identity without implicating some kind of environmental influence.

At the same time, biological factors also matter. For example, the shift towards masculinized gender role behavior seen in girls with CAH or pAIS are likely related, at least in part, to their exposure to prenatal androgens and the pattern is very similar to what has been observed in experimental studies of nonhuman primates (see, e.g., Goy, Bercovitch, & McBriar, 1988; Goy & McEwen, 1980). Although such effects on gender identity differentiation are likely more indirect, there is probably some alteration in the threshold, interacting with other factors, that results in an increased risk for gender change in some individuals (see Meyer-Bahlburg et al., 1996).

Despite the increasing recognition of the complex factors involved in psychosexual differentiation, the seductive lure of either psychosocial or biological reductionism remains a very strong temptation, to which many writers in the field succumb. Recent advances in the field of animal sexology point to the importance of paying serious attention to multifactorial models in our understanding of psychosexual differentiation in humans. For example, Juraska (1998) has shown how the typical sex difference in the number of neurons in the corpus callosum of rats is exquisitely sensitive to, and modified by, the rearing environment. In nonhuman primates, Wallen (1996) has recently summarized ways in which the social environment either attenuates or exacerbates typical sex differences in behavior, thus providing illustrations of how "nature needs nurture." Thus, it is in the more precise identification of the transactional nature of biological and psychosocial influences on gender identity differentiation in particular and psychosexual differentiation in general that future empirical inquiry must continue to invest its efforts.

1 The word normal is intentionally surrounded by quotation marks to alert the reader to one of the current controversies in the literature on physical intersex conditions, namely whether they should be considered "abnormal" or simply variants of typical or normal physical sex differentiation in ordinary biological males and females (see, e.g., Blackless et al., in press). I believe that physical intersex conditions are "abnormal," or, to use Money's (1994) descriptor, "sex errors of the body." I reach this conclusion by following King's (1945) definition of normal: "that which functions in accordance with its design." Given our understanding of the processes that govern normal physical sex differentiation and the processes that go awry in the differentiation of physical intersex conditions, it is hard to argue that such conditions are simply variants from the norm. In using the term abnormal, however, it is important, particularly from a moral point of view, to distinguish between the condition and the person who has the condition. In the 19th century, for example, hermaphroditism was viewed as a teratologic monstrosity and, at least for some people, the hermaphrodite was viewed as a "monster." The term teratology is of Greek origin (teras: monster or marvel). A less pernicious definition is that of an abnormality or serious deviation in the growth or structure of an organism. It is the blurring of the person and the person's condition that is particularly problematic. Of course, the psychological struggles that some people with physical intersex conditions experience cannot be comfortably distanced by referring to the past. Chase (1998a), the founder of a contemporary support group for intersexuals, the Intersex Society of' North America, lucidly describes her efforts in adulthood to over-come the feeling of being a "freak" (see also Holmes, 1994).

2 The terms "gender role and orientation" require some explication. Two years prior to the Money et al. (1957) article, Money (1955) had coined the term gender role, which was defined as "all those things that a person says or does to disclose himself or herself as having the status of boy or man, girl or woman, respectively. It includes, but is not restricted to, sexuality in the sense of eroticism" (p. 254). As I have noted in detail elsewhere (Zucker & Bradley, 1995, pp. 2-6), over the past several decades, Money's original use of the term gender role has been decomposed into three conceptually distinct component parts that are identified by the terms gender identity, gender role, and sexual orientation. Briefly, gender identity refers to a child's basic sense of self as a boy or as a girl; gender role refers to a child's preferential adoption of behaviors stereotypically associated with masculinity and femininity (e.g., in the domains of toy play, role play, peer preference, and so on), and sexual orientation refers to a person's responsiveness to sexual stimuli, of which the sex of one's partner is probably the most salient dimension. It should also be noted that the construct of sexual orientation needs to be further distinguished from the construct of sexual identity, which refers to a person's self-labeling as, for example, heterosexual, homosexual, or bisexual. Although often congruent with each other, sexual orientation and sexual identity are not always so. In this article, I will, where appropriate, attempt to use the more differentiated terms of gender identity, gender role, and sexual orientation when discussing the extant data on psychosexual differentiation in people with physical intersex conditions.

3 "Gender of rearing" would be the preferred description, but I will retain Money's use of the descriptor "sex of rearing" for the sake of historical continuity. But the reader should keep in mind that "rearing" really refers to putative gender-specific socialization techniques of the infant as a boy or a girl.

4 Occasionally, the diagnosis is made in childhood (e.g., if there is notable swelling in the perineum because of the undescended testes). On rare occasions, the diagnosis is suspected when the karyotyping of the sex chromosomes taken from amniotic fluid appears to contradict the genital configuration observed during fetal ultrasound.

5 Money (1991, Chapter 2, Case 2) reported on one patient, whose testes were discovered at 18 months (see also Crawford, 1970). A physician recommended gender reassignment on the grounds that physical masculinization could occur with exogenous hormone treatment. Unfortunately, the reassignment occurred during an era in which the natural history of cAIS was still unknown. The patient's body never masculinized. Although gender identity differentiated as male@ the patient was extremely isolated socially, and eventually committed suicide in adulthood. To my knowledge, this is the only case in the literature of male gender identity differentiation in a patient with cAIS albeit with a tragic outcome.

6 In many individuals with CAH, there is a deficiency in aldosterone, which causes low serum sodium levels, high serum potassium levels, and vascular collapse-the so-called "salt-wasting" crises in severe cases (New et al., 1996). Among patients with classical CAH, it has been traditional practice to classify them as either simple virilizers (SV; i.e., with no salt-wasting) or salt-wasters (SW). Although it is beyond the scope of this article to review the matter in detail, the SW-SV distinction is important from a psychosexual perspective because there is some evidence that the SW group is, on average, more severely physically masculinized genitally (Prader stages) at birth (for review, see Zucker et al., 1996), which likely indicates greater prenatal androgenization, and there is corresponding evidence for greater behavioral masculinization (see, e.g., Dittmann et al., 1990; Meyer-Bahlburg et al., 1999; Slijper, 1984; Zucker et al., 1996).

7 Prior to the discovery that two enzymes were involved in the pathophysiology, the syndrome was described simply as 5u,-reductase deficiency. Over the past few years, the syndrome has been called 5(x-reductase 2 deficiency. In genetic females with 5-ARD, it appears to have no substantive role in endocrine physiology and has no known adverse effects (Milewich et al., 1995; Wilson, Griffin, & Russell, 1993).

8 According to Wilson et al. (1993), the 19th-century French hermaphrodite, Herculine Barbin (Barbin, 1980), who changed her legal sex from female to male, but, because of a series of complex social events, committed suicide in adulthood, may well be the first published case in the literature of an individual with 5-ARD.

9 I have inserted here the term gender identity in place of "sexuality," because the discussion focuses on a more specific aspect of psychosexual differentiation.

References

A TRAGEDY YIELDS INSIGHT INTO GENDER. (1997, March 24). Newsweek, 66.

ABDULLAH, M. A., KATUGAMPOLA, M., AL-HABIB, S., AL-JURAYYAN, N., AL-SAMARRAi, A., AL-NUAIM, A., PATEL, P. J., & NiAzi, M. (1991). Ambiguous genitalia: Medical, socio-cultural and religious factors affecting management in Saudi Arabia. Annals of Tropical Paediatrics, 11, 343-348.

AL-ATTiA, H. M. (1996). Gender identity and role in a pedigree of Arabs with intersex due to 5 alpha reductase-2 deficiency. Psychoneuroendocrinology, 21, 651-657.

ALLEN, L. E., HARDY, B. E., & CHURCHILL, B. M. (1982). The surgical management of the enlarged clitoris. Journal of Urology, 128, 351-354.

AMERICAN ACADEMY OF PEDIATRICS. (1996). Timing of elective surgery on the genitalia of male children with particular reference to the risks, benefits, and psychological effects of surgery and anesthesia. Pediatrics, 97, 590-594.

AMERICAN PSYCHIATRIC ASSOCIATION. (1987). Diagnostic and statistical manual of men

tal disorders (3rd ed., rev.). Washington, DC: Author.

AMERICAN PSYCHIATRIC AssocIATION. (1994). Diagnostic and statistical manual of mental disorders (4th ed.). Washington, DC: Author.

ANGIER, N. (1997, March 14). Sexual identity not pliable after all, report says. New York Times, pp. Al, A10.

ASSAEL, M., LANCET, M., & SHANI, A. (1976). Incomplete testicular feminization syndrome: Clinical and psychiatric aspects. Psychiatria Clinica, 9, 118-129.

AzpiRoz, J. (1971). Agenesis of the penis. Journal of Pediatric Surgery, 6, 74.

BAILEY, J. M., & ZUCKER, K. J. (1995). Childhood sex-typed behavior and sexual orientation: A conceptual analysis and quantitative review. Developmental Psychology, 31, 43-55. BAKKER, A., vAN KESTEREN, P. J. M., GoOREN, L. J. G., & BEZEMER, P. D. (1993). The

prevalence of transsexualism in The Netherlands. Acta Psychiatrica Scandinavica, 87, 237-238.

BANCROFT, J. (1991). John Money: Some comments on his early work. Journal of Psychology & Human Sexuality, 4(2), 1-8.

BARBIN, H. (1980). Herculine Barbin: Being the recently discovered memoirs of a nineteenth-century French hermaphrodite (R. McDougall, Trans.). New York: Pantheon Books. BARTsCH, G., DECRISTOFORO, A., & SCHWEIKERT, U. (1987). Pseudovaginal perineoscrotal

hypospadias: Clinical, endocrinological and biochemical characterization of a patient. European Urology, 13, 386-389.

BAum, M. J. (1979). Differentiation of coital behavior in mammals: A comparative analysis. Neuroscience & Biobehavioral Reviews, 3, 265-284.

BEHESHTI, M., HARDY, B, E., CHURCHILL, B. M., & DANEMAN, D. (1983). Gender assignment in male pseudohermaphrodite children. Urology, 22, 604-607.

BEM, D. J. (1996). Exotic becomes erotic: A developmental theory of sexual orientation. Psychological Review, 103, 320-335.

BENJAMIN, J. T. (1997). Sex reassignment at birth [Letter to the editor]. Archives of Pediatrics and Adolescent Medicine, 151, 1062-1063.

BERENBAUM, S. A., & BAILEY, J. M. (1998, May). Variation in female gender identity: Evidence from girls with congenital adrenal hyperplasia, tomboys, and typical girls. Paper presented at the meeting of the Midwestern Psychological Association, Chicago, IL.

BERENBAUM, S. A., & HINES, M. (1992). Early androgens are related to childhood sex-typed toy preferences. Psychological Science, 3, 203-206.

BIN-ABBAS, B., CONTE, F. A., GRUMBACH, M. M., & KAPLAN, S. L. (1999). Congenital hypogonadotropic hypogonadism and micropenis: Effect of testosterone treatment on adult penile size-Why sex reversal is not indicated. Journal of Pediatrics, 134, 579-583.

BIOLOGICAL IMPERATIVES. (1973, January 8). T/me, 34.

BIRNHOLZ, J. C. (1983). Determination of fetal sex. New England Journal of Medicine, 309, 942-944.

BLACKLESS, M., CHARuvAsTRA, A., DERRYCK, A., FAuSTO-STERLING, A., LAVZANNE, K., & LEE, E. (in press). How sexually dimorphic are we? Review and synthesis. American Journal of Human Biology.

BLANCHARD, R., CLEMMENSEN, L. H., & STEINER, B. W. (1987). Heterosexual and homosexual gender dysphoria. Archives of Sexual Behavior, 16, 139-152.

BORNSTEIN, M. H. (Ed.). (1987). Sensitive periods in development: Interdisciplinary perspectives. Hillsdale, NJ: Erlbaum.

BORNSTEIN, M. H. (1989). Sensitive periods in development: Structural characteristics and causal interpretations. Psychological Bulletin, 105, 179-197.

BOUDON, C., LUMBROSO, S., LOBACCARO, J. M., SZARRAS-CZAPNIK, M., ROMER, T. E., GARANDEAU, R, MONTOYA, P., & SULTAN, C. (1995). Molecular study of the 5 (x-reductase type 2 gene in three European families with 5 (x-reductase deficiency. Journal of Clinical Endocrinology & Metabolism, 80, 2149-2153.

BOWLBY, J, (1969). Attachment and loss: Vol. 1. Attachment. New York: Basic Books. BRADLEY, S. J., OLIVER, G. D., CHERNICK, A. B., & ZUCKER, K. J. (1998). Experiment of

nurture: Ablatio penis at 2 months, sex reassignment at 7 months, and a psychosexual follow-up in young adulthood. Pediatrics, 102, E91-E95. (Available on the World Wide Web http://www.pediatrics. org/egi/content/ full/102/l/e9)

BROwN, T. R., SCHERER, P. A., CHANG, Y., MIGEON, C. J., GHIRRI, P., MURONO, K., & ZHou, Z. (1993). Molecular genetics of human androgen insensitivity. European Journal of Pediatrics, 152(Suppl. 2), S62-S69.

BRUCH, S. W., MEULI, M., & HARRISON, M. R. (1996). Immediate reconstruction for penile agenesis. Journal of Pediatric Surgery, 31, 1152-1154.

CANTU, J. M., CORONA-RIvERA, E., DIAZ, M., MEDINA, C., ESQUINCA, E., CORTES-GALLECOS, V., VACA, G., & HERNANDEZ, A. (1980). Post-pubertal female psychosexual orientation in incomplete male pseudohermaphroditism type 2 (5o,-reductase deficiency). Acta Endocrinologica, 94, 273-279.

CANTU, J. M., HERNANDEZ-MONTES, H., DEL CASTILLO, V., CORTES-GALLEGOS, V., SANDOVAL, R., ARMENDARES, S., & PARRA, A. (1976). Potential fertility in incomplete male pseudohermaphroditism type 2. Revista de Investigacion Clinica, 28, 177-182.

CARPENTER, T. 0., IMPERATo-McGINLEY, J., BoULWARE, S. D., WEISS, R. M., SHACKLETON, C., GRIFFIN, J. E., & WILSON, J. D. (1990). Variable expression of 5a-reductase deficiency: Presentation with male phenotype in a child of Greek origin. Journal of Clinical Endocrinology and Metabolism, 71, 318-322.

CARSON, D. J., OKLTNO, A., LEE, P. A., STETTEN, G., DIDOLKAR, S. M., & MIGEON, C. J.

(1982). Amniotic fluid steroid levels: Fetuses with adrenal hyperplasia, 46,XXY fetuses, and normal fetuses. American Journal of Diseases of Childhood, 136, 218-222.

CHAN-CUA, S., FREIDENBERG, G., & JONES, K. L. (1989). Occurrence of male phenotype in genotypic females with congenital virilizing adrenal hyperplasia. American Journal of Medical Genetics, 34, 406-412.

CHASE, C. (1998a). Hermaphrodites with attitude: Mapping the emergence of intersex political activism. GLQ, 4, 189-211.

CHASE, C. (1998b). Surgical progress is not the answer to intersexuality. Journal of Clinical Ethics, 9, 385-392.

CICCHETTI, D. (1993). Developmental psychopathology: Reactions, reflections, projections. Developmental Review, 13, 471-502.

COLAPINTO, J. (1997, December 11). The true story of John/Joan. Rolling Stone, 54-58, 60, 62, 64, 68, 72-73, 92, 94-97.

COLAPINTO, J. (in press). As nature made him: The boy who was raised as a girl. New York: HarperCollins.

COLLAER, M. L., & HINES, A (1995). Human behavioral sex difference: A role for gonadal hormones during early development? Psychological Bulletin, 118, 55-107.

CORRALL, R. J. M., WAKELIN, K., O'HARE, J. P., O'BRIEN, 1. A. D., IsHMAIL, A. A. A., & HONOUR, J. (1984). 5a-reductase deficiency: Diagnosis via abnormal plasma levels of reduced testosterone derivatives. Acta Endocrinologica, 107, 538-543.

COSTA, E. M. F., MENDONCA, B. B., INACIO, M., ARNHOLD, L J. R, SiLvA, F. A. Q., & LODOVICI, 0. (1997). Management of ambiguous genitalia in pseudohermaphrodites: New perspectives on vaginal dilation. Fertility and Sterility, 67, 229-232.

CRAWFORD, J. D. (1970). Syndromes of testicular feminization. Clinical Pediatrics, 9, 165-178.

DE MARNEFFE, D. (1997). Bodies and words: A study of young children's genital and gender knowledge. Gender & Psychoanalysis, 2, 3-33.

DELK, J. L., MADDEN, R. B., LIVINGSTON, M., & RYAN, T. T. (1986). Adult perceptions of the infant as a function of gender labeling and observer gender. Sex Roles, 15, 527-534. DESLYPERE, J. P., CoucKE, W., ROBBE, N., & VERMEULEN, A. (1985). 5 alpha-reductase

deficiency: An infrequent cause of male pseudohermaphroditism, Acta Clinica Belgica, 40, 240-246.

DEWHURST, J., CHAPMAN, M., MuRAm, D., & DONNISON, B. (1983). 5 alpha-reductase deficiency in 46XY sisters. Pediatric and Adolescent Gynecology, 1, 85-95.

DIAMOND, M. (1965). A critical evaluation of the ontogeny of human sexual behavior. Quarterly Review of Biology, 40, 147-175.

DIAMOND, M. (1982). Sexual identity, monozygotic twins reared in discordant sex roles and a BBC follow-up. Archives of Sexual Behavior, 11, 181-186.

DIAMOND, M. (1996a). Prenatal predisposition and the clinical management of some pediatric conditions. Journal of Sex & Marital Therapy, 22, 139-147.

DIAMOND, M. (1996b). Response: Considerations for sex assignment. Journal of Sex & Marital Therapy, 22, 161-174.

DIAMOND, M. (1997). Sexual identity and sexual orientation in children with traumatized or ambiguous genitalia. The Journal of Sex Research, 34, 199-211.

DIAMOND, M. (1999). Pediatric management of ambiguous and traumatized genitalia. Journal of Urology, 162, 1021-1028.

DIAMOND, M., & SIGMUNDSON, H. K. (1997a). Management of intersexuality: Guidelines for dealing with persons with ambiguous genitalia. Archives of Pediatrics and Adolescent Medicine, 151, 1046-1050.

DIAMOND, M., & SIGMUNDSON, H. K. (1997b). Sex reassignment at birth: Long-term review and clinical implications. Archives of Pediatrics and Adolescent Medicine, 151, 298304.

DicKEY, R., & STEINER, B. W. (1990). Hormone treatment and surgery. In. R. Blanchard & B. W. Steiner (Eds.), Clinical management of gender identity disorders in children and adults (pp. 137-158). Washington, DC: American Psychiatric Press.

DITTMANN, R. W (1998). Ambiguous genitalia, gender-identity problems, and sex reassignment. Journal of Sex & Marital Therapy, 24, 255-271.

DITTMANN, R. W, KAPPEs, M. E., & KAPPEs, M. H. (1992). Sexual behavior in adolescent and adult females with congenital adrenal hyperplasia. Psychoneuroendocrinology, 17, 153-170.

DITTMANN, R. W, KAPPEs, M. H., KAPPES, M. E., BORGER, D., MEYER-BAHLBURG, H. F. L., STEGNER, H., WILLIG, R. H., & WALLIS, H. (1990). Congenital adrenal hyperplasia 11: Gender-related behavior and attitudes in female salt-wasting and simple-virilizing patients. Psychoneuroendocrinology, 15, 421-434.

DIXSON, A. F. (1998). Primate sexuality: Comparative studies of the prosimians, monkeys, apes, and human beings. Oxford: Oxford University Press.

DONAHOE, P. K., CRAWFORD, J. D., & HENDRIN, W. H. (1977). Management of neonates and children with male pseudohermaphroditism. Journal of Pediatric Surgery, 12, 10451057.

DONAHOE, R K., & SCHNITZER, J. J. (1996). Evaluation of the infant who has ambiguous genitalia, and principles of operative management. Seminars in Pediatric Surgery, 5, 30-40.

DREGER, A. D. (1995a). Doubtful sex: Cases and concepts of hermaphroditism in France and Britain, 1868-1915. Unpublished doctoral dissertation, Indiana University.

DREGER, A. D. (1995b). Doubtful sex: The fate of the hermaphrodite in Victorian medicine. Victorian Studies, 38, 335-370.

DREGER, A. D. (1998a). "Ambiguous sex" or ambivalent medicine? Ethical issues in the treatment of intersexuality. Hastings Center Report, 28(3), 24-35.

DREGER, A. D. (1998b). Hermaphrodites and the medical invention of sex. Cambridge: Harvard University Press.

DREGER, A. D. (1998c). A history of intersexuality: From the age of gonads to the age of consent. Journal of Clinical Ethics, 9, 345-355.

DRURY, R. B., & SCHWARZELL, H. H. (1935). Congenital absence of penis. Archives of Surgery, 30, 236-242.

EHRHARDT, A. A., & BAKER, S. W. (1974). Fetal androgens, human central nervous system differentiation, and behavior sex differences. In R. C. Friedman, R. M. Richart, & R. L. Vande Wiele (Eds.), Sex differences in behavior (pp. 33-51). New York: Wiley.

EHRHARDT, A. A., EPSTEiN, R., & MONEY, J. (1968). Fetal androgens and female gender identity in the early-treated adrenogenital syndrome. Johns Hopkins Medical Journal,

122,160-167.

EHRHARDT, A. A., EVERS, K., & MONEY, J. (1968). Influence of androgen and some

aspects of sexually dimorphic behavior in women with the late-treated adrenogenital syndrome. Johns Hopkins Medical Journal 123, 115-122.

ELLIS, A. (1945). The sexual psychology of human hermaphrodites. Psychosomatic Medicine, 7, 108-125.

ELSAYED, S. M., AL-MAGHRABY, M., HAFEIZ, H. B., & TAHA, S. A. (1988). Psychological aspects of intersex in Saudi patients. Acta Psychiatrica Scandinavica, 77, 297-300.

EVANS, J. A., ERDILE, L. B., GREENBERG, C. R., & CHUDLEY, A. E. (1999). Agenesis of the penis: Patterns of associated malformations. American Journal of Medical Genetics, 84, 47-55.

FAGOT, B. 1. (1995). Psychosocial and cognitive determinants of early gender-role development. Annual Review of Sex Research, 6, 1-31.

FARAH, R., & RENO, G. (1972). Congenital absence of the penis. Journal of Urology, 107, 154-155.

FARKAS, A., & RoSLER, A. (1993). Ten years experience with masculinizing genitoplasty in male pseudohermaphroditism due to 170-hydroxysteroid dehydrogenase deficiency. European Journal of Pediatrics, 152(Suppl. 2), S88-S90.

FAuSTO-STERLING, A. (1993, March/April). The five sexes: Why male and female are not enough. The Sciences, 33, 20-25.

FEITZ, W F. J., VAN GRUNSVEN, E. J. K. J. E. M., FROELING, F. M. J. A., & DE VRIES, J. D. M. (1994). Outcome analysis of the psychosexual and socioeconomical development of adult patients born with bladder exstrophy. Journal of Urology, 152, 1417-1419.

FELDMAN, K. W, & SMITH, D. W (1975). Fetal phallic growth and penile standards for newborn male infants. Journal of Pediatrics, 86, 395-398.

FILLER, R. M. (1988). [Discussion]. Annals of Surgery, 208, 311-312.

FINDLAY, D. (1995). Discovering sex: Medical science, feminism and intersexuality. Canadian Review of Sociology and Anthropology, 32, 25-52.

FISHER, L. K., KOGUT, M. D., MOORE, R. J., GOEBELSMANN, U., WEITZMAN, J. J., ISAACS, H., GRIFFIN, J. E., & WILSON, J. D. (1978). Clinical, endocrinological, and enzymatic characterization of two patients with 5(x-reductase deficiency: Evidence that a single enzyme is responsible for the 5(x-reduction of cortisol and testosterone. Journal of Clinical Endocrinology and Metabolism, 47, 653-664.

FLETCHER, J. C., & EvANs, M. 1. (1983). Maternal bonding in early fetal ultrasound examinations. New England Journal of Medicine, 308, 392-393.

FOREST, M. G. (1985). Pitfalls in prenatal diagnosis of 21-hydroxylase deficiency by amniotic fluid steroid analysis? A six years experience in 102 pregnancies at risk. Annals of the New York Academy of Sciences, 458, 130-147.

FORTI, G., FALCHETTi, A., SANTORO, S., DAvis, D. L., WILSON, J. D., & RusSELL, D. W. (1996). Steroid 5a-reductase 2 deficiency: Virilization in early infancy may be due to partial function of mutant enzyme. Clinical Endocrinology, 44, 477-482.

FRANK, J. D. (1997). [Editorial comment]. British Journal of Urology, 79, 789. FRATiANNi, C. M., & IMPERATo-McGINLEY, J. (1994). The syndrome of 5a-reductase deficiency. Endocrinologist, 4, 302-314.

FRIDELL, S. R., ZUCKER, K. J., BRADLEY, S. J., & MAING, D. M. (1996). Physical attractiveness of girls with gender identity disorder. Archives of Sexual Behavior, 25, 17-31.

GAUTIER, T., SALIENT, J., PENA, S., IMPERATo-McGINLEY, J., & PETERSON, R. E. (1981). Testicular function in 2 cases of penile agenesis. Journal of Urology, 126, 556-557. GEARHART, J. P., & ROCK, J. A. (1989). Total ablation of the penis after circumcision

with electrocautery: A method of management and long-term follow-up. Journal of Urology, 142, 799-801.

GLASSBERG, K. 1. (1998). The intersex infant: Early gender assignment and surgical reconstruction. Journal of Pediatric and Adolescent Gynecology, 11, 151-154.

GLAsSBERG, K. 1. (1999). Editorial: Gender assignment and the pediatric urologist.

Journal of Urology, 161, 1308-1310.

GLUER, S., FUCHS, J., & MILDENBERGER, H. (1998). Diagnosis and current management of penile agenesis. Journal of Pediatric Surgery, 33, 628-631.

GOLDSCHMIDT, R. B. (1923). The mechanism and physiology of sex determination (W. J. Dakin, Trans.). London: Methuen.

GOREN, L., & COHEN-KETTENIS, P. T. (1991). Development of male gender identity/role and a sexual orientation towards women in a 46,XY subject with an incomplete form of the androgen insensitivity syndrome. Archives of Sexual Behavior, 20, 459-470.

GORMAN, C. (1997, March 24). A boy without a penis. Time, 31.

Goy, R. W., BERCOVITCH, F. B., & McBRIAR, M. C. (1988). Behavioral masculinization is independent of genital masculinization in prenatally androgenized female rhesus macaques. Hormones and Behavior, 22, 552-571.

Gov, R. W, & McEWEN, B. S. (1980). Sexual differentiation of the brain. Cambridge, MA: MIT Press.

GREEN, R. (1987). The "sissy boy syndrome" and the development of homosexuality. New Haven, CT: Yale University Press.

GREEN, R., STOLLER, R. J., & MAcANDREW, C. (1966). Sex assignment and reassignment: Physicians' attitudes in the management of hermaphroditic children. American Journal of Diseases of Childhood, 111, 524-528.

GREENBERG, J. A. (1999). Defining male and female: Intersexuality and the collision between law and biology. Arizona Law Review, 41, 265-328.

GRIFFIN, J. E. (1992). Androgen resistance: The clinical and molecular spectrum. New England Journal of Medicine, 326, 611-618.

GROSS, D. J., LANDAU, H., KOHN, G., FARKAS, A., ELRAYYFS, E., EL-SHAWWA, R., LASCH, E. E., & ROSLER, A. (1986). Male pseudohermaphroditism due to 17p-hydroxysteroid dehydrogenase deficiency: Gender reassignment in early infancy. Acta Endocrinologica, 112, 238-246.

H?ipSON, J. L., HAMPSON, J. G., & MONEY, J. (1955). The syndrome of gonadal agenesis (ovarian agenesis) and male chromosomal pattern in girls and women: Psychologic studies. Bulletin of the Johns Hopkins Hospital, 97, 207-226.

HAQQ, C. M,, & DoNAHOE, P. K. (1998). Regulation of sexual dimorphism in mammals. Physiological Reviews, 78, 1-33.

HARTocOLLIS, A. (1999, February 16). A turn toward tradition: Old standards are making a comeback in children's names. New York Times, p. C17.

HAYDEN, P. W, CHAPmAN, W H., & STEVENSON, J. K. (1973). Exstrophy of the cloaca. American Journal of Diseases of Childhood, 125, 879-883.

HERDT, G. (1990). Mistaken gender: 5-alpha reductase hermaphroditism and biological reductionism in sexual identity reconsidered. American Anthropologist, 92, 433-446. HERDT, G. (Ed.). (1994). Third sex, third gender: Beyond sexual dimorphism in culture and history. New York: Zone Books.

HERDT, G. H., & DAVIDSON, J. (1988). The Sambia "Turnim-man": Sociocultural and clinical aspects of gender formation in male pseudohermaphrodites with 5-alpha reductase deficiency in Papua New Guinea. Archives of Sexual Behavior, 17, 33-56.

HESS, E. H. (1973). Imprinting. New York: Van Nostrand.

HINMAN, F. (1951a). Advisability of surgical reversal of sex in female pseudohermaphrodism. Journal of the American Medical Association, 146, 423-429.

HINMAN, F. (1951b). Sexual trends in female pseudohermaphrodism. Journal of Clinical Endocrinology and Metabolism, 11, 477-486.

HINMAN, F. (1972). Microphallus: Characteristics and choice of treatment from a study of 20 cases. Journal of Urology, 107, 499-505.

HOCHBERG, Z., GARDOS, M., & BENDERLY, A. (1987). Psychosexual outcome of assigned females and males with 46,XX virilizing congenital adrenal hyperplasia. European Journal of Pediatrics, 146, 497-499.

HOCHBERG, Z., CHAYEN, R., REiss, N., FAME, Z., MAKLER, A., MUNICHOR, M., FARKAs, A.,

GOLDFARB, H., OHANA, N., & MORT, 0. (1996). Clinical, biochemical, and genetic findings in a large pedigree of male and female patients with 5(x-reductase 2 deficiency. Journal of Clinical Endocrinology and Metabolism, 81, 2821-2827.

HODGINS, M. B., CLAYTON, R. N., & LONDON, D. R. (1977). Androgen metabolism and binding in skin and fibroblasts from a case of incomplete male pseudohermaphroditism. Journal of Endocrinology, 75, 24P

HOENIG, J. (1985). The origin of gender identity. In B. W. Steiner (Ed.), Gender dysphoria: Development, research, management (pp. 11-32). New York: Plenum Press.

HOLMES, M. M. L. (1994). Medical politics and cultural imperatives: Intersexual identities beyond pathology and erasure. Unpublished master's thesis, York University, North York, Ontario.

HOWELL, C., CALDAMONE, A., SNYDER, H., ZIEGLER, M., & DUCKETT, J. (1983). Optimal management of cloacal exstrophy. Journal of Pediatric Surgery, 18, 365-369.

HURTIG, A. L. (1992). The psychosocial effects of ambiguous genitalia. Comprehensive Therapy, 18, 22-25.

IMPERATo-McGINLEY, J., GAUTIER, T., PETERSON, R. E., & SHACKLETON, C. (1986). The prevalence of 5a-reductase deficiency in children with ambiguous genitalia in the Dominican Republic. Journal of Urology, 136, 867-873.

ImPFRATo-McGINLEY, J., GUERRERO, L., GAUTIER, T., & PETERSON, R. E. (1974). Steroid 5(x-reductase deficiency in man: An inherited form of male pseudohermaphroditism. Science, 186, 1213-1215.

ImPERATo-McGINLEY, J., MILLER, M., WILSON, J. D., PETERSON, R. E., SHACKLETON, C., & GAJDUSEK, D. C. (1991). A cluster of male pseudohermaphrodites with 5a-reductase deficiency in Papua New Guinea. Clinical Endocrinology, 34, 293-298.

ImPERATo-McGINLEY, J., PETERSON, R. E., & GAUTIER, T. (1976). Gender identity and hermaphroditism [Letter]. Science, 191, 872.

ImPERATo-McGINLEY, J., PETERSON, R. E., GAUTIER, T., & STURLA, E. (1979). Androgens and the evolution of male-gender identity among male pseudohermaphrodites with 5(alphareductase deficiency. New England Journal of Medicine, 300, 1233-1237.

IMPERATO-McGINLEY, J., PETERSON, R. E., LESHIN, M., GRIFFIN, J. E., COOPER, G., DRAGHI, S., BERENYI, M., & WILSON, J. D. (1980). Steroid 5a-reductase deficiency in a 65year-old male pseudohermaphrodite: The natural history, ultrastructure of the testes, and evidence for inherited enzyme heterogeneity. Journal of Clinical Endocrinology and Metabolism, 50, 15-22.

ImPERATo-McGINLEY, J., PETERSON, R. E., STOLLER, R., & GOODWIN, W. E. (1979). Male pseudohermaphroditism secondary to 17]-hydroxysteroid dehydrogenase deficiency: Gender role change with puberty. Journal of Clinical Endocrinology and Metabolism, 49, 391-395.

IMPERATo-McGINLEY, J., PICHARDO, M., GAUTIER, T., VOYER, D., & BRYDEN, M. P. (1991). Cognitive abilities in androgen-insensitive subjects: Comparison with control males and females from the same kindred. Clinical Endocrinology, 34, 341-347.

INTERSEX SOCIETY OF NORTH AMERICA. (1999, October 25). Columbia high court restricts intersex genital mutilation. (Available on the World Wide Web http://www.isna.org)

INTONS-PETERSON, M. J., & REDDEL, M. (1984). What do people ask about a neonate? Developmental Psychology, 20, 358-359.

IVARSSON, S.-A., NIELSEN, M. D., & LINDBERG, T. (1988). Male pseudohermaphroditism due to 5(x-reductase deficiency in a Swedish family. European Journal of Pediatrics, 147, 532-535.

IZQUIERDO, G., & GLASSBERG, K. 1. (1993). Gender assignment and gender identity in patients with ambiguous genitalia. Urology, 42, 232-242.

JOHNSTON, W. G., YEATmAN, G. W., & WEIGEL, J. W. (1977). Congenital absence of the penis. Journal of Urology, 117, 508-511.

JONES, H. W., PARK, I. J., & ROCK, J. A. (1978). Technique of surgical sex reassignment for micropenis and allied conditions. American Journal of Obstetrics and Gynecology, 132, 870-877.

JURASKA, J. M. (1998). Neural plasticity and the development of sex differences. Annual Review of Sex Research, 9, 20-38.

KANDEMIR, N., & YORDAM, N. (1997). Congenital adrenal hyperplasia in Turkey: A review of 273 patients. Acta Paediatrica, 86, 22-25.

KATz, M. D., KLiGmAN, I., CAi, L., ZHU, Y, FRATIANNI, C. M., ZERVOUDAKIS, I., RoSENWAKS, Z., & IMPERATo-McGINLEY, J. (1997). Paternity by intrauterine insemination with sperm from a man with 5a-reductase-2 deficiency. New England Journal of Medicine, 336, 994-997.

KESSLER, S. J. (1990). The medical construction of gender: Case management of intersexed infants. Signs, 16, 3-26.

KESSLER, S. J. (1998). Lessons from the intersexed. New Brunswick, NJ: Rutgers University Press.

KESSLER, W 0.@ & MCLAUGHLIN, A. P. (1973). Agenesis of penis: Embryology and management. Urology, 1, 226-229.

KING, C. D. (1945). The meaning of normal. Yale Journal of Biology and Medicine, 17, 493-501.

KING, M. (1998, April 4-10). NZ sex-change guru in US ethics storm: John Money's only interview. New Zealand Listener, 18-21.

KIPNIS, K., & DIAMOND, M. (1998). Pediatric ethics and the surgical assignment of sex. Journal of Clinical Ethics, 9, 398-410.

KIRSHBAUM, J. D. (1950). Congenital absence of the external genitals (persistent primitive cloaca). Journal of Pediatrics, 37, 102-105.

KLEBS, T. A. E. (1876). Handbuch der pathologischen Anatomie [Handbook of pathological anatomy]. Berlin: A. Hirschwald.

KODMAN-JONES, C., & ZDERIC, S. (1999, April). The management of cloacal exstrophy: The CHOP experience. Paper presented at the Conference on Pediatric Gender Reassignment: A Critical Reappraisal, Dallas, TX.

KOHLBERG, L. (1966). A cognitive-developmental analysis of children's sex-role concepts and attitudes. In E. E. Maccoby (Ed.), The development of sex differences (pp. 82-173). Stanford, CA: Stanford University Press.

KUHNLE, U., & BULLINGER, M. (1997). Outcome of congenital adrenal hyperplasia. Pediatric Surgery International, 12, 511-515.

KUTTENN, F., MOWSZOWICA, I., WRIGHT, F., BAUDOT, N., JAFFIOL, C., ROBIN, M., & MAUvAis-JARvis, P. (1979). Male pseudohermaphroditism: A comparative study of one patient with 5(x-reductase deficiency and three patients with the complete form of testicular feminization. Journal of Clinical Endocrinology and Metabolism, 49, 861-865.

LEE, E. H. (1994). Producing sex: An interdisciplinary perspective on sex assignment decisions for intersexuals. Unpublished B.A. thesis, Brown University, Providence, Rl. LEE, P. A., MAZUR, T., DANisH, R., AmRHEIN, J., BLIZZARD, R. M., MONEY, J., & MIGEON,

C. J. (1980). Micropenis. I. Criteria, etiologies and classification. Johns Hopkins Medical Journal, 146, 156-163.

LIEBERSON, S., & BELL, E. 0. (1992). Children's first names: An empirical study of social taste. American Journal of Sociology, 98, 511-554.

LisA, L., HANAK, J., CERNY, M., FAFLOVA, H., KAFKA, V., & BRAzA, J. (1973). Agenesis of the penis. Journal of Pediatric Surgery, 8, 327-328.

MADDEN, J. D., WALSH, P. C., MAcDONALD, R C., & WILSON, J. D. (1975). Clinical and endocrinologic characterization of a patient with the syndrome of incomplete testicular feminization. Journal of Clinical Endocrinology and Metabolism, 41, 751-760.

MAES, M., SULTAN, C., ZERHOuNi, N., ROTHWELL, S. W., & MIGEON, C. J. (1979). Role of testosterone binding to the androgen receptor in male sexual differentiation of patients with 5a-reductase deficiency. Journal of Steroid Biochemistry, 11, 1385-1390.

MARK, E., CASTELLS, S., GLAsSBERG, K., CHOI, S. J., TOLETE-VELCEK, F., DAVID, K., & MIGEON, C. J. (1983). Deficiency of androgen receptors in male pseudohermaphroditism. Urology, 21, 168-171.

MASICA, D. N., MONEY, J., & EHRHARDT, A. A. (1971). Fetal feminization and female gender identity in the testicular feminizing syndrome of androgen insensitivity. Archives of Sexual Behavior, 1, 131-142.

MAUVAIS-JARVIS, P., KUTTENN, F., Mowszowicz, I., & WRIGHT, F. (1981). Different aspects of 5a-reductase deficiency in male pseudohermaphroditism and hypothyroidism. Clinical Endocrinology, 14, 459-469.

MCCAULEY, E. (1990). Disorders of sexual differentiation and development: Psychological aspects. Pediatric Clinics of North America, 37, 1405-1420.

MCCONAGHY, M. (1979). Gender permanence and the genital basis of gender: Stages in the development of constancy of gender identity. Child Development, 50, 1223-1226. McDERMID, S. A., ZucKER, K. J., BRADLEY, S. J., & MAING, D. M. (1998). Effects of phys

ical appearance on masculine trait ratings of boys and girls with gender@dentity disorder. Archives of Sexual Behavior, 27, 253-267.

MEDINA, M., CHAVEz, B., & PEREZ-PALACIOS, G. (1981). Defective androgen action at the cellular level in the androgen resistance syndromes. 1. Differences between the complete and incomplete testicular feminization syndromes. Journal of Clinical Endocrinology and Metabolism, 53, 1243-1246.

MENDEZ, J. P., ULLOA-AGUIRRE, A., IMPERATO-McGINLEY, J., BRUGMANN, A., DELFIN, M., CHAVEZ, B., SHACKLETON, C., KOFMAN-ALFARO, S., & PEREZ-PALAcios, G. (1995). Male pseudohermaphroditism due to primary 5u.-reductase deficiency: Variation in gender identity reversal in seven Mexican patients from five different pedigrees. Journal of Endocrinological Investigation, 18, 205-213.

MENDONCA, B. B., BLOISE, W, ARNHOLD, 1. J. P., BATisTA, M. C., DE ALMEIDA TOLEDO, S. P., DRUMMOND, M. C. F., NICOLAU, W, & MATTAR, E. (1987). Male pseudohermaphroditism due to nonsalt-losing 3@-hydroxysteroid dehydrogenase deficiency: Gender role change and absence of gynecomastia at puberty. Journal of Steroid Biochemistry, 28, 669-675.

MENDONCA, B. B., INACIO, M., COSTA, E. M. F., ARNHOLD, 1. J. R, SILVA, F. A. Q., NicoLAu, W., BLOISE, W., RUSSELL, D. W., & WILSON, J. D. (1996). Male pseudohermaphroditism due to steroid 5a-reductase 2 deficiency Medicine, 75, 64-76.

MESROBIAN, H. J., KELALIS, P. P., & KRAMER, S. A. (1986). Long-term followup of Cosmetic appearance and genital function in boys with exstrophy: Review of 53 patients. Journal of Urology, 136, 256-258.

MEYER-BAHLBURG, H. F. L. (1982). Hormones and psychosexual differentiation: Implications for the management of intersexuality, homosexuality and transsexuality. Clinics in Endocrinology and Metabolism, 11, 681-701.

MEYER-BAHLBURG, H. F. L. (1993). Gender identity development in intersex patients. Child and Adolescent Psychiatric Clinics of North America, 2, 501-512.

MEYER-BAHLBURG, H. F. L. (1994). Intersexuality and the diagnosis of gender identity disorder. Archives of Sexual Behavior, 23, 21-40.

MEYER-BAHLBURG, H. F. L. (1998). Gender assignment in intersexuality. Journal of Psychology & Human Sexuality, 10(2), 1-21.

MEYER-BAHLBURG, H. F. L. (1999a). Gender assignment and reassignment in 46,XY pseudohermaphroditism and related conditions. Journal of Clinical Endocrinology & Metabolism, 84, 3455-3458.

MEYER-BAHLBURG, H. F. L. (1999b). Variants of gender differentiation. In H.-C. Steinhausen & F. C. Verhulst (Eds.), Risks and outcomes in developmental psychopathology (pp. 298-313). New York: Oxford University Press.

MEYER-BAHLBURG, H. F. L. (1999c). What causes low rates of child-bearing in congenital adrenal hyperplasia? Journal of Clinical Endocrinology and Metabolism, 84, 18441847.

MEYER-BAHLBURG, H. F. L., DOLEZAL, C., BAKER, S. W, CARLSON, A., OBEID, J., VOGIATZI, M., & NEW, M. 1. (1999, June). Differences in behavioral masculinization between subtypes of classical congenital adrenal hyperplasia in girls. Poster presented at the meeting of the International Academy of Sex Research, Stony Brook, New York.

MEYER-BAHLBURG, H. F. L., EHRHARDT, A. A., PINEL, A., & GRUEN, R. (1989, October). Gender identity development in two 46,XY individuals with normal gonads raised female because of severe genital abnormalities. Paper presented at the meeting of the Society for Sex Therapy and Research, Cleveland, OH.

MEYER-BAHLBURG, H. F. L., GRUEN, R. S., NEW, M. I., BELL, J. J., MORISHIMA, A., SHIMSHI, M., BUENO, Y., VARGAS, I., & BAKER, S. W (1996). Gender change from female to male in classical congenital adrenal hyperplasia. Hormones and Behavior, 30, 319-332.

MIGEON, C. J., & WISNIEWSKi, A. B. (1998). Sexual differentiation: From genes to gender. Hormone Research, 50, 245-251.

MILEWICH, L., MENDONcA, B. B., ARNHOLD, I., WALLACE, A. M., DONALDSON, A D. C., WILSON, J. D., & RUSSELL, D. W. (1995). Women with steroid 5a-reductase 2 deficiency have normal concentrations of plasma 5u,-dihydroprogesterone during luteal phase. Journal of Clinical Endrocrinology and Metabolism, 80, 3136-3139.

MITCHELL, M. (1999, April). The challenge of cloacal exstrophy. Paper presented at the Conference on Pediatric Gender Reassignment: A Critical Reappraisal, Dallas, TX.

MONEY, J. W. (1952). Hermaphroditism: An inquiry into the nature of a human paradox. Unpublished doctoral dissertation, Harvard University.

MONEY, J. (1955). Hermaphroditism, gender and precocity in hyperadrenocorticism: Psychologic findings. Bulletin of the Johns Hopkins Hospital, 96, 253-264.

MONEY, J. (1965). Psychologic evaluation of the child with intersex problems. Pediatrics, 36, 51-55.

MONEY, J. (1975). Ablatio penis: Normal male infant sex-reassigned as a girl. Archives of Sexual Behavior, 4, 65-71.

MONEY, J. (1976). Gender identity and hermaphroditism [Letter]. Science, 191, 872. MONEY, J. (1979). [Letter to the editor]. Obstetrical and Gynecological Survey, 34, 770771.

MONEY, J. (1984a): Part I: Childhood coping and adult follow-up of micropenis syndrome in a case with female sex assignment. International Journal of Family Psychiatry, 5,317-339.

MONEY, J. (1984b). Part II: Transsexual versus homosexual coping in micropenis syndrome with male sex assignment. International Journal of Family Psychiatry, 5, 341-373. MONEY, J. (1985). The conceptual neutering of gender and the criminalization of

sex. Archives of Sexual Behavior, 14, 279-290.

MONEY, J. (1991). Biographies of gender and hermaphroditism in paired comparisons. Amsterdam: Elsevier.

MONEY, J. (1994). Sex errors of the body and related syndromes: A guide to counseling children, adolescents, and their families (2nd ed.). Baltimore, MD: Paul H. Brookes Publishing Co.

MONEY, J. (1998). Case consultation: Ablatio penis. Medicine and the Law, 17, 113-123. MONEY, J., & ANNECILLO, C. (1987). Crucial period effect in psychoendocrinology: Two

syndromes, abuse dwarfism and female (CVAH) hermaphroditism. In A H. Bornstein (Ed.), Sensitive periods in development: Interdisciplinary perspectives (pp. 145-158). Hillsdale, NJ: Erlbaum.

MONEY, J., & DALERY, J. (1976). latrogenic homosexuality: Gender identity in seven 46,XX chromosomal females with hyperadrenocortical hermaphroditism born with a penis, three reared as boys, four reared as girls. Journal of Homosexuality, 1, 357-371.

MONEY, J., DEVORE, H., & NORMAN, B. F. (1986). Gender identity and gender transposition: Longitudinal outcome study of 32 male hermaphrodites assigned as girls. Journal of Sex & Marital Therapy, 12, 165-181.

MONEY, J., & EHRHARDT, A. A. (1972). Man and woman, boy and girl: The differentiation and dimorphism of gender identity from conception to maturity. Baltimore, MD: Johns Hopkins Press.

MONEY, J., EHRHARDT, A. A., & MASICA, D. N. (1968). Fetal feminization induced by androgen insensitivity in the testicular feminizing syndrome: Effect on marriage and

maternalism. Johns Hopkins Medical Journal, 123, 105-114.

MONEY, J., HAmpsON, J. G., & HAmPSON, J. L. (1955). An examination of some basic sexual concepts: The evidence of human hermaphroditism. Bulletin of the Johns Hopkins Hospital, 97, 301-319.

MONEY, J., HAMPSON, J. G., & HAmPSON, J. L. (1957). Imprinting and the establishment of gender role. Archives of Neurology and Psychiatry, 77, 333-336.

MONEY, J., LEHNE, G. K, & PIERRE-JEROME, F. (1985). Micropenis: Gender, erotosexual coping strategy, and behavioral health in nine pediatric cases followed to adulthood. Comprehensive Psychiatry, 26, 29-42.

MONEY, J., & MAZUR, T. (1977). Microphallus: The successful use of a prosthetic phallus in a 9-year-old boy. Journal of Sex and Marital Therapy, 3, 187-196.

MONEY, J., MAZUR, T., ABRAms, C., & NoRmAN, B. F. (1981). Micropenis, family mental health, and neonatal management: A report on 14 patients reared as girls. Journal of Preventive Psychiatry, 1, 17-27.

MONEY, J., & NORMAN, B. F. (1988). Pedagogical handicap associated with micropenis and other CHARGE syndrome anomalies of embryogenesis: Four 46,XY cases reared as girls. American Journal of Psychotherapy, 42, 354-379.

MONEY, J., & OGUNRO, C. (1974). Behavioral sexology: Ten cases of genetic male intersexuality with impaired prenatal and pubertal androgenization. Archives of Sexual Behavior, 3, 181-205.

MONTAGNINO, B., CZYZEwsKi, D. I., RUNYAN, R. D., BERKMAN, S. ROTH, D. R., & GONZALEs, E. T. (1998). Long-term adjustment issues in patients with exstrophy. Journal of Urology, 160, 1471-1474.

MOORE, K. L., & BARR, M. L. (1955). Smears from the oral mucosa in the detection of chromosomal sex. Lancet, 2, 57-58.

MoRRIS, J. M. (1953). The syndrome of testicular feminization in male pseudohermaphrodites. American Journal of Obstetrics & Gynecology, 65, 1192-1211.

MoRRIs, J. M., & MAHEsH, V. B. (1963). Further observations on the syndrome, "testicular feminization." American Journal of Obstetrics & Gynecology, 87, 731-748.

MULAIKAL, R. M., MIGEON, C. J., & ROCK, J. A. (1987). Fertility rates in female patients with congenital adrenal hyperplasia due to 21-hydroxylase deficiency. New England Journal of Medicine, 316, 178-182.

MUREAU, M. A. M., SLIJPER, F. M. E., NIJMAN, R. J. M., VAN DER MEULEN, J. C., VERHuLsT, F. C., & SLOB, A. K. (1995). Psychosexual adjustment of children and adolescents after different types of hypospadias surgery: A norm-related study. Journal of Urology, 154, 1902-1907.

MUREAU, M. A. M., SLIJPER, F. M. E., SLOB, A. K., & VERHULST, F. C. (1995). Genital perception of children, adolescents, and adults operated on for hypospadias: A comparative study. The Journal of Sex Research, 32, 289-298.

MuREAu, M. A. M., SLIJPER, F. M. E., SLOB, A. K., & VERHULsT, F. C. (1997). Psychosocial functioning of children, adolescents, and adults following hypospadias surgery: A comparative study. Journal of Pediatric Psychology, 22, 371-387.

MUREAU, M. A. M., SLIJPER, F. M. E., SLOB, A. K., VERHULsT, F. C., & NuMAN, R. J. M. (1996). Satisfaction with penile appearance after hypospadias surgery: The patient and surgeon view. Journal of Urology, 155, 703-706.

NEw, M. L, GHIZZONI, L., & SPEISER, R W. (1996). Update on congenital adrenal hyperplasia. In F. Lifshitz (Ed.), Pediatric endocrinology (3rd ed., pp. 305-320). New York: Marcel Dekker.

NEw, M. I., & Josso, N. (1988). Disorders of gonadal differentiation and congenital adrenal hyperplasia. Endocrinology and Metabolism Clinics of North America, 17, 339366.

NG, W K., TAYLOR, N. F., HUGHES, 1. A., TAYLOR, J., RANsLEY, P. G., & GRANT, D. B. (1990). 5a-reductase deficiency without hypospadias. Archives of Disease in Childhood, 65,1166-1167.

NORDENSKJOLD, A., MAGNUS, 0., AAGENAES, 0., & KNUDTZON, J. (1998). Homozygous mutation (A228T) in the 5a-reductase type 2 gene in a boy with 5a-reductase deficiency: Genotype-phenotype correlations. American Journal of Medical Genetics, 80, 269-272.

NowAKowSKi, H., & LENZ, W. (1961). Genetic aspects in male hypogonadism. Recent Progress in Hormone Research, 17, 53-95.

OCHOA, B. (1998). Trauma of the external genitalia in children: Amputation of the penis and emasculation. Journal of Urology, 160, 1116-1119.

ODAME, I., DONALDSON, M. D. C., WALLACE, A. M., COCHRAN, W., & SMITH, P. J. (1992). Early diagnosis and management of 5(x-reductase deficiency. Archives of Disease in Childhood, 67, 720-723.

OESCH, I. L., PINTER, A., & RANSLEY, P. G. (1987). Penile agenesis: A report of six cases. Journal of Pediatric Surgery, 22, 172-174.

OKON, E., LivNi, N., ROSLER, A., YORKONI, S., SEGAL, S., KOHN, G., & SCHENKER, J. G. (1980). Male pseudohermaphroditism due to 5cc-reductase deficiency: Ultrastructure of the gonads. Archives of Pathology and Laboratory Medicine, 104, 363-367.

O'NEILL, J. A., HOLCOMB, G. W., SCHNAUFER, L., TEMPLETON, J. M., BISHOP, H. C., Ross, A. J., DUCKETT, J. W., NORWOOD, W. I., ZIEGLER, M. M., & Koop, C. E. (1988). Surgical experience with thirteen conjoined twins. Annals of Surgery, 208, 299-310.

OPITZ, J. M., SIMPSON, J. L., SARTO, G. E., SUMMITT, R. L., NEW, M., & GERMAN, J.

(1972). Pseudovaginal perineoscrotal hypospadias. Clinical Genetics, 3, 1-26.

PAGON, R. A. (1987). Diagnostic approach to the newborn with ambiguous genitalia. Pediatric Clinics of North America, 34, 1019-1031.

PANG, S. (1997). Congenital adrenal hyperplasia. Endocrinology and Metabolism Clinics of North America, 26, 853-891.

PAOLETTi, J. B., & THOMPSON, S. (1987, May). Gender differences in infants' rompers, 1910-1930. Paper presented at the meeting of the Costume Society of America, Richmond, VA.

PERLMUTTER, A. D. (1994). [Editorial comment]. Urology, 43, 374.

PEREZ-PALACIOS, G., BRUGMANN, A., IMPERATO-McGINLEY, J., BASSOL, S., VALDES, E., MENDEZ, J. P., & ULLOA-AGUIRRE, A. (1984). Variability on the gender identity in the syndrome of 5(x-steroid reductase deficiency. Prostate, 5, 355.

PEREZ-PALACios, G., CHAvEz, B., MENDEZ, J. R, IMPERATo-McGINLEY, J., & ULLOAAGUIRRE, A. (1987). The syndromes of androgen resistance revisited. Journal of Steroid Biochemistry, 27, 1101-1108.

POHLANDT, F., KUHN, H., TELLER, W., & THomA, H. (1974). Penisagenesie. Weibliche geschlechtszuweisung unter psychotherapeutischer Betreuung der Eltern [Penile agenesis. Female sex reassignment and psychotherapeutic management of parents]. Deutsche Medizinische Wochenschrift, 99, 2166-2172.

PREVES, S. E. (1998). For the sake of the children: Destigmatizing intersexuality. Journal of Clinical Ethics, 9, 411-420.

PRICE, P., WASS, J. A. H., GRIFFIN, J. E., LESHIN, M., SAVAGE, M. 0., LARGE, D. M., BU'LOCK, D. E., ANDERSON, D. C., WILSON, J. D., & BESSER, G. M. (1984). High dose androgen therapy in male pseudohermaphroditism due to 5oc-reductase deficiency and disorders of the androgen receptor. Journal of Clinical Investigation, 74, 1496-1508.

QUATTRIN, T., ARONICA, S., & MAZUR, T. (1990). Management of male pseudohermaphroditism: A case report spanning twenty-one years. Journal of Pediatric Psychology, 15, 699709.

QUIGLEY, C. A., DE BELLIS, A., MARSCHKE, K. B., EL-AWADY, M. K., WILSON, E. M., & FRENCH, F. S. (1995). Androgen receptor defects: Historical, clinical, and molecular perspectives. Endocrine Reviews, 16, 271-321.

RADMAYR, C., CULIG, Z., GLATZL, J., NEUSCHMID-KAsPAR, F., BARTSCH, G., & KLocKER, H. (1997). Androgen receptor point mutations as the underlying molecular defect in 2 patients with androgen insensitivity syndrome. Journal of Urology, 158, 1553-1556.

RANDOLPH, J., HUNG, W., & RATHLEV, M. C. (1981). Clitoroplasty for females born with

ambiguous genitalia: A long-term study of 37 patients. Journal of Pediatric Surgery, 16, 882-887.

REILLY, J. M., & WOODHOUSE, C. R. J. (1989). Small penis and the male sexual role. Journal of Urology, 142, 569-571.

REINER, W. G. (1997a). [Letter to the editor]. Archives of Pediatrics and Adolescent Medicine, 151, 1064.

REINER, W. G. (1997b). Sex assignment in the neonate with intersex or inadequate genitalia. Archives of Pediatrics and Adolescent Medicine, 151, 1044-1045.

REINER, W (1997c). To be male or female-that is the question [Editorial]. Archives of Pediatrics and Adolescent Medicine, 151, 224-225.

REINER, W G., GEARHART, J. P, & JEFFS, R. (1999). Psychosexual dysfunction in males with genital anomalies: Late adolescence, Tanner stages IV to VI. Journal of the American Academy of Child and Adolescent Psychiatry, 38, 865-872.

REKERS, G. A., & YATES, C. E. (1976). Sex-typed play in feminoid boys vs. normal boys and girls. Journal of Abnormal Child Psychology, 4, 1-8.

RicHART, R., & BENIRSCHKE, K. (1960). Penile agenesis: Report of case, review of world literature, and discussion of pertinent embryology. Archives of Pathology, 70, 252-260. RicKHAm, P. R (1960). Vesico-intestinal fissure. Archives of Disease in Childhood, 35, 97-102.

ROBERTS, S. (1996, March 3). Forget Jos6. There's an America of Kevins and Ashleys aborning. New York Times, p. 3 (Section 4).

ROSENBLUM, R. R., & TURNER, W. R. (1973). Congenital absence of the penis. Journal of the South Carolina Medical Association, 69, 178-179.

RoSLER, A., & KOHN, G. (1983). Male pseudohermaphroditism due to 17-hydroxysteroid dehydrogenase deficiency: Studies on the natural history of the defect and effect of androgens on gender role. Journal of Steroid Biochemistry, 19, 663-674.

RUBLE, D. N., & MARTIN, C. L. (1998). Gender development. In W. Damon (Ser. ed.) & N. Eisenberg (Vol. ed.), The handbook of child psychology (5th ed.). Vol. 3: Social, emotional, and personality development (pp. 933-1016). New York: Wiley.

RUSSELL, D. W., & WILSON, J. D. (1994). Steroid 5(x-reductase: Two genes/two enzymes. Annual Review of Biochemistry, 63, 25-61.

RUTGERS, J. L., & ScuLLY, R. E. (1991). The androgen insensitivity syndrome (testicular feminization): A clinicopathologic study of 43 cases. International Journal of Gynecological Pathology, 10, 126-144.

SAENGER, P., GOLDMAN, A. S., LEVINE, L. S., KORTH-SCHUTZ, S., MUECKE, E. C., KATsumATA, M., DOBERNE, Y, & NEw, M. 1. (1978). Prepubertal diagnosis of steroid 5a-reductase deficiency. Journal of Clinical Endocrinology and Metabolism, 46, 627-634.

SANDBERG, D. E., MEYER-BAHLBURG, H. F. L., ARANOFF, G. S., SCONZO, J. M., & HENSLE, T. W (1989). Boys with hypospadias: A survey of behavioral difficulties. Journal of Pediatric Psychology, 14, 491-514.

SANDBERG, D. E., MEYER-BAHLBURG, H. F. L., YAGER, T. J., HENSLE, T. W., LEVITT, S. B., KOGAN, S. J., & REDA, E. F. (1995). Gender development in boys born with hypospadias. Psychoneuroendocrinology, 20, 693-709.

SAVAGE, M. 0., PREECE, M. A., JEFFCOATE, S. L., RANSLEY, P. G., RumSBY, G., MANSFIELD, M. D., & WILLIAMS, D. 1. (1980). Familial male pseudohermaphroditism due to deficiency of 5a-reductase. Clinical Endocrinology, 12, 397-406.

SCHOBER, J. M. (1998a). A surgeon's response to the intersex controversy. Journal of Clinical Ethics, 9, 393-397.

SCHOBER, J. M. (1998b). Early feminizing genitoplasty or watchful waiting. Journal of Pediatric and Adolescent Gynecology, 11, 154-156.

SCHOBER, J. M. (1999). Long-term outcomes and changing attitudes to intersexuality. BJU International, 83(Suppl. 3), 39-50.

SCHONFELD, W A. (1943). Primary and secondary sexual characteristics: Study of their development in males from birth through maturity, with biometric study of penis and

testes. American Journal of Diseases of Childhood, 65, 535-549.

ScHwARz, H. P. (1997). Sex reassignment at birth [Letter to the editor]. Archives of Pediatrics and Adolescent Medicine, 151, 1064.

SHAKIN, M., SHAKIN, D., & STERNGLANZ, S. H. (1985). Infant clothing: Sex labeling for strangers. Sex Roles, 12, 955-964.

SHEARMAN, R. R (1982). Intersexuality. In R J. V. Beumont & G. D. Burrows (Eds.), Handbook ofpsychiatry and endocrinology (pp. 325-354). Amsterdam: Elsevier Biomedical Press.

SIMPSON, J. L., NEW, M., PETERSON, R. E., & GERMAN, J. (1971). Pseudovaginal perineoscrotal hypospadias (PPSH) in sibs. Birth Defects, 7, 140-144.

SINNECKER, G. H. G., HIORT, 0., DIBBFLT, L., ALBERS, N., DORR, H. G., HAus, H., HEINRICH, U., HEMMINGHAUS, M., HOEPFFNER, W., HOLDER, M., SCHNABEL, D., & KRUSE, K. (1996). Phenotypic classification of male pseudohermaphroditism due to steroid 5(X-reduetase 2 deficiency. American Journal of Medical Genetics, 63, 223-230.

SIZONENKO, P. C. (1993). Sexual differentiation. In J. Bertrand, R. Rappaport, & P. C. Sizonenko (Eds.), Pediatric endocrinology: Physiology, pathophysiology, clinical aspects (2nd ed., pp. 88-99). Baltimore: Williams & Wilkins.

SKOOG, S. J., & BELMAN, A. B. (1989). Aphallia: Its classification and management. Journal of Urology, 141, 589-592.

SLIJPER, F. M. E. (1984). Androgens and gender role behaviour in girls with congenital adrenal hyperplasia (CAH). Progress in Brain Research, 61, 417-422.

SLIJPER, F. M. E., DROP, S. L. S., MOLENAAR, J. C., & DF MUINCK KEIZER-SCHRAmA, S. M. P. F. (1998). Long-term psychological evaluation of intersex children. Archives of Sexual Behavior, 27, 125-144.

SLOB, A. K., VAN DER WERFF TEN BOSCH, J. J., VAN HALL, E. V., DE JONG, F. H., WEIJMAR SCHULTZ, W C. M., & EIKELBOOM, F. A. (1993). Psychosexual functioning in women with complete testicular feminization: Is androgen replacement therapy preferable to estrogen? Journal of Sex & Marital Therapy, 19, 201-209.

SMITH, D. W. (1977). Micropenis and its management. Birth Defects, 13, 147-154. SRIPATHI, V., AHMED, S., SAKATi, N., & AL-ASHWAL, A. (1997). Gender reversal in 46XX congenital virilizing adrenal hyperplasia. British Journal of Urology, 79, 785-789.

SROUFE, L. A. (1990). Considering normal and abnormal together: The essence of developmental psychopathology. Development and Psychopathology, 2, 335-347.

STEIN, R., STOCKLE, M., FISCH, M., NAKAI, H., MULLER, S. C., & HOHENFELLNER, R.

(1994). The fate of the adult exstrophy patient. Journal of Urology, 152, 1413-1416. STOLAR, C. J. H., WIENER, E. S., HENSLE, T. W., SILEN, M. L., SUKAROCHANA, K., SIEBER,

W. K., GOLDSTEIN, H. R., & PETTIT, J. (1987). Reconstruction of penile agenesis by a posterior sagittal approach. Journal of Pediatric Surgery, 22, 1076-1080.

STOLLER, R. J. (1964). The hermaphroditic identity of hermaphrodites. Journal of Nervous and Mental Disease, 139, 453-457.

STOLLER, R. J. (1965). The sense of maleness. Psychoanalytic Quarterly, 34, 207-218. STOLLER, R. J., GARFINKEL, H., & ROSEN, A. C. (1962). Psychiatric management of intersexed patients. California Medicine, 96, 30-34.

SwAAB, D. F., GOOREN, L. J. G., & HOFMAN, M. A. (1992). Gender and sexual orientation in relation to hypothalamic structures. Hormone Research, 38(Suppl. 2), 51-61.

TAHA, S. A. (1994). Male pseudohermaphroditism: Factors determining the gender of rearing in Saudi Arabia. Urology, 43, 370-374.

TALwAp, S., & KAPOOR, R. (1988). Aphallia. Indian Pediatrics, 25, 579-581.

TANK, E. S., & LINDENAUER, S. M. (1970). Principles of management of exstrophy of the cloaca. American Journal of Surgery, 119, 95-98.

TETER, J., & BocZKOWSKI, K. (1966). Testicular feminization with and without clitoral enlargement. American Journal of Obstetrics & Gynecology, 94, 813-819.

THIGPEN, A. E., DAvis, D. L., MILATOVICH, A., MENDONCA, B. B., IMPERATo-McGINLEY, J., GRIFFIN, J. E., FRANcKE, U., WILSON, J. D., & RUSSELL, D. W (1992). Molecular genetics of

steroid 5(x-reductase 2 deficiency. Journal of Clinical Investigation, 90, 799-809.

THORNE, B. (1993). Gender play: Girls and boys in school. New Brunswick, NJ: Rutgers University Press.

TUFFIER, T., & LAPOINTE, A. (1911). L'hermaphrodisme: Ses varietes et ses consequences pour la pratique medicale (d'apres un cas personnel). Revue de Gyn6coloigie et de Chiurgie Abdominale, 17, 209-268.

VAN SETERS, A. R, & SLOB, A. K. (1988). Mutually gratifying heterosexual relationship with micropenis of husband. Journal of Sex & Marital Therapy, 14, 98-107.

VAN WYK, J. J., & CALIKOGLIJ, A. A. (1999). Should boys with micropenis be reared as girls? Journal of Pediatrics, 134, 537-538.

VILAIN, E., & MCCABE, E. R. B. (1998). Mammalian sex determination: From gonads to brain. Molecular Genetics and Metabolism, 65, 74-84.

WALLEN, K. (1996). Nature needs nurture: The interaction of hormonal and social influences on the development of behavioral sex differences in rhesus monkeys. Hormones and Behavior, 30, 364-378.

WALSH, P. C., MADDEN, J. D., HARROD, M. J., GOLDSTEIN, J. L., MACDONALD, P. C., & WILSON, J. D. (1974). Familial incomplete male pseudohermaphroditism, type 2: Decreased dihydrotestosterone formation in pseudovaginal perineoscrotal hypospadias. New England Journal of Medicine, 291, 944-949.

WARNE, G. L., & ZAjAc, J. D. (1998). Disorders of sexual differentiation. Endocrinology and Metabolism Clinics of North America, 27, 945-967.

WEBSTER'S SEVENTH NEW COLLEGIATE DICTIONARY. (1963). Springfield, MA: G. & C. Merriam.

WEDELL, A. (1998). Molecular genetics of congenital adrenal hyperplasia (21-hydroxylase deficiency): Implications for diagnosis, prognosis and treatment. Acta Paediatrica, 87, 159-164.

WIEACKER, P, FLECKEN, U., & BRECKWOLDT, M. (1992). Ein Fall von pseudovaginaler, perineoskrotaler Hypospadie mit 5a-reduktase-defizienz [A case of pseudovaginal, perineoscrotal hypospadia with 5(x-reductase deficiency]. Geburtshilfe und Frauenheilkunde, 52, 126-128.

WILKINS, L., LEWIS, R. A., KLEIN, R., & RoSEMBERG, E. (1950). The suppression of androgen secretion by cortisone in a case of congenital adrenal hyperplasia. Bulletin of the Johns Hopkins Hospital, 86, 249-252.

WILSON, B. E., & REINER, W. G. (1998). Management of intersex: A shifting paradigm. Journal of Clinical Ethics, 9, 360-369.

WILSON, J. D., GRIFFIN, J. E., & RusSELL, D. W. (1993). Steroid 5a-reductase 2 deficiency. Endocrine Reviews, 14, 577-593.

WILSON, R. C., MERCADO, A. B., CHENG, K. C., & NEW, M. 1. (1995). Steroid 21-hydroxylase deficiency: Genotype may not predict phenotype. Journal of Clinical Endocrinology and Metabolism, 80, 2322-2329.

WINESTINE, M. C. (1989). To know or not to know: Some observations on women's reactions to the availability of prenatal knowledge of their babies' sex. Journal of the American Psychoanalytic Association, 37, 1015-1030.

WOOLLETT, A., WHITE, D., & LYON, L. (1982). Observations of fathers at birth. In N. Beail & J. McGuire (Eds.), Fathers: Psychological perspectives (pp. 71-91). London: Junction Books.

YOUNG, H. H., CocKETT, A. T. K., STOLLER, R., ASHLEY, F. L., & GOODWIN, W. E. (1971). The management of agenesis of the phallus. Pediatrics, 47, 81-87.

ZACHARIN, M. (1999). Fertility and its complications in a patient with salt losing congenital adrenal hyperplasia. Journal of Pediatric Endocrinology & Metabolism, 12, 89-94. ZDERIC, S. A., CANNING, D. A., SNYDER, H. M., & CARR, M. C. (Eds.). (in press). Pediatric

gender assignment: A critical reappraisal. New York: Plenum.

ZUCKER, K. J. (1996). Commentary on Diamond's "Prenatal Predisposition and the Clinical Management of Some Pediatric Conditions." Journal of Sex & Marital Therapy,

22,148-160.

ZUCKER, K. J., & BRADLEY, S. J. (1995). Gender identity disorder and psychosexual prob

lems in children and adolescents. New York: Guilford Press.

ZUCKER, K J., BRADLEY, S. J., Kums, M., PECORE, K., BIRKENFELD-ADAms, A., DOERING, R. W., MITCHELL, J. N., & WILD, J. (in press). Gender constancy judgments in children with gender identity disorder: Evidence for a developmental lag. Archives of Sexual Behavior.

ZUCKER, K J., BRADLEY, S. J., LowRy SULLIVAN, C. B., KuKsis, M., BIRKENFELD-ADAMS, A., & MITCHELL, J. N. (1993). A gender identity interview for children. Journal of Personality Assessment, 61, 443-456.

ZUCKER, K. J., BRADLEY, S. J., OLIVER, G., BLAKE, J., FLEMING, S., & HoOD, J. (1996). Psychosexual development of women with congenital adrenal hyperplasia. Hormones and Behavior, 30, 300-318.

ZUCKER, K. J., DOERING, R. W., BRADLEY, S. J., & FINEGAN, J. K. (1982). Sex-typed play in gender-disturbed children: A comparison to sibling and psychiatric controls. Archives of Sexual Behavior, 11, 309-321.

ZUGER, B. (1970). Gender role determination: A critical review of the evidence from hermaphroditism. Psychosomatic Medicine, 32, 449-463.

ZWIREN, G. T., & PATTERSON, J. H. (1965). Extrophy of the cloaca: Report of a case treated surgically. Pediatrics, 35, 687-692.

Copyright Society for the Scientific Study of Sex 1999
Provided by ProQuest Information and Learning Company. All rights Reserved

Return to Pseudohermaphroditism
Home Contact Resources Exchange Links ebay