Skip to main content
Molecular Endocrinology logoLink to Molecular Endocrinology
. 2007 Sep 20;22(1):1–9. doi: 10.1210/me.2007-0250

SRY and the Standoff in Sex Determination

Leo DiNapoli 1, Blanche Capel 1
PMCID: PMC2725752  PMID: 17666585

Abstract

SRY was identified as the mammalian sex-determining gene more than 15 yr ago and has been extensively studied since. Although many of the pathways regulating sexual differentiation have been elucidated, direct downstream targets of SRY are still unclear, making a top down approach difficult. However, recent work has demonstrated that the fate of the gonad is actively contested by both male-promoting and female-promoting signals. Sox9 and Fgf9 push gonads towards testis differentiation. These two genes are opposed by Wnt4, and possibly RSPO1, which push gonads toward ovary differentiation. In this review, we will discuss the history of the field, current findings, and exciting new directions in vertebrate sex determination.

X, Y, AND THE GONAD

MOST ADULT MAMMALS exhibit clear sexual dimorphism that manifests externally and internally. Early studies of the sexes catalogued anatomical and endocrinological differences between males and females while the big question remained elusive: What makes them different to begin with? The first genetic clue came when Theophilus Shickel Painter (1) discovered that all males were XY and all females were XX. This led to the conclusion that genetic sex is determined when an X- or Y-bearing sperm meets an X-bearing egg. However, during early embryonic development, the two sexes are indistinguishable. The link between the heteromorphic sex chromosomes and gender was a big step, but the relationship between the sex chromosomes and the mechanism that determines male vs. female fate was still unclear.

The dimorphic development of male and female embryos originates with a fate decision that occurs in the fetal gonad. This unique bipotential organ begins as an indifferent tissue. At about 7–8 wk of development in humans, or midgestation in the mouse, the gonad adopts a testis or ovary fate. The importance of the gonad as the regulator of sexual development was first demonstrated by Alfred Jost (2). Working in France at the end of World War II, Jost performed a seminal embryological experiment that revealed the connection between the gonad and sexual fate and set the tone for the study of sex determination for decades. Jost removed undifferentiated gonads from fetal rabbits in utero and found that all of the embryos, regardless of their X/Y composition, developed female anatomy including the development of an oviduct, a uterus, a vagina, and female external genitalia. From these experiments, he concluded that male sexual development requires a gonad whereas female sexual development could be considered a default state that can be established in the absence of a gonad.

Jost demonstrated that there are at least two active factors associated with the testis that promote male sexual development. One of them is testosterone, which actively promotes male development of the epididymis, vas deferens, and male genitalia. However, Jost also showed that a second factor produced by the testis [now known to be anti-Müllerian hormone (3)] is responsible for inducing regression of the Müllerian duct, the antecedent of the oviduct, uterus, and vagina in females. Based on these experiments, Jost predicted the existence of a male sex-determining pathway the primary role of which was to regulate development of the gonad as a testis. The idea that male sexual development was active and female sexual development was passive emerged from this work and formed the conceptual framework for subsequent research.

THE Y OF SRY

How do the X- and Y-chromosomes alter the sexual fate of the gonad? After the work of Painter and Jost, it was not clear whether the presence of the Y or the lack of two X’s was responsible for male development. The answer came in 1959 when two groups examined the human conditions known as Turner’s Syndrome and Klinefelter’s Syndrome. Individuals with Turner’s Syndrome are XO (having only one sex chromosome) and are phenotypically female, whereas individuals with Klinefelter’s Syndrome are XXY and phenotypically male (4,5). This positively identified the Y-chromosome as the factor that engenders maleness, and the study of how the Y was directing male development began in earnest. Taking a genetic approach, groups began to whittle down the Y-chromosome looking for the testis-determining region (called TDYin humans, or testis-determining factor, Tdf, in mice). In 1990, the Y was pared down to a single gene SRY (sex-determining region of the Y; Sry in mice) thought to hold the secret of the switch between male and female development (6).

Here the sex chromosome mapping analysis and embryological experiments began to converge. Consistent with the implications of Jost’s experiments, Sry was shown to be expressed in the XY gonad (7,8). Conclusive proof came shortly thereafter with the aid of transgenic mice. XY mice lacking Sry develop an ovary and follow female sexual development pathways (9,10), and XX mice with a transgenic autosomal copy of Sry develop a testis and male secondary sex characteristics, although they are sterile in the absence of a Y-chromosome, which is required for spermatogenesis (11,12). These experiments firmly established the relationship between the Y-chromosome, the gonad, and sex determination.

SRY IN SUPPORTING CELLS

Mouse chimera experiments conducted before the time that Sry was isolated predicted that the male sex determinant would be required only in the Sertoli cell lineage (13,14). These chimeras, constructed between XX and XY embryos, showed a strong bias for XY cells in the Sertoli cell lineage, but not in other lineages of the testis (14). Many experiments suggest that, consistent with theoretical predictions, the cells that are precursors to the Sertoli cell lineage are the same cells that can give rise to follicle (granulosa) cells in the ovary (12,13,14,15,16,17). Therefore, this lineage is referred to as the “supporting cell lineage” based on its role in supporting development of germ cells in both sexes.

Sry is expressed in these supporting cell lineage precursors in the XY gonad. Primary sex determination, at least in mammals, appears to be focused on the cell fate decision that occurs in this precursor population when cells chose to differentiate as either Sertoli or follicle cells. Both the spatial and temporal regulation of Sry levels are critical for its ability to pattern the gonad. Sry is expressed in XY gonads, specifically in cells that initiate differentiation as Sertoli cells, at the stage when the gonad first forms as a bipotential primordium (7,8,15,19,20,21,22). The gonad represents a unique environment because ectopic expression of Sry outside this tissue does not lead to differentiation of Sertoli cells (23). The 5′-region of the Sry gene has been shown to be hypomethylated in the gonad, but hypermethylated in other tissues (24). Low or delayed expression leads to the development of a gonad containing a mixture of male and female tissues (an ovotestes) (20,25). Although expression of Sry may be transient in individual cells, based on the persistent Sry transgenic reporters, Sry-enhanced green fluorescent protein and Sry-alkaline phosphatase, it is likely that all Sertoli cells express Sry at some time during their differentiation (15,21).

SRY THE PROTEIN

SRY was the founding member of the Sox [Sry-related high-mobility group (HMG) box] family of transcription factors, although it appears to have evolved from an X-linked member of the family, Sox3 (26,27,28). The defining characteristic of this family is the presence of a HMG DNA-binding domain (29). SRY is thought to bind and bend DNA and may also be involved in chromatin remodeling (29,30,31,32). Although a binding site for the protein was predicted through PCR selection assays, the 6-bp sequence occurs so frequently in the genome that it has not been useful in defining targets of SRY in vivo (31,32,33). Consistent with the importance of the DNA binding function of SRY, most sex-reversing mutations occur within the HMG domain (33). Furthermore, this domain is the only conserved feature across mammalian SRYs. Regions outside this domain have evolved so rapidly that they are difficult to align between species (34).

As would be expected of a transcription factor, SRY is imported into the nucleus, and both importin-β and calmodulin have been shown to play a role in its translocation in vitro (35,36). From there, the water becomes murkier because SRY has been proposed to act as both an activator (37,38) and a repressor (39,40,41) of transcription. One report has even suggested that SRY may act as a direct modulator of splicing (42).

BELOW THE SWITCH

Only 10% of sex reversal cases are linked to mutations in SRY, indicating that other genes must also be critical in the sex determination pathway. The first gene known to be expressed downstream of Sry is Sox9, a closely related family member. Sox9 is expressed in all Sertoli cells, and both loss and gain of function evidence indicates that it is critical to initiate the testis pathway. Deletion of the Sox9 gene leads to male-to-female sex reversal in humans (43,44,45) and mice (46), whereas duplication or overexpression of Sox9 leads to female-to-male sex reversal (47,48,49). These experiments indicate that Sox9 can function as a substitute for Sry itself, even producing fertile males when expressed at sufficient levels in XY Sry null embryos (50). Sox9 may be the only direct target of SRY; however, direct evidence for regulation of Sox9 by SRY is still missing (16,21).

Interestingly, only an average of 90% of Sertoli cells in XX↔XY chimeras were XY cells (14). The occurrence of XX cells in this population indicated that non-cell-autonomous factors must also operate during mammalian sex determination and must be capable of recruiting XX cells to differentiate along the Sertoli pathway. Although at one time this finding was puzzling, it has proved to be conceptually enlightening. Several extracellular signaling pathways have been implicated in recruiting the cells of the gonad to the testis pathway. For example, prostaglandin D2 (PDG2) promotes the Sertoli cell fate and induces Sox9 expression in XX cells in vitro (16,51,52,53,54), whereas fibroblast growth factor 9 (Fgf9) has been shown to induce Sox9 in XX cells in vitro and is required for maintenance of Sox9 expression and testis differentiation in vivo (17,55,56). Further support for the role of extracellular signaling molecules in development of the gonad comes from the observation that in XX↔XY chimeras where the XY component is more than 30%, the gonad still differentiates as a testis (14). These findings strongly argue for the ability of extracellular signals to recruit cells to the testis developmental pathway.

Proliferation is also critical to testis development and is up-regulated in XY gonads soon after Sry expression begins. This results in the expansion of Sertoli precursor cells (56,57,58). Disruption of proliferation in vivo through disruption of the Fgf9 pathway, or through the use of proliferation inhibitors, leads to male-to-female sex reversal (17,58). Taken together, the XX↔XY chimera experiments, along with the data showing that reduced expression of Sry leads to sex reversal (20,59,60,61), have led to the interpretation that there is a critical threshold effect in the gonad when the primordium is balanced between ovarian and testis fates. Both the level of SRY produced by individual cells and the number of cells producing it may contribute to this threshold effect.

This precarious balance of the gonad between these two developmental pathways is likely what confers its dual potential. Normally, the system employs complex cell signaling loops that reinforce a single fate decision in the supporting cell lineage and recruit all gonadal cells behind the testicular or ovarian pathway. Defects in these reinforcing signaling loops may explain many disorders of incomplete sexual development that manifest as gonadal dysgenesis, ovotestis formation, ambiguous ductal or genitalia development, or a combination of these features.

THE TROUBLE WITH XX MALES

When SRY was discovered in 1990 and shown to encode a DNA-binding protein, it fit well with the prediction that the sex-determining gene would be a master regulator that activated many downsteam pathways leading to testis development. However, the persistent fly in the ointment was the occurrence of XX males that show no evidence of the SRY gene. If the male sex-determining gene is required in the bipotential gonad to activate testis development, how does an XX gonad establish testis organogenesis in the absence of SRY? A parsimonious explanation proposed by McElreavey et al. (39) and Goodfellow and Lovell-Badge (62) was that SRY actually acts as a negative regulator of a hypothetical gene Z, the role of which is to repress the testis pathway. In such a double-negative system, it is easy to explain how a single mutation rendering Z inactive could lead to initiation of testis development in the absence of SRY. Recent experiments in the field have developed this concept of inhibition further.

REMODELING THE Z HYPOTHESIS: SEX IN THE BALANCE

The Fgf- and Wnt-signaling pathways have been implicated in regulating the fate of the gonad. Mutations in Fgf9 lead to male-to-female sex reversal, whereas mutations in Wnt4 lead to partial female-to-male sex reversal (17,63,64,65). In the absence of Fgf9, Sry is expressed normally in the XY gonad and Sox9 expression is initiated. However Sox9 expression is rapidly silenced in the absence of Fgf9, and the cells of the XY gonad express genes characteristic of the female pathway, namely, Wnt4, Follistatin, and Bmp2 (17,65). These data indicate that Fgf9 is required to promote the male pathway and suppress the female pathway.

Wnt4 appears to play a reciprocal role in the female pathway. In the absence of Wnt4, Fgf9 and Sox9 expression are transiently up-regulated in the XX gonad. Activation of Sox9 occurs in this case in the absence of the Sry gene, indicating that expression of Sox9 does not require Sry but can be accomplished simply by down-regulating Wnt4. This suggests that Wnt4, like the hypothetical gene Z, normally represses the male pathway (namely, Sox9) in the XX gonad. However, unlike Z, Wnt4 is also required for male development (64). The current model is an evolution from a simple repressor of a repressor to a model of balanced opposition, both sides of which may be important to establish the initial bipotential field of cells.

Sex reversal in Wnt4 null XX gonads is incomplete, suggesting that there may be additional factors that reinforce the female pathway. The canonical Wnt signaling pathway acts by stabilizing β-catenin, which then acts as a cofactor for transcriptional activation. Recently, mutations in the R-SPONDIN1 (RSPO1) gene were identified in human XX patients with testis development (66). Importantly, this is the first human mutation that results in complete female-to-male sex reversal. RSPO1 has also been shown to activate the β-catenin signaling pathway (67), which raises the possibility that Wnt4 and Rspo1 act cooperatively to block the male pathway in XX gonads. Duplication of the human chromosome 1 region, where both WNT4 and RSPO1 map, led to disruption of testis development in humans (68). Surprisingly, overexpression of Wnt4 had only minor effects on testis development in mice (68,69,70). This could be related to an inappropriate spatial or temporal expression of these transgenes, or it could mean that ectopic expression of both Wnt4 and Rspo1 would be required.

There is a precedent for the regulation of cell fate decisions through this seemingly disparate collection of molecules. During chondrogenesis, Wnt signaling activates β-catenin, whereas and Fgf signaling activates Sox9 signaling, respectively, activate β-catenin and SOX9 in chondrocyte precursors (71,72,73,74,75). In chondrocytes, β-catenin and SOX9 have been shown to bind together and trigger the degradation of both proteins until one of the two prevails (73,74).

By analogy, this raises the possibility of an intracellular stand-off between these pathways in supporting cell precursors during the bipotential stage of gonad development (Fig. 1). Primary sex determination might occur when one of these pathways gains the upper hand, silences the other pathway, and triggers a cell fate decision resulting in the differentiation of either Sertoli or follicle cells. Although competition between SOX9 and β-catenin is an appealing concept from many points of view, Wnt4 can also signal through noncanonical pathways (76,77), and no evidence has been presented thus far to indicate that Wnt signaling acts through β-catenin in the gonad.

Figure 1.

Figure 1

Reinforcing Loops that Propagate Testis or Ovary Fate

The somatic precursors of the supporting cell lineage can become Sertoli cells or follicle cells dependent upon whether male- (FGF9) or female- (WNT4) promoting signals dominate the field. In a male environment, these cells become Sertoli cells, organize testis development, promote Leydig cell development, and regulate germ cells that arrest in G0/G1 [PGC (G1)]. In a female environment, the somatic precursors become follicle cell precursors. Their organization into follicle structures is dependent on meiotic germ cells [PGC (Mei)]. It is likely that reinforcing signals (dotted lines) exist between fetal follicle or Sertoli cells and theca or Leydig cells, respectively, although this has not yet been shown. Question marks denote putative interactions between Leydig or theca cells and germ cells.

These observations paint a picture that is markedly different from the classic male-active/female-passive model. It now appears that the bipotential gonad is the battleground between two active and opposing signaling pathways that converge on the regulation of the Sox9/Fgf9 loop (Fig. 2). Sox9 is a conserved element of the male pathway in all vertebrates examined, and even in one species of fly (78). In mammals, Sry tips this balanced Fgf/Wnt antagonism toward testis development by boosting expression of Sox9. This model provides an easy explanation for the occurrence of XX individuals who develop a testis in the absence of Sry. It is also easy to imagine how this plastic system could respond to diverse regulators such as temperature or behavioral cues that control sex determination in other species.

Figure 2.

Figure 2

Model of Opposed Signals in Mammalian Sex Determination

A, In the bipotential gonad, male-promoting (SOX9 and FGF9) and female-promoting (WNT4 and possibly RSPO1) hold each other in check. B, The presence of SRY (XY) reinforces the positive feedback between SOX9 and FGF9, which then out-competes the female signals and drives testis differentiation. In the absence of SRY (XX), the female-promoting signals shut down the male loop and drive ovarian differentiation.

SPECIES WITHOUT SRY

SRY is critical for male development in most mammals, but what about species in which there is no SRY? A few rodents appear to have lost SRY, and even a Y-chromosome, but retain fertile males and females (79). Also, to be precise, only therian mammals have SRY. Monotremes have X- and Y-chromosomes (in fact, the platypus has five Xs and Ys), but no SRY gene (80,81,82). Because SRY likely evolved after the therian/monotreme split between 310 and 130 million years ago (80,83), other vertebrates must have SRY-independent sex-determining mechanisms.

Many alternative sex-determining mechanisms do exist. Interestingly, the primary switches controlling sex determination are highly divergent across species, far more than would be expected considering the conservation among other developmental pathways. Two commonly cited examples of strong evolutionary conservation are the Pax6/eyeless genes that govern development of the eye from Drosophila to humans (84) and the Hox genes that regulate the body axis across phyla (85). However, sex determination stands apart as an emphatic exception to this rule. The reason for the radical variations in sex-determining mechanisms, even among closely related species, is unclear and often debated (86).

Traditionally, sex-determining mechanisms are broadly spit into two categories: genetic sex determination (GSD) and environmental sex determination (ESD). However, even within a single category, there is diversity. For example, mammals and birds use GSD, but mammals have male heterogamety (XX/XY), whereas birds have female heterogamety (ZW/ZZ) (87). Furthermore, the sex chromosomes of mammals and birds appear to have evolved from entirely different autosomes (88,89). On the other side, there is variation in temperature-dependent sex determination, a common form of ESD in which the temperature of egg incubation determines the sex. Temperature-dependent sex determination is found in all crocodilians and many turtles, but the ranges of temperatures, or the end result of a given temperature, are quite disparate (90,91,92). To further complicate matters, there is also no phylogenetic consistency in sex-determining mechanisms across groups of vertebrates. ESD and GSD (either XX/XY or ZW/ZZ) appear to have evolved independently multiple times within fish, lizard, and amphibian lineages (86).

It was once proposed that GSD and ESD were so different as to be mutually exclusive (93,94). This has been reconsidered in light of evidence suggesting that all sex-determining mechanisms may have underlying genetic components (95,96). Indeed, ESD and GSD have even been shown to coexist within some species (97). Now, rather than being seen as Manichean opposites, GSD and ESD are probably best viewed as distinct engines with partially overlapping mechanisms (86,98).

So what, if anything, can these radically different methods of sex determination tell us about mammalian development? Despite the variability of the switches at the top of the sex-determining cascade, the structure of the adult testis is morphologically well conserved in all vertebrates. This has led to the idea that the pathways downstream of the switch are conserved. That is to say, regardless of how evolution has altered the top of the cascade, the movers and shakers down the chain that direct gonadal differentiation are the same. Comparative analysis between mammals, the most well studied, and other vertebrates have lent support to this model. Almost every mammalian gene known to play a role in early gonad development has been found to have similar expression patterns in birds and reptiles (87,91,99,100,101). Notably, Sox9, the gene that is the likely vortex of the fate determination decision in the supporting cells of the gonad, appears to be conserved in testis development in every vertebrate examined to date.

One attractive aspect of the current opposing pathways model is that it can parsimoniously explain the existence of a wide array of sex-determining mechanisms. The evolutionary appearance or disappearance of mechanisms that affect the balance between Sox9/Fgf9 and Wnt4/Rspo1 could be very rapid. For example, temperature might affect the biochemical activity of a component in the Wnt or Fgf pathway. Alternatively, or additionally, allelic variants might evolve that segregate in the population and push the gonad in one direction or the other, eventually founding a new sex chromosome (102). Multiple mechanisms may simultaneously exist within a species, one of which emerges as the predominant modus operandi at any one time (86,97,98). However, the underlying system could remain very plastic in its ability to respond to a variety of sex-determining switches.

PROPAGATING THE DECISION

Sex determination is a strongly canalized process. Once a fate decision is made within the supporting cell lineage of the gonad, the effects of feedback loops at hormonal and physiological levels reinforce the male or female pathway. This may also be true at the intracellular and intercellular levels. Feed-forward loops that reinforce intracellular decisions have proved to be a common theme across many developing tissues, and the gonad is no exception. As previously stated, Sox9 works in a feed-forward loop with Fgf9. These two genes up-regulate each other and generate a cellular momentum moving toward Sertoli cell fate. Layered onto this intracellular loop is an extracellular mechanism that works to recruit other cells in the gonad to the developmental plan through the expression of additional secreted signals. Paracrine signals, such as prostaglandin D2 (51,52,53,54), also reinforce the male pathway. Leydig cells, the male steroidogenic cells, develop as the result of desert hedgehog (Dhh) and platelet-derived growth factor A (PdgfA) signals produced by Sertoli cells (103,104). Furthermore, morphogenetic changes, such as testis cord formation, appear to reinforce the differentiation of Sertoli cells and the establishment of the male pathway. Wnt signaling represses this male loop and drives the differentiation of follicle cells. Although it is less well understood, this promotes a female signaling network between follicle cells, the female steroidogenic precursors (theca cells), and germ cells. This establishes an ovarian environment hostile to male development (Ref.105 and Fig. 2). These local reinforcing environments in the gonad likely explain the observation that ovarian and testicular portions of an ovotestis occupy distinct domains.

GERM CELLS IN THE CONTESTED PRIMORDIUM

Although germ cells are not critical for testis morphogenesis, they are required for follicle formation in the ovary (106). Most experiments in mammals are consistent with the idea that the ovarian pathway initiates normally in the absence of germ cells but fails to result in follicle formation and a functional ovary at birth. When germ cells are absent, the production of testosterone from the testis is not lost; however, the cycling hormonal balance in the ovary is completely disrupted.

Indirect evidence indicates that germ cells initiate dimorphic development soon after they enter the gonad. Male and female germ cells show differences in sensitivity to mutations in Fgf9 and to Mvh (mouse vasa homolog) by 11.5 d post coitum (65,107). Germ cells that enter an ovary initiate meiosis soon after 13.5 d post coitum. Retinoic acid, which is likely produced by the adjacent mesonephros, has been shown to induce this process in XX gonads (18,108). In contrast, germ cells that enter a testis arrest in G0/G1 of the mitotic cycle until after birth. Sertoli cells play a critical role in blocking the entry of germ cells into meiosis in XY gonads, perhaps through the production of Cyp26b1, an enzyme that degrades retinoic acid (18). There is evidence that meiotic germ cells can disrupt testis development in vitro, inhibiting male-specific migration and cord formation (105). Thus development of germ cells in concordance with the sexual fate of the somatic cells in the gonad may be important not only for fertility, but may be critical to reinforce the fate of the gonad and further canalize development of a testis or ovary.

FINAL REMARKS

Many questions are still outstanding about how these intracellular and extracellular pathways work to establish sex determination in the gonad. It is clear that the same pathways that regulate the development of many other organs are also involved in the development of the gonad. However, the system is adapted to regulate the unique dimorphic developmental fate of this primordial organ. During the bipotential phase of gonad development, opposing pathways hold the cells of the gonad in a bipotential state until it is launched down the male or female pathway. SRY remains central as the primary trigger for the initiation of testis development in mammals. Whether there is a parallel transcription factor regulating the initiation of ovary development (i.e. a female-determining gene) is unknown, but not theoretically required. Regardless of whether or not one is found, it is clear that initiation of the ovarian pathway involves the active regulation of many genes and is not simply a passive/default developmental process.

Although it has been valuable to place SRY within this signaling network in the last few years, the mechanism through which the gene functions has not yet been revealed. Future work will address this important unresolved issue and will focus on the integration between the intracellular/intercellular signaling pathways that regulate cell fate and structural morphogenesis of the testis and ovary. Current efforts to establish networks of gene interactions will be critical to provide a framework for future investigations.

Acknowledgments

We thank Steve Munger, Danielle Maatouk, and Tony DeFalco for careful reading of the manuscript, and other members of the laboratory for useful discussions.

Footnotes

Disclosure Statement: The authors have nothing to disclose.

This work was supported by grants (to B.C.) from the National Institutes of Health (HL630540 and HD039963), the National Science Foundation (NSF 0317234), and the Lance Armstrong Foundation.

First Published Online September 20, 2007

Abbreviations: ESD, Environmental sex determination; GSD, genetic sex determination; HMG, high-mobility group; Sox, Sry-related HMG box; SRY, sex-determining region of the Y.

References

  1. Painter TS 1923 Studies in mammalian spermatogenesis. II. The spermatogenesis of man. J Exp Zool 37:291–338 [Google Scholar]
  2. Jost A 1947 Recherches sur la diff′erenciation sexuelle de l’embryon de lapin. Arch Anat Microsc Morph Exp 36:271–315 [Google Scholar]
  3. Josso N, Cate RL, Picard JY, Vigier B, di Clemente N, Wilson C, Imbeaud S, Pepinsky RB, Guerrier D, Boussin L, Legeai L, Carre-Eusebe D 1993 Anti-mullerian hormone: the Jost factor. Recent Prog Horm Res 148:1–59 [DOI] [PubMed] [Google Scholar]
  4. Ford CE, Jones KW, Polani PE, De Almeida JC, Briggs JH 1959 A sex-chromosome anomaly in a case of gonadal dysgenesis (Turner’s syndrome). Lancet 1:711–713 [DOI] [PubMed] [Google Scholar]
  5. Jacobs PA, Strong JA 1959 A case of human intersexuality having a possible XXY sex-determining mechanism. Nature 183:302–303 [DOI] [PubMed] [Google Scholar]
  6. Sinclair AH, Berta P, Palmer MS, Hawkins JR, Griffiths BL, Smith MJ, Foster JW, Frischauf AM, Lovell-Badge R, Goodfellow PN 1990 A gene from the human sex-determining region encodes a protein with homology to a conserved DNA-binding motif. Nature 346:240–244 [DOI] [PubMed] [Google Scholar]
  7. Koopman P, Munsterberg A, Capel B, Vivian N, Lovell-Badge R 1990 Expression of a candidate sex-determining gene during mouse testis differentiation. Nature 348:450–452 [DOI] [PubMed] [Google Scholar]
  8. Hacker A, Capel B, Goodfellow P, Lovell-Badge R 1995 Expression of Sry, the mouse sex determining gene. Development 121:1603–1614 [DOI] [PubMed] [Google Scholar]
  9. Lovell-Badge R, Robertson E 1990 XY female mice resulting from a heritable mutation in the primary testis-determining gene, Tdy. Development 109:635–646 [DOI] [PubMed] [Google Scholar]
  10. Gubbay J, Vivian N, Economou A, Jackson D, Goodfellow P, Lovell-Badge R 1992 Inverted repeat structure of the Sry locus in mice. Proc Natl Acad Sci USA 89:7953–7957 [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Koopman P, Gubbay J, Vivian N, Goodfellow P, Lovell-Badge R 1991 Male development of chromosomally female mice transgenic for Sry. Nature 351:117–121 [DOI] [PubMed] [Google Scholar]
  12. Eicher EM, Shown EP, Washburn LL 1995 Sex reversal in C57BL/6J-YPOS mice corrected by a Sry transgene. Philos Trans R Soc Lond B Biol Sci 350:263–268 [DOI] [PubMed] [Google Scholar]
  13. Burgoyne PS, Buehr M, McLaren A 1988 XY follicle cells in ovaries of XX-XY female mouse chimaeras. Development 104:683–688 [DOI] [PubMed] [Google Scholar]
  14. Palmer SJ, Burgoyne PS 1991 In situ analysis of fetal, prepuberal and adult XX↔XY chimaeric mouse testes: Sertoli cells are predominantly, but not exclusively, XY. Development 112:265–268 [DOI] [PubMed] [Google Scholar]
  15. Albrecht K, Eicher E 2001 Evidence that Sry is expressed in pre-Sertoli cells and Sertoli and granulosa cells have a common precursor. Dev Biol 240:92–107 [DOI] [PubMed] [Google Scholar]
  16. Wilhelm D, Martinson F, Bradford S, Wilson MJ, Combes AN, Beverdam A, Bowles J, Mizusaki H, Koopman P 2005 Sertoli cell differentiation is induced both cell-autonomously and through prostaglandin signaling during mammalian sex determination. Dev Biol 287:111–124 [DOI] [PubMed] [Google Scholar]
  17. Kim Y, Kobayashi A, Sekido R, DiNapoli L, Brennan J, Chaboissier MC, Poulat F, Behringer RR, Lovell-Badge R, Capel B 2006 Fgf9 and Wnt4 act as antagonistic signals to regulate mammalian sex determination. PLoS Biol 4:e187 [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Bowles J, Knight D, Smith C, Wilhelm D, Richman J, Mamiya S, Yashiro K, Chawengsaksophak K, Wilson MJ, Rossant J, Hamada H, Koopman P 2006 Retinoid signaling determines germ cell fate in mice. Science 312:596–600 [DOI] [PubMed] [Google Scholar]
  19. Bullejos M, Koopman P 2001 Spatially dynamic expression of Sry in mouse genital ridges. Dev Dyn 221:201–205 [DOI] [PubMed] [Google Scholar]
  20. Bullejos M, Koopman P 2005 Delayed Sry and Sox9 expression in developing mouse gonads underlies B6-Y(DOM) sex reversal. Dev Biol 278:473–481 [DOI] [PubMed] [Google Scholar]
  21. Sekido R, Bar I, Narvaez V, Penny G, Lovell-Badge R 2004 SOX9 is up-regulated by the transient expression of SRY specifically in Sertoli cell precursors. Dev Biol 274:271–279 [DOI] [PubMed] [Google Scholar]
  22. Salas-Cortes L, Jaubert F, Barbaux S, Nessmann C, Bono MR, Fellous M, McElreavey K, Rosemblatt M 1999 The human SRY protein is present in fetal and adult Sertoli cells and germ cells. Int J Dev Biol 43:135–140 [PubMed] [Google Scholar]
  23. Kidokoro T, Matoba S, Hiramatsu R, Fujisawa M, Kanai-Azuma M, Taya C, Kurohmaru M, Kawakami H, Hayashi Y, Kanai Y, Yonekawa H 2005 Influence on spatiotemporal patterns of a male-specific Sox9 activation by ectopic Sry expression during early phases of testis differentiation in mice. Dev Biol 278:511–525 [DOI] [PubMed] [Google Scholar]
  24. Nishino K, Hattori N, Tanaka S, Shiota K 2004 DNA methylation-mediated control of Sry gene expression in mouse gonadal development. J Biol Chem 279:22306–22313 [DOI] [PubMed] [Google Scholar]
  25. Taketo T, Lee CH, Zhang J, Li Y, Lee CY, Lau YF 2005 Expression of SRY proteins in both normal and sex-reversed XY fetal mouse gonads. Dev Dyn 233:612–622 [DOI] [PubMed] [Google Scholar]
  26. Collignon J, Sockanathan S, Hacker A, Cohen-Tannoudji M, Norris D, Rastan S, Stevanovic M, Goodfellow PN, Lovell-Badge R 1996 A comparison of the properties of Sox-3 with Sry and two related genes, Sox-1 and Sox-2. Development 122:509–520 [DOI] [PubMed] [Google Scholar]
  27. Graves JA 1998 Interactions between SRY and SOX genes in mammalian sex determination. Bioessays 20:264–269 [DOI] [PubMed] [Google Scholar]
  28. Bowles J, Bullejos M, Koopman P 2000 Screening for novel mammalian sex-determining genes using expression cloning and microarray approaches. Aust Biochemist 31:4–6 [Google Scholar]
  29. Ferrari S, Harley VR, Pontiggia A, Goodfellow PN, Lovell-Badge R, Bianchi ME 1992 SRY, like HMG1, recognizes sharp angles in DNA. EMBO J 11:4497–4506 [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Giese K, Cox J, Grosschedl R 1992 The HMG domain of lymphoid enhancer factor 1 bends DNA and facilitates assembly of functional nucleoprotein structures. Cell 69:185–195 [DOI] [PubMed] [Google Scholar]
  31. Giese K, Pagel J, Grosschedl R 1994 Distinct DNA-binding properties of the high mobility group domain of murine and human SRY sex-determining factors. Proc Natl Acad Sci USA 91:3368–3372 [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Pontiggia A, Rimini R, Harley VR, Goodfellow PN, Lovell-Badge R, Bianchi ME 1994 Sex-reversing mutations affect the architecture of SRY-DNA complexes. EMBO J 13:6115–6124 [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Harley VR, Jackson DI, Hextall PJ, Hawkins JR, Berkovitz GD, Sockanathan S, Lovell-Badge R, Goodfellow PN 1992 DNA binding activity of recombinant SRY from normal males and XY females. Science 255:453–456 [DOI] [PubMed] [Google Scholar]
  34. Whitfield LS, Lovell-Badge R, Goodfellow PN 1993 Rapid sequence evolution of the mammalian sex-determining gene SRY. Nature 364:713–715 [DOI] [PubMed] [Google Scholar]
  35. Harley VR, Layfield S, Mitchell CL, Forwood JK, John AP, Briggs LJ, McDowall SG, Jans DA 2003 Defective importin β recognition and nuclear import of the sex-determining factor SRY are associated with XY sex-reversing mutations. Proc Natl Acad Sci USA 100:7045–7050 [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Sim H, Rimmer K, Kelly S, Ludbrook LM, Clayton AH, Harley VR 2005 Defective calmodulin-mediated nuclear transport of the sex-determining region of the Y chromosome (SRY) in XY sex reversal. Mol Endocrinol 19:1884–1892 [DOI] [PubMed] [Google Scholar]
  37. Cohen DR, Sinclair AH, McGovern JD 1994 Sry protein enhances transcription of Fos-related antigen 1 promoter constructs. Proc Natl Acad Sci USA 91:4372–4376 [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Dubin RA, Ostrer H 1994 Sry is a transcriptional activator. Mol Endocrinol 8:1182–1192 [DOI] [PubMed] [Google Scholar]
  39. McElreavey K, Vilain E, Cotinot C, Payen E, Fellous M 1993 Control of sex determination in animals. Eur J Biochem 218:769–783 [DOI] [PubMed] [Google Scholar]
  40. Desclozeaux M, Poulat F, Barbara PD, Capony JP, Turowski P, Jay P, Mejean C, Moniot B, Boizet B, Berta P 1998 Phosphorylation of an N-terminal motif enhances DNA-binding activity of the human SRY protein. J Biol Chem 273:7988–7995 [DOI] [PubMed] [Google Scholar]
  41. Oh HJ, Li Y, Lau YF 2005 Sry associates with the heterochromatin protein 1 complex by interacting with a KRAB domain protein. Biol Reprod 72:407–415 [DOI] [PubMed] [Google Scholar]
  42. Ohe K, Lalli E, Sassone-Corsi P 2002 A direct role of SRY and SOX proteins in pre-mRNA splicing. Proc Natl Acad Sci USA 99:1146–1151 [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Wagner T, Wirth J, Meyer J, Zabel B, Held M, Zimmer J, Pasantes J, Bricarelli FD, Keutel J, Hustert E, Wolf U, Tommerup N, Schempp W, Scherer G 1994 Autosomal sex reversal and campomelic dysplasia are caused by mutations in and around the SRY-related gene SOX9. Cell 79:1111–1120 [DOI] [PubMed] [Google Scholar]
  44. Foster JW 1996 Mutations in SOX9 cause both autosomal sex reversal and campomelic dysplasia. Acta Paediatr Jpn 38:405–411 [DOI] [PubMed] [Google Scholar]
  45. Smyk M, Obersztyn E, Nowakowska B, Bocian E, Cheung SW, Mazurczak T, Stankiewicz P 2007 Recurrent SOX9 deletion campomelic dysplasia due to somatic mosaicism in the father. Am J Med Genet A 143:866–870 [DOI] [PubMed] [Google Scholar]
  46. Chaboissier MC, Kobayashi A, Vidal VI, Lutzkendorf S, van de Kant HJ, Wegner M, de Rooij DG, Behringer RR, Schedl A 2004 Functional analysis of Sox8 and Sox9 during sex determination in the mouse. Development 131:1891–1901 [DOI] [PubMed] [Google Scholar]
  47. Huang B, Wang S, Ning Y, Lamb AN, Bartley J 1999 Autosomal XX sex reversal caused by duplication of SOX9. Am J Med Genet 87:349–353 [DOI] [PubMed] [Google Scholar]
  48. Bishop CE, Whitworth DJ, Qin Y, Agoulnik AI, Agoulnik IU, Harrison WR, Behringer RR, Overbeek PA 2000 A transgenic insertion upstream of Sox9 is associated with dominant XX sex reversal in the mouse. Nat Genet 26:490–494 [DOI] [PubMed] [Google Scholar]
  49. Vidal V, Chaboissier M, de Rooij D, Schedl A 2001 Sox9 induces testis development in XX transgenic mice. Nat Genet 28:216–217 [DOI] [PubMed] [Google Scholar]
  50. Qin Y, Bishop CE 2005 Sox9 is sufficient for functional testis development producing fertile male mice in the absence of Sry. Hum Mol Genet 14:1221–1229 [DOI] [PubMed] [Google Scholar]
  51. Adams IR, McLaren A 2002 Sexually dimorphic development of mouse primordial germ cells: switching from oogenesis to spermatogenesis. Development 129:1155–1164 [DOI] [PubMed] [Google Scholar]
  52. Malki S, Nef S, Notarnicola C, Thevenet L, Gasca S, Mejean C, Berta P, Poulat F, Boizet-Bonhoure B 2005 Prostaglandin D2 induces nuclear import of the sex-determining factor SOX9 via its cAMP-PKA phosphorylation. EMBO J 24:1798–1809 [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Wilhelm D, Hiramatsu R, Mizusaki H, Widjaja L, Combes AN, Kanai Y, Koopman P 2007 SOX9 regulates prostaglandin D synthase gene transcription in vivo to ensure testis development. J Biol Chem 282:10553–10560 [DOI] [PubMed] [Google Scholar]
  54. Malki S, Bibeau F, Notarnicola C, Roques S, Berta P, Poulat F, Boizet-Bonhoure B 2007 Expression and biological role of the prostaglandin D synthase/SOX9 pathway in human ovarian cancer cells. Cancer Lett 255:182–193 [DOI] [PubMed] [Google Scholar]
  55. Colvin JS, Green RP, Schmahl J, Capel B, Ornitz DM 2001 Male-to-female sex reversal in mice lacking fibroblast growth factor 9. Cell 104:875–889 [DOI] [PubMed] [Google Scholar]
  56. Schmahl J, Kim Y, Colvin JS, Ornitz DM, Capel B 2004 Fgf9 induces proliferation and nuclear localization of FGFR2 in Sertoli precursors during male sex determination. Development 131:3627–3636 [DOI] [PubMed] [Google Scholar]
  57. Karl J, Capel B 1998 Sertoli cells of the mouse testis originate from the coelomic epithelium. Dev Biol 203:323–333 [DOI] [PubMed] [Google Scholar]
  58. Schmahl J, Eicher E, Washburn L, Capel B 2000 Sry induces cell proliferation in the mouse gonad. Development 127:65–73 [DOI] [PubMed] [Google Scholar]
  59. Nagamine CM, Morohashi K, Carlisle C, Chang DK 1999 Sex reversal caused by Mus musculus domesticus Y chromosomes linked to variant expression of the testis-determining gene Sry. Dev Biol 216:182–194 [DOI] [PubMed] [Google Scholar]
  60. Albrecht KH, Capel B, Washburn LL, Eicher EM 2000 Defective mesonephric cell migration is associated with abnormal testis cord development in C57BL/6JMus domesticus mice. Dev Biol 225:26–36 [DOI] [PubMed] [Google Scholar]
  61. Washburn LL, Albrecht KH, Eicher EM 2001 C57BL/6J-T-associated sex reversal in mice is caused by reduced expression of a Mus domesticus Sry allele. Genetics 158:1675–1681 [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Goodfellow PN, Lovell-Badge R 1993 SRY and sex determination in mammals. Annu Rev Genet 27:71–92 [DOI] [PubMed] [Google Scholar]
  63. Vainio S, Heikkila M, Kispert A, Chin N, McMahon AP 1999 Female development in mammals is regulated by Wnt-4 signalling. Nature 397:405–409 [DOI] [PubMed] [Google Scholar]
  64. Jeays-Ward K, Dandonneau M, Swain A 2004 Wnt4 is required for proper male as well as female sexual development. Dev Biol 276:431–440 [DOI] [PubMed] [Google Scholar]
  65. DiNapoli L, Batchvarov J, Capel B 2006 FGF9 promotes survival of germ cells in the fetal testis. Development 133:1519–1527 [DOI] [PubMed] [Google Scholar]
  66. Parma P, Radi O, Vidal V, Chaboissier MC, Dellambra E, Valentini S, Guerra L, Schedl A, Camerino G 2006 R-spondin1 is essential in sex determination, skin differentiation and malignancy. Nat Genet 38:1304–1309 [DOI] [PubMed] [Google Scholar]
  67. Kim KA, Zhao J, Andarmani S, Kakitani M, Oshima T, Binnerts ME, Abo A, Tomizuka K, Funk WD 2006 R-Spondin proteins: a novel link to β-catenin activation. Cell Cycle 5:23–26 [DOI] [PubMed] [Google Scholar]
  68. Jordan BK, Mohammed M, Ching ST, Delot E, Chen XN, Dewing P, Swain A, Rao PN, Elejalde BR, Vilain E 2001 Up-regulation of WNT-4 signaling and dosage-sensitive sex reversal in humans. Am J Hum Genet 68:1102–1109 [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Jeays-Ward K, Hoyle C, Brennan J, Dandonneau M, Alldus G, Capel B, Swain A 2003 Endothelial and steroidogenic cell migration are regulated by WNT4 in the developing mammalian gonad. Development 130:3663–3670 [DOI] [PubMed] [Google Scholar]
  70. Jordan BK, Shen JH, Olaso R, Ingraham HA, Vilain E 2003 Wnt4 overexpression disrupts normal testicular vasculature and inhibits testosterone synthesis by repressing steroidogenic factor 1/β-catenin synergy. Proc Natl Acad Sci USA 100:10866–10871 [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Weksler NB, Lunstrum GP, Reid ES, Horton WA 1999 Differential effects of fibroblast growth factor (FGF) 9 and FGF2 on proliferation, differentiation and terminal differentiation of chondrocytic cells in vitro. Biochem J 342:677–682 [PMC free article] [PubMed] [Google Scholar]
  72. Rozenblatt-Rosen O, Mosonego-Ornan E, Sadot E, Madar-Shapiro L, Sheinin Y, Ginsberg D, Yayon A 2002 Induction of chondrocyte growth arrest by FGF: transcriptional and cytoskeletal alterations. J Cell Sci 115:553–562 [DOI] [PubMed] [Google Scholar]
  73. Akiyama H, Lyons JP, Mori-Akiyama Y, Yang X, Zhang R, Zhang Z, Deng JM, Taketo MM, Nakamura T, Behringer RR, McCrea PD, de Crombrugghe B 2004 Interactions between Sox9 and β-catenin control chondrocyte differentiation. Genes Dev 18:1072–1087 [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Hill TP, Spater D, Taketo MM, Birchmeier W, Hartmann C 2005 Canonical Wnt/β-catenin signaling prevents osteoblasts from differentiating into chondrocytes. Dev Cell 8:727–738 [DOI] [PubMed] [Google Scholar]
  75. Yano F, Kugimiya F, Ohba S, Ikeda T, Chikuda H, Ogasawara T, Ogata N, Takato T, Nakamura K, Kawaguchi H, Chung UI 2005 The canonical Wnt signaling pathway promotes chondrocyte differentiation in a Sox9-dependent manner. Biochem Biophys Res Commun 333:1300–1308 [DOI] [PubMed] [Google Scholar]
  76. Maurus D, Heligon C, Burger-Schwarzler A, Brandli AW, Kuhl M 2005 Noncanonical Wnt-4 signaling and EAF2 are required for eye development in Xenopus laevis. EMBO J 24:1181–1191 [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Osafune K, Takasato M, Kispert A, Asashima M, Nishinakamura R 2006 Identification of multipotent progenitors in the embryonic mouse kidney by a novel colony-forming assay. Development 133:151–161 [DOI] [PubMed] [Google Scholar]
  78. DeFalco TJ, Verney G, Jenkins AB, McCaffery JM, Russell S, Van Doren M 2003 Sex-specific apoptosis regulates sexual dimorphism in the Drosophila embryonic gonad. Dev Cell 5:205–216 [DOI] [PubMed] [Google Scholar]
  79. Sutou S, Mitsui Y, Tsuchiya K 2001 Sex determination without the Y chromosome in two Japanese rodents Tokudaia osimensis osimensis and Tokudaia osimensis spp. Mamm Genome 12:17–21 [DOI] [PubMed] [Google Scholar]
  80. Foster JW, Brennan FE, Hampikian GK, Goodfellow PN, Sinclair AH, Lovell-Badge R, Selwood L, Renfree MB, Cooper DW, Graves JA 1992 Evolution of sex determination and the Y chromosome: SRY-related sequences in marsupials. Nature 359:531–533 [DOI] [PubMed] [Google Scholar]
  81. Grutzner F, Rens W, Tsend-Ayush E, El-Mogharbel N, O’Brien PC, Jones RC, Ferguson-Smith MA, Marshall Graves JA 2004 In the platypus a meiotic chain of ten sex chromosomes shares genes with the bird Z and mammal X chromosomes. Nature 432:913–917 [DOI] [PubMed] [Google Scholar]
  82. Rens W, Grutzner F, O’brien PC, Fairclough H, Graves JA, Ferguson-Smith MA 2004 Resolution and evolution of the duck-billed platypus karyotype with an X1Y1X2Y2X3Y3X4Y4X5Y5 male sex chromosome constitution. Proc Natl Acad Sci USA 101:16257–16261 [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Griffiths R 1991 The isolation of conserved DNA sequences related to the human sex-determining region Y gene from the lesser black-backed gull (Larus fuscus). Proc Biol Sci 244:123–128 [DOI] [PubMed] [Google Scholar]
  84. Kozmik Z 2005 Pax genes in eye development and evolution. Curr Opin Genet Dev 15:430–438 [DOI] [PubMed] [Google Scholar]
  85. Lemons D, McGinnis W 2006 Genomic evolution of Hox gene clusters. Science 313:1918–1922 [DOI] [PubMed] [Google Scholar]
  86. Janzen FJ, Phillips PC 2006 Exploring the evolution of environmental sex determination, especially in reptiles. J Evol Biol 19:1775–1784 [DOI] [PubMed] [Google Scholar]
  87. Smith CA, Sinclair AH 2004 Sex determination: insights from the chicken. Bioessays 26:120–132 [DOI] [PubMed] [Google Scholar]
  88. Nanda I, Shan Z, Schartl M, Burt DW, Koehler M, Nothwang H, Grutzner F, Paton IR, Windsor D, Dunn I, Engel W, Staeheli P, Mizuno S, Haaf T, Schmid M 1999 300 Million years of conserved synteny between chicken Z and human chromosome 9. Nat Genet 21:258–259 [DOI] [PubMed] [Google Scholar]
  89. Nanda I, Zend-Ajusch E, Shan Z, Grutzner F, Schartl M, Burt DW, Koehler M, Fowler VM, Goodwin G, Schneider WJ, Mizuno S, Dechant G, Haaf T, Schmid M 2000 Conserved synteny between the chicken Z sex chromosome and human chromosome 9 includes the male regulatory gene DMRT1: a comparative (re)view on avian sex determination. Cytogenet Cell Genet 89:67–78 [DOI] [PubMed] [Google Scholar]
  90. Wibbels T, Cowan J, LeBoeuf R 1998 Temperature-dependent sex determination in the red-eared slider turtle, Trachemys scripta. J Exp Zool 281:409–416 [DOI] [PubMed] [Google Scholar]
  91. Pieau C, Dorizzi M, Richard-Mercier N 1999 Temperature-dependent sex determination and gonadal differentiation in reptiles. Cell Mol Life Sci 55:887–900 [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Western PS, Sinclair AH 2001 Sex, genes, and heat: triggers of diversity. J Exp Zool 290:624–631 [DOI] [PubMed] [Google Scholar]
  93. Bull JJ 1980 Sex determination in reptiles. Q Rev Biol 55:3–21 [Google Scholar]
  94. Janzen FJ, Paukstis GL 1991 Environmental sex determination in reptiles: ecology, evolution, and experimental design. Q Rev Biol 66:149–179 [DOI] [PubMed] [Google Scholar]
  95. Wilkins AS 1995 Moving up the hierarchy—a hypothesis on the evolution of a genetic sex determination pathway. Bioessays 17:71–77 [DOI] [PubMed] [Google Scholar]
  96. Wilkins AS 2002 Sex determination. In: Pagel M, ed. Encyclopedia of evolution. Vol 2. Oxford, UK: Oxford University Press; 1033–1037 [Google Scholar]
  97. Nakamura M, Kobayashi T, Chang XT, Nagahama Y 1998 Gonadal sex differentiation in teleost fish. J Exp Zool 281:362–372 [Google Scholar]
  98. Sarre SD, Georges A, Quinn A 2004 The ends of a continuum: genetic and temperature-dependent sex determination in reptiles. Bioessays 26:639–645 [DOI] [PubMed] [Google Scholar]
  99. Yao HH, Capel B 2005 Temperature, genes, and sex: a comparative view of sex determination in Trachemys scripta and Mus musculus. J Biochem 138:5–12 [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Morrish BC, Sinclair AH 2002 Vertebrate sex determination: many means to an end. Reproduction 124:447–457 [DOI] [PubMed] [Google Scholar]
  101. Shoemaker C, Ramsey M, Queen J, Crews D 2007 Expression of Sox9, Mis, and Dmrt1 in the gonad of a species with temperature-dependent sex determination. Dev Dyn 236:1055–1063 [DOI] [PubMed] [Google Scholar]
  102. Graves JA 2006 Sex chromosome specialization and degeneration in mammals. Cell 124:901–914 [DOI] [PubMed] [Google Scholar]
  103. Yao HH, Whoriskey W, Capel B 2002 Desert hedgehog/patched 1 signaling specifies fetal Leydig cell fate in testis organogenesis. Genes Dev 16:1433–1440 [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Brennan J, Tilmann C, Capel B 2003 Pdgfr-α mediates testis cord organization and fetal Leydig cell development in the XY gonad. Genes Dev 17:800–810 [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Yao HH, DiNapoli L, Capel B 2003 Meiotic germ cells antagonize mesonephric cell migration and testis cord formation in mouse gonads. Development 130:5895–5902 [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Guigon CJ, Coudouel N, Mazaud-Guittot S, Forest MG, Magre S 2005 Follicular cells acquire Sertoli cell characteristics after oocyte loss. Endocrinology 146:2992–3004 [DOI] [PubMed] [Google Scholar]
  107. Tanaka SS, Toyooka Y, Akasu R, Katoh-Fukui Y, Nakahara Y, Suzuki R, Yokoyama M, Noce T 2000 The mouse homolog of Drosophila vasa is required for the development of male germ cells. Genes Dev 14:841–853 [PMC free article] [PubMed] [Google Scholar]
  108. Koubova J, Menke DB, Zhou Q, Capel B, Griswold MD, Page DC 2006 Retinoic acid regulates sex-specific timing of meiotic initiation in mice. Proc Natl Acad Sci USA 103:2474–2479 [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Molecular Endocrinology are provided here courtesy of The Endocrine Society

RESOURCES