Next Article in Journal
Transcriptome, Ectopic Expression and Genetic Population Analysis Identify Candidate Genes for Fiber Quality Improvement in Cotton
Next Article in Special Issue
Comprehensive Analysis of YTH Domain-Containing Genes, Encoding m6A Reader and Their Response to Temperature Stresses and Yersinia ruckeri Infection in Rainbow Trout (Oncorhynchus mykiss)
Previous Article in Journal
Computational Modeling Analysis of Kinetics of Fumarate Reductase Activity and ROS Production during Reverse Electron Transfer in Mitochondrial Respiratory Complex II
Previous Article in Special Issue
HuR Promotes the Differentiation of Goat Skeletal Muscle Satellite Cells by Regulating Myomaker mRNA Stability
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Overview of Avian Sex Reversal

1
Department of Animal Genetics and Breeding, College of Animal Science and Technology, China Agricultural University, Beijing 100193, China
2
National Engineering Laboratory for Animal Breeding and Key Laboratory of Animal Genetics, Breeding and Reproduction, Ministry of Agriculture and Rural Affairs, China Agricultural University, Beijing 100193, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(9), 8284; https://doi.org/10.3390/ijms24098284
Submission received: 27 March 2023 / Revised: 28 April 2023 / Accepted: 29 April 2023 / Published: 5 May 2023
(This article belongs to the Special Issue Molecular Genetics and Breeding Mechanisms in Domestics Animals)

Abstract

:
Sex determination and differentiation are processes by which a bipotential gonad adopts either a testicular or ovarian cell fate, and secondary sexual characteristics adopt either male or female developmental patterns. In birds, although genetic factors control the sex determination program, sex differentiation is sensitive to hormones, which can induce sex reversal when disturbed. Although these sex-reversed birds can form phenotypes opposite to their genotypes, none can experience complete sex reversal or produce offspring under natural conditions. Promising evidence indicates that the incomplete sex reversal is associated with cell autonomous sex identity (CASI) of avian cells, which is controlled by genetic factors. However, studies cannot clearly describe the regulatory mechanism of avian CASI and sex development at present, and these factors require further exploration. In spite of this, the abundant findings of avian sex research have provided theoretical bases for the progress of gender control technologies, which are being improved through interdisciplinary co-operation and will ultimately be employed in poultry production. In this review, we provide an overview of avian sex determination and differentiation and comprehensively summarize the research progress on sex reversal in birds, especially chickens. Importantly, we describe key issues faced by applying gender control systems in poultry production and chronologically summarize the development of avian sex control methods. In conclusion, this review provides unique perspectives for avian sex studies and helps scientists develop more advanced systems for sex regulation in birds.

1. Introduction

Birds are important agricultural species and long-standing development model animals [1,2]. For a long time, the global poultry industry has sought effective methods to control the sex of hatchlings to achieve different purposes. In the layer industry, for example, only female chickens are required, and day-old male hatchlings are cruelly culled, which seriously increases ethical and legal concerns [3,4,5,6,7]. Deciphering the mechanisms of avian sex determination and differentiation can contribute to achieving sex control in chickens and avoiding such issues. Furthermore, owing to the high accessibility of chicken embryos, which can be genetically manipulated in vitro, they are ideal models for gonadal research [8,9,10,11]. Given the conservation of gonadal morphogenesis, studies on chicken embryos have shed light on the developmental patterns of the human reproductive system and the causes of sex disorders [12]. Thus, understanding the cell biology and genetics of gonadal formation during avian embryogenesis is crucial for improving our knowledge in diverse areas, from livestock breeding to human sex development.
In birds, the sex of offspring is determined by genetic factors. Male determinant genes activate and maintain the cascade reaction of testicular development, whereas female sex-related pathways are responsible for the proliferation and differentiation of the ovary [13,14,15]. However, avian sex differentiation is regulated by multiple factors, including genetics, epigenetics, and sex hormones [16,17]. Under natural conditions, for example, disease-induced gene mutation and alteration of estrogen levels can significantly divert the direction of sex differentiation, leading to sex reversal [17,18,19]. Although adult sex-reversed chickens form almost identical phenotypes to their opposite genetic sexes, they cannot experience complete sex reversal and produce offspring, which is caused by the cell autonomous sex identity (CASI) of avian cells [16,17,20,21]. Promising investigations indicates that hereditary factors are responsible for avian CASI. The research on the sex development of gynandromorphic birds suggests that the CASI is governed by the cellular sex chromosome combination, and is vital to the maintenance of genetic sex [21]. Until now, we have failed to clarify the underlying basis of this biological feature in birds, which needs further research.
Fundamental research on avian sex determination and differentiation can inspire the development of sex control technologies for the poultry industry, ultimately enabling the production of single-sex hatchlings. In the early years, several embryonic sex detection and selection methods based on various biological characteristics, such as egg shape, egg odors, and genetic information, were invented [22,23,24,25,26]. Although the development of these systems is limited by many challenges, such as economic benefits and animal welfare issues, they are still being applied and improved upon [27,28]. With the recent boom in CRISPR/Cas9 systems, direct editing of sex-determining genes has become possible [14,29]. This will help poultry industries to achieve precise gender regulation during production and provide insights for the study of more diversified sex control strategies.
In this review, we describe the entire process of sex determination and differentiation in birds, particularly emphasizing genetic and epigenetic regulation. Importantly, we focus on research on sex reversal in chickens and speculate that CASI might contribute to incomplete sex reversal in birds. Finally, we highlight the contribution of sex research to the advancement in gender control technologies and clarify the developmental progress of avian sex control systems. In summary, this review will help researchers understand the sex development of birds from the perspective of genetics and epigenetics, and provide theoretical support for the research of sex control methods in poultry industries.

2. Avian Sex Determination

2.1. Mechanism of Vertebrate Sex Determination

Under natural conditions, the mechanisms of vertebrate sex determination can be divided into two categories: environmental sex determination (ESD) and genetic sex determination (GSD) [30,31]. The sex-determining pattern of most reptiles, such as crocodiles, turtles, and lizards, is consistent with ESD [32,33,34,35,36,37,38]. One of the most common ESD mechanisms is temperature-dependent sex determination (TSD), which indicates that the sex of offspring is determined by ambient temperature during incubation [39]. For example, in the red-eared slider turtle, eggs of Trachemys scripta elegans incubated at 26 °C produce male hatchlings, whereas higher temperatures (32 °C) result in the birth of females [40]. However, GSD usually occurs in higher vertebrates, such as mammals and birds. In these species, the sex of the offspring is controlled by the combination of sex chromosomes carried by the sperm and ovum at fertilization [31]. Most mammals, such as mice, have an XX/XY sex chromosome system, which is a feature of the male heterogamety (XY) [41,42]. The Y chromosome-linked Sex-Determining Region Y (SRY) or male-specific repetitive DNA sequences directly govern the program of sex determination [43,44,45,46,47]. In contrast, birds have a ZZ/ZW sex chromosome system, and avian heterogametic sex is discovered in females (ZW) [48]. Studies have only identified a few groups of bona fide genes in the W chromosome, and their biological functions remain unclear [49,50,51,52].

2.2. Genetic and Epigenetic Regulation in Avian Sex Determination

Previous evidence has shown that avian sex is controlled by dominant female determinant genes located on the W chromosome or governed by the Z chromosome dosage effect. The former hypothesis suggests that the W chromosome carries one or several female sex determination-related genes, which are akin to the SRY gene on the Y chromosome [53]. These W-linked genes are female-specific due to recombination suppression and adaptive processes, and are likely conducive to the activation of female developmental pathways [54,55,56,57,58,59]. Although the W-linked Histidine Triad Nucleotide Binding Protein (HINTW) and Female Expressed Transcript 1 (FET1) are the most promising candidates, attempts to verify their functions in ovarian development have proven unsuccessful [60,61,62,63]. Therefore, it is possible that the W chromosome may not be at the center of female sex determination. Nevertheless, the presence of the W chromosome is indispensable for maintaining female differentiation [14]. Studies have suggested that aneuploid birds with ZZW sex chromosomal karyotypes form feminized phenotypes, whereas Z0 (only possessing a single sex chromosome) develop into masculine phenotypes [64,65,66]. Hence, although current research has failed to identify one female sex-determining gene on the W chromosome, it may carry elements that stimulate female development.
Unlike X chromosome inactivation regulated by the X Inactive Specific Transcript (XIST) in mammals, avian species lack global Z chromosome inactivation, which causes sex determination in avian species to be governed by the dosage effect of genes in the Z chromosome [67,68,69,70,71,72,73]. On average, male birds have two Z chromosomes; however, their female counterparts have only one. This implies that the expression level of some Z-linked genes, which are widely expressed in various tissues, is twice as high in males than in females [70,72,74,75,76]. Concerning gonads, most high-throughput studies have focused on the Doublesex and Mab-3 Related Transcription Factor 1 (DMRT1), a Z-specific gene that encodes a zinc-finger-like transcription factor similar to other DMRT family members [13,14]. The DMRT1 gene is highly conserved across lower and higher vertebrates, and its homologs have been reported to play crucial roles in vertebrate sex development [77,78,79,80,81,82,83,84,85,86]. For instance, its duplicated copy, the DM-domain gene on the Y chromosome (DMY), is a promising sex determination-related gene in medaka fish (Oryzias latipes) (XX) [87,88,89,90]. As expected, this effect is also confirmed in birds. DMRT1 has two copies in males (ZZ) but one copy in females, and has been reported to be unaffected by the dosage compensation effect [91]. The locus of DMRT1 has been shown to generate a number of diverse alternatively spliced transcripts; however, a recent investigation suggests that only a single DMRT1 transcript is expressed in the developing chicken gonads [92,93]. DMRT1 is first detectable at E3.5 and is transcribed exclusively in the urogenital system [94,95,96,97,98,99]. Overexpression of DMRT1 results in the masculinization of female embryos, characterized by the activation of testicular differentiation-related genes, SRY-box transcription factor 9 (SOX9), and anti-Müllerian hormone (AMH) [13,98,100]. Importantly, knocking out the half copy of DMRT1 in male chickens leads to the formation of ovaries, which are characterized by the expression of female development-related markers, such as Cytochrome P450 Family 19 Subfamily A Member 1 (CYP19A1) and Forkhead Box L2 (FOXL2) [13,14,100]. This implies that two functional copies of DMRT1 are necessary to stimulate masculinized development, and one copy of DMRT1 is insufficient to initiate the cascade reaction of male sex determination and repress female developmental pathways in birds (Figure 1).
Owing to the pivotal role of DMRT1, tight control of its spatiotemporal expression is crucial for regulating normal sex developmental programs. Evidence has confirmed that the transcriptional activity of DMRT1 is heavily controlled by epigenetic factors, including long noncoding RNA (lncRNA) and histone modifications. For example, the Male HyperMethylated (MHM) region, a 2.2 kb repeat sequence located on the Z chromosome that is only transcribed in female cells from a particular strand into lncRNAs, is considered to play a crucial role in the dosage compensation and regulation of DMRT1 expression [103,104,105]. This region is hypermethylated in males and hypomethylated in females [106,107,108,109,110,111]. The global overexpression of MHM inhibits the transcriptional activity of DMRT1 in adult chicken gonads, leading to male-to-female sex reversal, suggesting that lncRNAs transcribed from this locus are likely associated with sex determination in birds [103,104,112]. In addition, current research indicates that H3K27ac, an indicator of active enhancers, is densely occupied near the DMRT1 locus in males and is highly distributed around MHM in females, which is consistent with the expression pattern of DMRT1 [113]. Therefore, based on evidences above, the absence of the dosage compensation effect highlights the importance of Z-linked DMRT1, which is fine-tuned by genetic and epigenetic elements, in determining avian sex.

3. Avian Sex Differentiation

3.1. Morphological Changes in Reproductive Organs during Avian Sex Differentiation

Sex differentiation is an essential part of sex development. DMRT1 activates sex determination as a master switch between the male and female cell fate of the gonad, and numerous genes and hormones are responsible for maintaining the subsequent process of sex differentiation underlying gonadal development in birds (Figure 1). Considering chicken as a model, gonadal progenitors appear from precursors of the coelomic epithelium and mesonephros at approximately embryonic day 2 (E2, Hamburger Hamilton Stage, HH6) [114,115,116,117,118]. The undifferentiated gonad (the so-called genital ridge), which develops on the ventromedial surface of the mesonephric kidney, primarily consists of an outer cortex and inner medulla [12,119]. Their proliferation and differentiation are mainly maintained by asymmetric cell division [9,120]. Daughter cells produced by the division remain in the coelomic epithelium and mesonephros, and the rest migrate into the genital ridge to form functional gonadal cell lineages [9,120]. At the primary stage, three types of cells are encased in the gonad: supporting cell precursors, steroidogenic progenitors, and primordial germ cells (PGCs) [121,122].
The supporting cell is the first bipotent somatic lineage to be differentiated under the signal of sex determination in embryonic gonads [102,123]. In chickens, embryonic mesonephros gives rise to supporting cell precursors, forming Sertoli and Leydig cells in males or pre-granulosa and theca cells in females (Figure 1) [101]. PGCs ingress into gonads, proliferating and differentiating into primary spermatocytes or oocytes, via the bloodstream from the extra-embryonic germinal crescent [124]. The morphological difference of bilateral gonads is first macroscopically visible at E5.5 (HH28) [119]. In males, the bipotential gonads develop into symmetrical testes, which are characterized by the progressive proliferation of sex cords in medullae, providing suitable niches for the differentiation of PGCs, and the flattening of outer cortical layers (Figure 1) [124,125,126]. However, the situation in females is quite different. The cortical layer of the female left gonad is well-developed, with many PGCs condensed inside, and the inner medulla appears largely unstructured, containing fluid-filled vacuoles known as lacunae (Figure 1) [127,128,129,130]. In contrast, cortical regions of the right side are only encircled by a simple epithelial layer and experience arrested proliferation in later stages [16,128,131].
Unlike gonads, the reproductive ducts of birds derive from two primary embryonic structures: Wolffian and Müllerian ducts [132,133,134]. In males, Wolffian ducts give rise to the vas deferens, whereas Müllerian ducts progressively degenerate under the influence of AMH at E9 (HH35) [135,136,137,138]. Conversely, in females, only the Müllerian duct on the left side differentiates into a functional oviduct, whereas the right Müllerian duct and bilateral Wolffian ducts gradually regress [133,139]. During ovarian formation, estrogens provide a suitable environment to prevent the left oviduct from being affected by AMH [140,141]. Notably, in these two seemingly independent structures, the differentiation of Müllerian ducts is closely associated with Wolffian ducts [142,143]. Ablation of Wolffian ducts in the early embryonic stages leads to the failure of Müllerian ducts [144]. Therefore, it is considered that the local signaling molecule produced by intact Wolffian ducts is required for cell proliferation and caudal migration of the Müllerian ducts [145,146].

3.2. Genetic and Epigenetic Regulation in Male Sex Differentiation

In male birds, SOX9, AMH, and HEMOGN (HEMGN) are essential factors that maintain sex development (Figure 1). At the onset of gonadal differentiation, DMRT1 acts as a pioneer transcription factor (pTF) to stimulate the downstream highly conserved genes, SOX9 and AMH [13,14,98]. Chicken SOX9 is characterized by a high mobility group (HMG) box DNA-binding domain and is upregulated during testicular differentiation [147,148,149,150,151,152]. It is mainly expressed in supporting cells and triggers the differentiation program toward pre-Sertoli cells [101,153]. The function of SOX9 is conserved in mammals and birds. A previous study confirmed that AMH is the only SOX9 target gene shared in mouse and chicken male gonads (mainly Sertoli cells) at identical embryonic periods through comparative SOX9 chromatin immunoprecipitation sequencing (ChIP-seq) analysis [154]. This suggests that the transcriptional activity of chicken AMH may be regulated by SOX9, similar to that observed in mice. However, the expression of chicken AMH is first detected at E4.5 (HH25), preceding SOX9 (E6.5, HH30) [137,155]. Therefore, SOX9 may not directly activate AMH expression but may maintain it in birds. Focusing on the function of AMH, it is vital to the formation of avian Wolffian ducts and testes, but the overexpression or knockdown of AMH in embryonic chickens causes abnormal gonadal development [156]. Therefore, the proper activation of AMH is essential for forming the reproductive system in birds.
Another promising male sex differentiation-related gene is Z-linked HEMGN [157]. Research has indicated that HEMGN is expressed in primitive blood cells and functions in hematopoiesis in mice [158]. Avian HEMGN, nevertheless, is highly expressed in Sertoli cells of gonads [157]. Its expression is first detected at E5.5 (HH28), then increases significantly to a peak at E8.5 [157]. The overexpression of HEMGN in female embryos through retroviruses induces the upregulation of DMRT1 and masculinization of gonads [157]. In addition, overexpression of DMRT1 in female embryonic gonads induces the activation of HEMGN [98]. Hence, this evidence illustrates that a positive feedback loop exists between DMRT1 and HEMGN and that the male-specific expression of HEMGN is pivotal to the differentiation of testes in birds.
Recently, studies found that the Transducin-Like Enhancer of Split 4 (TLE4Z1), which presents significant male preference at E4.5, can induce gonadal defeminization when it is overexpressed in female embryos and is involved in the masculinization of embryonic gonads [159]. Moreover, SPINDLIN1-Z (SPIN1Z), a Z-linked gene, is important for initiating male development [160]. Evidence has shown that overexpression of SPIN1Z increases the transcriptional activity of SOX9 and AMH in females; however, its knockdown shows a reversed phenomenon [160]. In addition, a recent investigation pointed out that the SMAD family member 2 (SMAD2) is a testicular differentiation-related gene in chickens, and its disruption inhibits the expression of DMRT1 and SOX9 in male gonads [161]. In summary, the findings for these novel genes illustrate that sex differentiation in male birds is a complex process jointly regulated by a set of genetic factors (Figure 1).
In addition to the genetic regulatory network mentioned above, substantial evidence suggests that epigenetics is also involved in male sex differentiation. The current study revealed significant differences in transcriptome-wide m6A landscapes between E7 chicken female and male left gonads by methylated RNA immunoprecipitation sequencing (MeRIP-seq) [162]. Experiments have shown that the m6A-recognized protein YTH Domain Containing 2 (YTHDC2) can regulate the expression of genes related to male sex differentiation, such as SOX9 and HEMGN [162]. Similarly, studies have revealed striking differences in the landscape of genomic chromatin accessibility between female and male embryonic left gonads [163]. An in-depth investigation has confirmed that the variation in chromatin accessibility corresponds to the alteration of gene expression [163]. For instance, compared to females, the transcriptional activity of DMRT1 in males is upregulated, which is accompanied by increased chromatin accessibility [163]. Based on these findings, we conclude that epigenetic regulation plays a key role in controlling the activation and maintenance of masculine developmental pathways in birds.

3.3. Genetic and Epigenetic Regulation in Female Sex Differentiation

In females, two conventional developmental pathways are considered to act in parallel to promote ovarian development. The first pathway is the FOXL2/aromatase (CYP19A1)/estrogen-signaling pathway (Figure 1). FOXL2 is expressed in a female-specific manner and initially detected in the gonadal medulla at approximately E5.5 (HH28), just slightly ahead of the onset of sex differentiation [15,164]. In later embryonic stages, the expression of FOXL2 also becomes detectable in a group of cortical cells [15]. The mis-expression of FOXL2 in male gonads represses the differentiation of the Sertoli cell lineage and abolishes the local expression of testicular differentiation-related genes, such as DMRT1, SOX9, and AMH [15]. However, the knockdown of FOXL2 induces ectopic activation of SOX9 in females [15,165]. Therefore, FOXL2 may act as a master activator of ovarian development and maintain an antagonistic relationship with male pathways.
Downstream CYP19A1 encodes an enzyme responsible for aromatizing androgens to form estrogens [166,167,168,169,170,171]. It exhibits a sexually dimorphic expression pattern at approximately E5.5 (HH28) and is mainly detected in pre-granulosa cells in female chickens [168,169,172]. The interference of CYP19A1 promotes the development of masculinized medullae and inhibits the growth of cortices in female embryonic gonads, accompanied by the upregulation of SOX9, whereas the overexpression of CYP19A1 induces male embryos feminization [172]. Previous studies assumed that CYP19A1 is controlled by FOXL2 in birds, similar to the case in mammals [173,174]. However, convincing experiments have confirmed that chicken FOXL2 does not directly regulate CYP19A1 [175]. The mis-expression of FOXL2 is insufficient to stimulate CYP19A1 in male embryonic gonads, and the disruption of FOXL2 in females fails to influence the activation of CYP19A1 [172]. Conversely, overexpression of CYP19A1 increases the transcriptional activity of FOXL2 in both sexes, and the inhibition of aromatase activity causes a reduction in FOXL2 levels in female gonads [172,176]. Therefore, these results indicate that avian FOXL2 is unlikely to govern the expression of CYP19A1, but the CYP19A1 can control the transcriptional activity of FOXL2. However, it is also possible that avian CYP19A1 is regulated by FOXL2 in later stages [15].
The fine-tuning of avian CYP19A1 is highly dependent on dynamic epigenetic modifications. In chickens, several epigenetic markers have found in the promoter region of CYP19A1, involving DNA methylation and histone lysine methylation [177]. These modifications exhibit different distribution patterns in male and female individuals, corresponding to the expression levels of CYP19A1 [177]. Notably, experiments have shown that these epigenetic marks can be induced to form feminized modes in estrogen-mediated male-to-female sex-reversed chickens, indicating that DNA methylation is closely involved in governing the transcriptional activity of CYP19A1 [177,178].
In birds, estrogen is an absolute requirement for female sex differentiation, as it is essential for the growth of the ovary and acquisition of secondary sexual characteristics, such as the wattle, comb, leg spurs, and feathering patterns [11,17,20,179]. Under natural conditions, estrogens stimulate the elongation and growth of Müllerian ducts, regulating the formation of tubular glands and the differentiation of the oviductal epithelium into ciliated and goblet cells [180,181,182,183]. The exogenous addition of estrogens can induce the feminization of male chickens, whereas inhibiting the production of estrogens can masculinize female chicken [16,17,91,184]. During female gonadal development, Estrogen Receptor 1 (ESR1) is a signal transducer of estrogens [11]. ESR1 is asymmetrically expressed in both sexes and is restricted to cortical regions of the female left gonad at E6 (HH29) [185]. The activation of ESR1 by propyl-pyrazole-triol (PPT) causes the formation of left-side ovotestis and retention of Müllerian ducts in male embryos [186]. Therefore, estrogens are required for female gonadal sex differentiation in birds, and their receptor ESR1 plays an important role in ovarian development.
Another canonical pathway related to ovarian development is the Wnt Family Member 4 (WNT4)/ R-spondin1 (RSPO1)/β-catenin signaling pathway (Figure 1). WNT morphogens are a highly conserved family of signaling molecules that play a crucial role in the differentiation of female gonads in several animals [12,187,188]. The binding of WNT4 with its coreceptor LDL Receptor Related Protein 6 (LRP6) and RSPO1 allows the formation of a complex that can promote the expression of β-catenin by inhibiting its phosphorylation [189,190,191]. In chickens, WNT4 is expressed in bipotential gonads of both sexes at E4.5 (HH25) [192]. Its expression level gradually increases in female gonads from E6.5 (HH30) to E8.5, which is mainly restricted to pre-granulosa cells in the left cortex and decreases in males [192]. Chicken RSPO1 shows a sexually dimorphic pattern from early stages. Its expression is detectable in female embryonic gonads as early as E4.5 (HH25) and increases strongly from E8.5 [192]. RSPO1 is co-expressed with WNT4 in cortical regions of the female left gonad and loses its expression on the right side since E8.5 [192]. As for β-catenin, its expression is detected in the female left gonad at E6.5 (HH30) and increases significantly in cortical regions at E13.5 [193]. Our knowledge of this pathway derives mainly from mammals and humans. Mutation of the WNT4 gene in humans causes female-to-male sex reversal, which is coupled with renal, adrenal, and lung dysgenesis [194]. Similarly, the lack of WNT4 and RSPO1 in mice induces the differentiation of ovotestes in females and the maldevelopment of testes in males [195,196]. Moreover, knocking out RSPO1 in female mice results in decreased levels of β-catenin signaling molecules and female-to-male sex reversal [197]. However, in birds, we can only speculate that the function of β-catenin is identical to that in mammals. The amino acid sequence of the β-catenin gene is highly conserved in vertebrates, with 99% similarity between chicken and mice [198]. Therefore, it is likely that β-catenin plays a conserved role in chicken ovarian development and female sex differentiation, which needs to be proven in future studies.
Although these two pathways related to ovarian development are expressed in different gonadal regions, FOXL2 and CYP19A1 are expressed mainly in the medulla, whereas the WNT4/RSPO1/β-catenin pathway is detected in the cortex. It has been shown that there are some interplays between them [193]. An investigation found that inhibiting the synthesis of estrogens through aromatase inhibitors curbs the expression of RSPO1 in embryonic ovaries [192]. However, it is unclear whether this effect is caused by the direct interaction of estrogens with RSPO1 gene transcription or the lack of well-developed cortical regions in sex-reversed gonads, which results in leaving a few groups of cells to express RSPO1.
Several genes, independent of these pathways, have recently been identified to play crucial roles in ovarian development (Figure 1). Experiments have shown that the Jun proto-oncogene, AP-1 transcription factor subunit (JUN) functions as an ovarian differentiation-related regulator in embryonic periods [161]. Overexpression of JUN inhibits the expression of SMAD2, DMRT1, and SOX9 in male embryonic gonads while inducing the expression of FOXL2, ESR1, and CYP19A1 [161]. However, the knockdown of JUN leads to the masculinization of female embryonic gonads [161]. A recent study identified a novel regulator of juxtacortical medulla differentiation in female embryonic gonads, termed TGF-β Induced Factor Homeobox 1 (TGIF1) [199]. It is mainly expressed in the cortex and the pre-granulosa cell lineage [199]. Although estrogen-mediated male-to-female sex reversal can induce ectopic activation, results from targeted mis-expression and knockdown of TGIF1 indicate that it is only required, but insufficient, for proper ovarian cortex formation [199]. These results suggest that ovarian differentiation is a polygenic regulatory process. Two signaling pathways, FOXL2/CYP19A1/estrogen, and WNT4/RSPO1/β-catenin, play significant roles in the developmental program.

4. Avian Sex Reversal

4.1. Occurrence of Vertebrate Sex Reversal

In birds and some lower vertebrates, sex differentiation is often affected by various factors, leading to sex reversal. In chicken flocks, the population’s social structure can influence the sex of individuals, similar to the case of fish [200,201,202,203,204]. For instance, when there is no rooster in a chicken flock, a hen may experience sex reversal to maintain the reproduction of the population [135]. This is an adaptive change that arises during the long-term evolutionary process. However, no further research has been conducted to prove this phenomenon. In addition, diseases are a major cause of sex reversal. Ectopic activation of aromatase and accumulation of estrogens as a result of the henny-feathering trait, an autosomal dominant mutation, can feminize the feathering patterns of male chicken [18,205]. However, this disease-mediated sex reversal might not influence gonadal morphology because aromatase is mainly mis-expressed in extragonadal tissues [18,205].

4.2. Mechanism of Avian Sex Reversal

Bulk instances of sex reversal in birds are associated with the high sensitivity to alterations in sex hormone levels [16,206]. This biological characteristic of birds is closely related to their evolutionary processes. From the viewpoint of evolution, mechanisms of mammalian sex determination and differentiation evolved from synapsid reptiles, whereas diapsid reptiles gave rise to crocodiles, lizards, and birds 150 million years ago [125,207]. In eutherian mammals, gonadal sex differentiation appears independent of sex steroid hormones and can proceed without steroidogenesis [208,209]. Environmentally insensitive gonads of highly evolved mammalian embryos occur in the maternal womb, which is a potentially dangerous place rich in various hormones (Figure 2) [210]. Hence, the highly evolved placenta and maternal internal pregnancy pattern may have forced eutherians to abandon estrogens as components of the sex-determining cascade. However, sex differentiation of birds occurs in an extra-maternal environment and has characteristics similar to those of lower vertebrates (Figure 2). Therefore, it is likely that they do not have the mechanism to resist the alteration of exogenous estrogen levels, which is usually unlikely to occur under normal hatching conditions. Therefore, the evolutionary position of birds and the biological characteristics of extra-maternal embryogenesis indicate that avian sex differentiation is prone to being affected by environmental hormones.

4.3. Avian Sex Reversal Induced by Transplant Treatment

Sex-reversed birds can be obtained by altering endogenous hormone secretion through gonadal grafting. Transplantation of E13 (HH39) chicken whole testes into E3 female chicken extra-embryonic coelom can induce gonadal sex reversal [211,212]. Under the stimulation of exogenous testes, the left gonad differentiates into a testis instead of an ovary, and germ cells migrate into sex cords. However, they do not experience a meiotic process [212]. Moreover, grafting E2 female (male) sections of the presumptive mesoderm, which gives rise to gonads, to replace the equivalent tissue of male (female) at the same growth stage, induces the formation of mixed-sex chimeras [21]. Therefore, artificial transplantation of exogenous tissues can affect avian sex differentiation, leading to sex reversal.

4.4. Avian Male-to-Female Sex Reversal Induced by Estrogens Treatment

In most cases, sex-reversed birds are created by adding estrogens or aromatase inhibitors to manipulate endogenous hormone levels (Figure 3) [16,17]. The administration of estrogen to quail eggs during the first half of embryonic life (before the onset of sex differentiation) can feminize genetic males, characterized by the hyperproliferation of the left gonadal cortex and vacuolized inner medulla [16]. Moreover, injecting a 17a-Ethinylestradiol (EE2) emulsion into E3 quail eggs can demasculinize male individuals, which will lose typical masculine sexual behavior and form asymmetric testes with decreased areas of androgen-dependent cloacal glands after maturity [213]. Recently, research has found that treating chicken eggs with this emulsion can feminize male chickens to different degrees [177,178]. The left gonads of these individuals progressively form female-like cortical regions and fluid-filled medullae from low to high degrees of sex reversal, while the characterized testicular tissues are almost lost [178]. However, the effect of estrogen-mediated male-to-female sex reversal is transient. Some reversed individuals revert to normal male phenotypes after hatching, whereas others can persist in female phenotypes for not more than 1 year [125]. In addition, estrogen-regulated sex reversal is not limited to gonads but is also presented in secondary sexual characteristics. Injection of estradiol into leg muscles of adult male chickens can feminize feathering patterns, including reducing saddle feather length, increasing plumulaceous segments, and altering feathering colors (Figure 3) [214]. Similarly, these changes are not permanent and gradually disappear with decreased estrogen levels in the body. Taken together, the addition of estrogens can significantly affect the differentiation of gonadal structures during embryonic periods and influence the formation of secondary sexual characteristics in adult periods; however, these effects are only maintained for a short time.

4.5. Avian Female-to-Male Sex Reversal Induced by Aromatase Inhibitors Treatment

Injecting aromatase inhibitors, which reduce the production of gonadal estrogens, into avian eggs can induce female-to-male sex reversal (Figure 3) [17,215,216,217]. A previous study found that treating E3 chicken eggs with fadrozole, an aromatase inhibitor, can masculinize bilateral female embryonic gonads, which fail to form stratified cortexes, but develop functional medullae enclosing germ cells at E9.5 [218]. Notably, aromatase inhibitors-treated chickens exhibit varying degrees of sex reversal (Figure 3) [184]. The gonadal cortex is reduced in highly sex-reversed individuals, and germ cells are relocated into developed medullary structures [184]. During the development of sex-reversed female embryos, DMRT1, SOX9, and AMH expression is significantly upregulated in gonadal medullae, whereas FOXL2 and CPY19A1 are downregulated in gonadal cortexes with fewer residues in juxtacortical medullae [91,218,219]. In addition, the primary AMH Receptor, AMH receptor type-II (AMHR2), is upregulated in fadrozole-treated females and co-locates with DMRT1 in Sertoli cells [220]. Moreover, gonadal morphology and inner structures of sex-reversed chickens continue to change after hatching. In 1-day-old sex-reversed chickens (D1), both follicles and tubular structures are visible in the gonads [221]. These tubular structures differentiate into abnormal seminiferous tubules at D11 and form atypical tubules with areas of loose connective tissue at D21 [221]. The left and right gonads grow into small ovotestes in D42 chickens, containing greatly enlarged atypical seminiferous tubules and fewer normal appearing seminiferous tubules [221]. This consecutive alteration is due to the permanent effect of aromatase inhibitors on chicken sex differentiation, which is likely achieved by cell reprogramming during the embryonic period. Research has emphasized that fadrozole-mediated female-to-male sex reversal includes a crucial event, namely, pre-granulosa cells trans-differentiate into undifferentiated PAX2+ supporting cells before forming Sertoli cells, which is lost in estrogens-mediated male-to-female sex reversal [222]. However, current evidences are still unable to clarify the potential mechanism of this phenomenon, which needs to be studied in detail.
The permanent effect of aromatase inhibitors on avian sex differentiation can significantly influence the development of reproductive tissues and the appearance of birds. Adult sex-reversed chickens form almost symmetrical gonads and small testes with regressed oviducts on the left side [17,107,223]. In addition, injecting aromatase inhibitors before sex differentiation during the embryonic period or after sexual maturity in the adult period can induce testosterone production while inhibiting estradiol synthesis in the blood [223,224]. The sex-revered females progressively form masculinized hackles, saddle feathers, combs, and wattles. In contrast, their body weights, especially muscle mass and fat pad weight, remain in line with normal females, suggesting that growth performance, unlike secondary sexual characteristics, may be controlled directly by genetic factors and independent of hormones [214,221,223,224,225]. Notably, under the influence of an aromatase inhibitor, Wolffian ducts in sex-reversed chickens develop into vas deferens, and gonads can produce sperms carrying the W chromosome [226]. Further investigation has shown that W-carrying sperms have oocyte-activating potency and can induce the formation of male and female pronuclei [227,228]. In summary, aromatase inhibitors can significantly disrupt estrogen synthesis and permanently influence the process of sex differentiation, causing the formation of sperm-carrying small testes and masculinized secondary sexual characteristics in adult sex-reversed birds.

4.6. Avian Sex Reversal and Cell Autonomous Sex Identity

Sex reversal can be induced in birds, but neither does the male bird has the ability to form well-functional ovaries, nor do female birds show normal masculinized courtship behavior and successfully produce fertile sperm after sexual maturity [17,229]. This incomplete sex reversal may be regulated by the CASI of avian somatic cells, which is a cellular inherent sex identity that can help individuals maintain their genetic sex independently of the impact of hormones [21,74]. Current research points out that the CASI is determined by the genetic information, especially that carried by sex chromosomes, and epigenetic factors, such as DNA methylation and histone lysine methylation [21,177]. At early developmental stages when primary gonads have not been induced to differentiation, sexually dimorphic gene expression can be detected in several tissues between male and female birds, indicating that the CASI is widely involved in organogenesis programs [230,231]. Studies of the CASI mainly focus on gynandromorphic birds, which display a significant bilateral asymmetry; one side of the body appears to be male, but the other appears to be female [74]. In gynandromorphic chickens, most cells on the side with a female appearance are of the ZW genotype, and most cells on the other side with a male appearance are of the ZZ genotype [21]. Notably, the gonadal morphologies of these gynandromorphic chickens conform to their cellular compositions, testicular and ovarian appearances are mainly composed of ZZ and ZW genotypic cells, whereas ovotestis is composed of a mixture of both cells [21]. Consistently, male-determining genes, such as DMRT1 and SOX9, are highly expressed in Sertoli cells of the seminiferous cords on the masculinized side, where aromatase is not detectable [232]. However, the gonads of the feminized side show areas of peripheral aromatase expression, together with areas of SOX9-positive and DMRT1-positive seminiferous tubules [232]. Moreover, the concept of CASI has been confirmed in mixed-sex chicken chimeras. Cells from donor females are restricted to the interstitial tissue between the sex cords and are not recruited into functional Sertoli cells in testes. Similarly, donor male cells cannot be recruited to form granulosa cells in ovaries [21]. That is, there is no interaction between GFP+ donor female cells and AMH+ host male cells in the testes, and GFP+ donor male cells and CYP19A1+ host female cells in the ovaries, suggesting that the differentiation of these cells is confirmed by their genetic fates [21]. Notably, epigenetic factors are likely to contribute to maintaining gonadal fate. The chromatin accessibility of DMRT1 in embryonic female-to-male sex-reversed chickens cannot be increased to the normal male level, corresponding to its transcriptional activity [163]. In addition, the pattern of DNA methylation around CYP19A1 in estrogen-mediated sex-reversed chicken gonads fails to be induced to form the same as that in control females, suggesting that epigenetic elements are involved in the establishment of avian gonadal CASI [177,178].
In addition, studies have found that the CASI also affects the growth of avian secondary sexual characteristics [233,234]. In a gynandromorphic finch, although both halves of the brain are exposed to the same hormone environment, the neural song circuit of the male side has a more masculine phenotype than that of the female side [76]. Strikingly, recent research found that DMRT1-knockout male chicken, which had ovaries instead of testes, can form the normal masculinized appearance, featuring large combs and wattles, hackle feathers, and long leg spurs in adult periods [14]. Similarly, in gynandromorphic chickens, the female side forms sex-linked feathering patterns, smaller spurs, and wattles than those on the male side after sexual maturity [232]. These results suggest that avian CASI may be critical in differentiating sex phenotypes.

5. Avian Sex Control

5.1. Critical Issues and Challenges of Avian Sex Control

One of the most important purposes of avian sex research is to implement it in the poultry industry to solve practical production problems [235,236,237]. At present, billions of 1-day-old male hatchlings are culled globally every year, and their carcasses are further processed into feed by machines or directly buried into soils [3]. Although those methods save the feeding cost, they severely violate animal welfare regulations advocated by a number of countries, and untreated animal carcasses are very likely to cause biosafety hazards which can endanger human health [238,239]. Inspired by gender control systems applied in dairy farms, the development of avian sex control technologies is a promising means to solve the abovementioned concerns and improve breeding efficiency [240,241]. Investigations of avian sex determination, sex differentiation, and sex reversal have provided abundant insights into the development of early sex detection and sex control technologies [14,101,242]. However, sex research in birds is a huge puzzle, full of various unknown regulatory networks, and we cannot clearly explain the detailed molecular biological mechanisms of sex determination and differentiation in birds at present. This creates huge barriers to the progress of avian gender control systems.
In addition, the assumption of the application process in poultry industries greatly limits the research direction and fields of sex control technologies. To ensure the economic benefits of the poultry industry, the applied technologies cannot affect the hatchability of eggs and the growth of chicken embryos, nor can they influence the egg production and survival rate of laying hens [27,243,244]. As a successful gender control strategy must help improve production efficiency, the system should be simply operated and highly mechanized, which can be used in intensive breeding [245,246]. Moreover, biosafety, food security, and animal welfare must be considered in the research process of gender control strategies. Those methods must fit into the political environment and be acceptable from a humanitarian and ethical point of view [28,247,248,249]. For example, sex control technology must be applied before chicken embryos’ pain perception has evolved. The initial sensory afferent nerves develop in the chicken embryo on E4 of incubation, but a synaptic connection to the spinal cord is not present before E7 of incubation, which makes nociception impossible in the first third of incubation, suggesting that the intervention of embryonic sex development must be earlier than E7 in chickens [250,251]. In summary, the development of sex control approaches has been, and is currently, still being measured according to the abovementioned criteria, with many challenges to be overcome on the way from an idea to practical application.

5.2. Research Progress and Application of Avian Sex Control Technologies

With the in-depth development of basic research and interdisciplinary communication, several gender control technologies have been developed, and some systems have even been used in poultry production [23,252,253]. In the early years, the focus of research was, for instance, to investigate the connection between the outer shape of the eggshell and the sex of hatchlings, as well as the association between the egg odor and the embryonic gender (Figure 4) [25,26]. Those studies have provided pivotal perspectives for developing avian sex control systems. However, there is no corresponding data on the likelihood of in ovo sex control based on egg shape and odor attributes [254].
In addition to using morphometric and chemical signals as biomarkers for sex detection, different optical and imaging methods have been successfully performed in birds. A great advantage of optical methods is their contactless nature. Raman spectroscopy, a representative subfield of vibrational spectroscopy based on the so-called “Raman effect”, has been widely employed in biological detection and clinical research [255,256,257]. It is unique to each molecule and is often called “molecular fingerprints.” As the biochemical composition of female and male cells is significantly different, Raman spectroscopy allows for sex identification in ovo according to the spectral signature of germinal or blood cells (Figure 4) [258,259]. For example, sex-specific molecules obtained from the E3 or E5 developing chicken vascular system when embryos have the ability to resist the impact of external stimuli and are earlier than the connection between afferent nerves and spinal cord can be used to distinguish between male and female eggs, respectively [259]. In addition to Raman spectroscopy, some promising spectra, such as Fourier transform infrared spectroscopy and time-resolved laser-induced fluorescence spectroscopy, are also being tested and applied to the sex control of poultry industries [253,260]. With the advance of these optical systems, the Canadian company “Egg Farmers of Ontario” funded a new technology called “Hypereye” which uses hyperspectral imaging to identify infertility and gender of day-of-lay eggs. As those eggs are essentially the same as regular table eggs, early identification could indicate a new source of eggs, including but not limited to the production of human vaccines and anti-depressants [261,262]. Moreover, the US company “Vital Farms” and Israeli company “Novatrans” jointly developed a new technology named “TeraEgg”. This system uses infrared light combined with complex algorithms to analyze the signs of sex and reproduction from the early stage of the embryonic development of chicken eggs. This implies that farms and intensive breeding industries can distinguish between male and infertile eggs, which can be brought back to the market and sold to the public. Multiple systems, including the two instructive sex control methods mentioned above, are being consistently improved with the research progress of avian sex development and spectroscopy and will finally be applied in the poultry industry.
In addition to achieving sex control using optical and imaging methods, the thriving genetically engineered technology represented by the CRISPR/Cas9 system provides a direct way to govern the gender of embryonic chickens. As genes regulate the determination of avian sex, the genetic marking of sex chromosomes has also been discussed as a possible route for in ovo sex control in birds (Figure 4) [14,21]. Based on previous knowledge, several studies have focused on the production of genetically engineered hens and have described the marking of the Z chromosome of breeding hens with various fluorescent proteins [3,263]. By labeling the Z chromosome of breeding hens with fluorescent biomarkers, when mated to non-genetically modified roosters, male-determined eggs can be identified based on their fluorescence signals before incubation or at early-stage incubation. Notably, this method was successfully used for sex control in layers, with sex being detected from sex-specific patterns of germinal disc fluorescence in non-incubated eggs [264]. However, according to current knowledge, this strategy cannot be operated once and for all, because female chickens hatched from such mating do not show any genetically modified genes.
Furthermore, based on gene-editing technologies, research has found that environmental factors can also activate or inhibit the expression of sex-determining genes, thus achieving avian sex control [265,266]. The US company “ONCE” has developed a “genetic switch” that can be used to selectively activate or deactivate light absorption centers in sex-determining genes using narrowband light-emitting diodes. In this way, the sex of oviparous embryos can be influenced at early differentiation stages. It should be possible to selectively produce only offspring of the desired sex (females in laying hen production and males in broiler production). Similarly, the Israeli company “Soos Technology” has invented a solution to affect the gene expression of the reproductive system in genetic male embryos, which can turn males into functional females capable of laying eggs. The artificial intelligence-guided incubation system developed by this company uses a patented combination of temperature, humidity, CO2, and sonic vibrations to control the sex development process in poultry embryos and turn genetically male chickens into functional females. However, no reliable results have been published, and no information is available concerning trials or the practical use of those methods. Nevertheless, those non-contact gene expression regulation modes provide new ideas for follow-up research on gender control technologies and are expected to open up more advanced sex control strategies in the near future.

6. Conclusions and Future Perspectives

The sex development of birds is a precise process. Research has indicated that the dosage effect of DMRT1 is the master switch that controls the activation of avian sex determination [13,14]. However, subsequent sex differentiation is a complex process that involves multiple genetic factors. To date, studies have identified several important genes related to this program, e.g., SOX9, AMH, and HEMGN in males and FOXL2, CYP19A1, WNT4, and RSPO1 in females, but have failed to elaborate the regulatory network between these genes and the molecular basis of the everlasting antagonistic relationship between testicular and ovarian pathways [13,15,187]. Therefore, future studies are needed to explore the internal regulatory mechanisms between avian sex-related genes.
In birds, sex reversal can be induced by various methods, including but not limited to the addition of exogenous estrogens and aromatase inhibitors. Sex-reversed birds are important for studying avian sex determination and differentiation because of their unique sex development modes [17,20,176]. Notably, regardless of the type of treatment used, the chicken cannot achieve complete sex reversal, even during embryonic periods. Current evidence indicates that this is due to the CASI of avian somatic cells [21]. The alteration of hormone levels can only cause partial or temporary sex reversal in birds. The original CASI of avian cells is the main force to resist external stimuli during the entire developmental stage, a biological characteristic that evolved over a long period of time [21,74]. However, the molecular mechanism of avian CASI still remains unclear, thus warranting further studies.
In recent years, sex research in mammals has suggested that epigenetic regulation plays a pivotal role in governing the expression of sex-related genes [267,268]. Epigenetic marks are reversible modifications that show varying patterns between males and females before and after sex development [269,270,271]. Previous studies have shown that several epigenetic elements are involved in sex development, such as lncRNA, DNA methylation, histone modification, and RNA methylation [107,162,177]. Current research has enabled the establishment of a regulatory relationship between epigenetics and avian sex determination and differentiation. However, the regulatory mechanism of epigenetic modifications on the activation of sex-related genes remains unclear. With the development of new technologies, such as high-throughput chromosome conformation capture (Hi-C), researchers can build three-dimensional chromatin landscapes at different time points of avian sex development and identify sex-related enhancer hubs [272,273]. Those results will help us investigate the 3D regulatory landscape of sex determination and the molecular basis of the transition from an initially biopotential gonad to either alternative fate in birds. Therefore, future sex research should pay more attention to the influence of 3D chromatin interactions on gene expression and look for epigenetic modifications that truly regulate sex-related gene expression in birds.
The ultimate goal of both genetic and epigenetic research in avian sex development is to employ it in the poultry industry and help solve problems, such as the uncertainty of sex-determining factors and barriers to the development of ideal systems, related to sex control [235,236,237]. Regarding gender control technologies, the most important thing is to find a “source of information” that provides reliable information about the sex of embryos as early as possible and cooperate with machines to identify single-sex fertilized eggs. Biomarkers employed by avian sex control have gradually improved from eggshell features and egg odors to biological signals revealed by spectroscopic technologies [25,26,258,259]. Advancements in this process emphasizes the importance of basic research and interdisciplinary communication. In the future, with the development of spectroscopic technologies and the reduction in machine costs, it will be possible to introduce spectroscopic detection equipment into laying hen hatcheries on a large scale and even introduce magnetic resonance imaging technology for early sex detection [22,274]. In addition, the direct impact of gene-editing systems and environmental stimuli on activating and suppressing sex-determining genes provides new perspectives for sex regulation in the poultry industry. Those non-invasive and efficient methods reduce the cost of gender detection and truly achieve one-step gender control in poultry production. However, whether the gender-control technologies that affect gene expression will cause biosafety problems and how to address corresponding ethical issues still need further discussion.

Author Contributions

Conceptualization, X.Z. and J.L.; writing—original draft preparation, X.Z.; writing—review and editing, X.Z., J.L., S.C., N.Y. and J.Z.; funding acquisition, N.Y. and J.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key Research and Development Program of China (2022YFF1000204) and National Key Research and Development Program of China (2022YFD1300100).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tizard, M.L.; Jenkins, K.A.; Cooper, C.A.; Woodcock, M.E.; Challagulla, A.; Doran, T.J. Potential benefits of gene editing for the future of poultry farming. Transgenic Res. 2019, 28, 87–92. [Google Scholar] [CrossRef] [PubMed]
  2. Zhu, R.; Fogelholm, M.; Jalo, E.; Poppitt, S.D.; Silvestre, M.P.; Møller, G.; Huttunen-Lenz, M.; Stratton, G.; Sundvall, J.; Macdonald, I.A.; et al. Animal-based food choice and associations with long-term weight maintenance and metabolic health after a large and rapid weight loss: The PREVIEW study. Clin. Nutr. 2022, 41, 817–828. [Google Scholar] [CrossRef] [PubMed]
  3. Doran, T.J.; Morris, K.R.; Wise, T.G.; O’Neil, T.E.; Cooper, C.A.; Jenkins, K.A.; Tizard, M.L.V. Sex selection in layer chickens. Anim. Prod. Sci. 2018, 58, 476. [Google Scholar] [CrossRef]
  4. Doran, T.J.; Cooper, C.A.; Jenkins, K.A.; Tizard, M.L. Advances in genetic engineering of the avian genome: “Realising the promise”. Transgenic Res. 2016, 25, 307–319. [Google Scholar] [CrossRef]
  5. Sinclair, M.; Zhang, Y.; Descovich, K.; Phillips, C.J.C. Farm Animal Welfare Science in China-A Bibliometric Review of Chinese Literature. Animals 2020, 10, 540. [Google Scholar] [CrossRef]
  6. McColl, K.A.; Clarke, B.; Doran, T.J. Role of genetically engineered animals in future food production. Aust. Vet. J. 2013, 91, 113–117. [Google Scholar] [CrossRef]
  7. Cooper, C.A.; Doran, T.J.; Challagulla, A.; Tizard, M.L.V.; Jenkins, K.A. Innovative approaches to genome editing in avian species. J. Anim. Sci. Biotechnol. 2018, 9, 15. [Google Scholar] [CrossRef]
  8. Hirst, C.E.; Serralbo, O.; Ayers, K.L.; Roeszler, K.N.; Smith, C.A. Genetic Manipulation of the Avian Urogenital System Using In Ovo Electroporation. Methods Mol. Biol. 2017, 1650, 177–190. [Google Scholar] [CrossRef]
  9. Yoshino, T.; Murai, H.; Saito, D. Hedgehog-BMP signalling establishes dorsoventral patterning in lateral plate mesoderm to trigger gonadogenesis in chicken embryos. Nat. Commun. 2016, 7, 12561. [Google Scholar] [CrossRef]
  10. Sekido, R.; Lovell-Badge, R. Mechanisms of gonadal morphogenesis are not conserved between chick and mouse. Dev. Biol. 2007, 302, 132–142. [Google Scholar] [CrossRef]
  11. Guioli, S.; Zhao, D.; Nandi, S.; Clinton, M.; Lovell-Badge, R. Oestrogen in the chick embryo can induce chromosomally male ZZ left gonad epithelial cells to form an ovarian cortex that can support oogenesis. Development 2020, 147, dev181693. [Google Scholar] [CrossRef]
  12. Chue, J.; Smith, C.A. Sex determination and sexual differentiation in the avian model. FEBS J. 2011, 278, 1027–1034. [Google Scholar] [CrossRef] [PubMed]
  13. Smith, C.A.; Roeszler, K.N.; Ohnesorg, T.; Cummins, D.M.; Farlie, P.G.; Doran, T.J.; Sinclair, A.H. The avian Z-linked gene DMRT1 is required for male sex determination in the chicken. Nature 2009, 461, 267–271. [Google Scholar] [CrossRef]
  14. Ioannidis, J.; Taylor, G.; Zhao, D.; Liu, L.; Idoko-Akoh, A.; Gong, D.; Lovell-Badge, R.; Guioli, S.; McGrew, M.J.; Clinton, M. Primary sex determination in birds depends on DMRT1 dosage, but gonadal sex does not determine adult secondary sex characteristics. Proc. Natl. Acad. Sci. USA 2021, 118, e2020909118. [Google Scholar] [CrossRef] [PubMed]
  15. Major, A.T.; Ayers, K.; Chue, J.; Roeszler, K.; Smith, C. FOXL2 antagonises the male developmental pathway in embryonic chicken gonads. J. Endocrinol. 2019, 243, 211–228. [Google Scholar] [CrossRef] [PubMed]
  16. Scheib, D. Effects and role of estrogens in avian gonadal differentiation. Differentiation 1983, 23, S87–S92. [Google Scholar] [CrossRef]
  17. Elbrecht, A.; Smith, R.G. Aromatase enzyme activity and sex determination in chickens. Science 1992, 255, 467–470. [Google Scholar] [CrossRef]
  18. George, F.W.; Wilson, J.D. pathogenesis of the henny feathering trait in the Sebright bantam chicken. Increased conversion of androgen to estrogen in skin. J. Clin. Investig. 1980, 66, 57–65. [Google Scholar] [CrossRef]
  19. Warren, W.C.; Hillier, L.W.; Tomlinson, C.; Minx, P.; Kremitzki, M.; Graves, T.; Markovic, C.; Bouk, N.; Pruitt, K.D.; Thibaud-Nissen, F.; et al. A New Chicken Genome Assembly Provides Insight into Avian Genome Structure. G3 Genes Genomes Genet. 2017, 7, 109–117. [Google Scholar] [CrossRef]
  20. Vaillant, S.; Dorizzi, M.; Pieau, C.; Richard-Mercier, N. Sex reversal and aromatase in chicken. J. Exp. Zool. 2001, 290, 727–740. [Google Scholar] [CrossRef]
  21. Zhao, D.; McBride, D.; Nandi, S.; McQueen, H.A.; McGrew, M.J.; Hocking, P.M.; Lewis, P.D.; Sang, H.M.; Clinton, M. Somatic sex identity is cell autonomous in the chicken. Nature 2010, 464, 237–242. [Google Scholar] [CrossRef] [PubMed]
  22. Galli, R.; Preusse, G.; Uckermann, O.; Bartels, T.; Krautwald-Junghanns, M.E.; Koch, E.; Steiner, G. In ovo sexing of chicken eggs by fluorescence spectroscopy. Anal. Bioanal. Chem. 2017, 409, 1185–1194. [Google Scholar] [CrossRef] [PubMed]
  23. Weissmann, A.; Reitemeier, S.; Hahn, A.; Gottschalk, J.; Einspanier, A. Sexing domestic chicken before hatch: A new method for in ovo gender identification. Theriogenology 2013, 80, 199–205. [Google Scholar] [CrossRef] [PubMed]
  24. Weissmann, A.; Förster, A.; Gottschalk, J.; Reitemeier, S.; Krautwald-Junghanns, M.-E.; Preisinger, R.; Einspanier, A. In ovo-gender identification in laying hen hybrids: Effects on hatching and production performance. Eur. Poult. Sci. 2014, 78, 25. [Google Scholar] [CrossRef]
  25. Yılmaz-Dıkmen, B.; Dikmen, S. A Morphometric Method of Sexing White Layer Eggs. Braz. J. Poult. Sci. 2013, 15, 203–210. [Google Scholar] [CrossRef]
  26. Webster, B.; Hayes, W.; Pike, T.W. Avian egg odour encodes information on embryo sex, fertility and development. PLoS ONE 2015, 10, e0116345. [Google Scholar] [CrossRef] [PubMed]
  27. Chen, Q.; Saatkamp, H.W.; Cortenbach, J.; Jin, W. Comparison of Chinese Broiler Production Systems in Economic Performance and Animal Welfare. Animals 2020, 10, 491. [Google Scholar] [CrossRef]
  28. He, S.; Lin, J.; Jin, Q.; Ma, X.; Liu, Z.; Chen, H.; Ma, J.; Zhang, H.; Descovich, K.; Phillips, C.J.C.; et al. The Relationship between Animal Welfare and Farm Profitability in Cage and Free-Range Housing Systems for Laying Hens in China. Animals 2022, 12, 90. [Google Scholar] [CrossRef]
  29. Ran, F.A.; Hsu, P.D.; Wright, J.; Agarwala, V.; Scott, D.A.; Zhang, F. Genome engineering using the CRISPR-Cas9 system. Nat. Protoc. 2013, 8, 2281–2308. [Google Scholar] [CrossRef]
  30. Gamble, T.; Zarkower, D. Sex determination. Curr. Biol. 2012, 22, R257–R262. [Google Scholar] [CrossRef]
  31. Bachtrog, D.; Mank, J.E.; Peichel, C.L.; Kirkpatrick, M.; Otto, S.P.; Ashman, T.L.; Hahn, M.W.; Kitano, J.; Mayrose, I.; Ming, R.; et al. Sex determination: Why so many ways of doing it? PLoS Biol. 2014, 12, e1001899. [Google Scholar] [CrossRef] [PubMed]
  32. Lang, J.W.; Andrews, H.V. Temperature-dependent sex determination in crocodilians. J. Exp. Zool. 1994, 270, 28–44. [Google Scholar] [CrossRef]
  33. Radder, R.S.; Quinn, A.E.; Georges, A.; Sarre, S.D.; Shine, R. Genetic evidence for co-occurrence of chromosomal and thermal sex-determining systems in a lizard. Biol. Lett. 2008, 4, 176–178. [Google Scholar] [CrossRef] [PubMed]
  34. Pieau, C. Effects of estradiol on the genital apparatus of the embryo of the Mauresque turtle (Testudo graceca L.). Arch. D’anatomie Microsc. Morphol. Exp. 1970, 59, 295–318. [Google Scholar]
  35. Pieau, C. Sex ratio in the embryos of 2 chelonians (Testudo graeca L. and Emys orbicularis L.) born of artificially incubated ova. C R Acad. Hebd. Seances Acad. Sci. D 1971, 272, 3071–3074. [Google Scholar]
  36. Ge, C.; Ye, J.; Zhang, H.; Zhang, Y.; Sun, W.; Sang, Y.; Capel, B.; Qian, G. Dmrt1 induces the male pathway in a turtle species with temperature-dependent sex determination. Development 2017, 144, 2222–2233. [Google Scholar] [CrossRef]
  37. Ge, C.; Ye, J.; Weber, C.; Sun, W.; Zhang, H.; Zhou, Y.; Cai, C.; Qian, G.; Capel, B. The histone demethylase KDM6B regulates temperature-dependent sex determination in a turtle species. Science 2018, 360, 645–648. [Google Scholar] [CrossRef]
  38. Weber, C.; Zhou, Y.; Lee, J.G.; Looger, L.L.; Qian, G.; Ge, C.; Capel, B. Temperature-dependent sex determination is mediated by pSTAT3 repression of Kdm6b. Science 2020, 368, 303–306. [Google Scholar] [CrossRef]
  39. Charnier, M. Action of temperature on the sex ratio in the Agama agama (Agamidae, Lacertilia) embryo. C R Seances Soc. Biol. Fil. 1966, 160, 620–622. [Google Scholar]
  40. Czerwinski, M.; Natarajan, A.; Barske, L.; Looger, L.L.; Capel, B. A timecourse analysis of systemic and gonadal effects of temperature on sexual development of the red-eared slider turtle Trachemys scripta elegans. Dev. Biol. 2016, 420, 166–177. [Google Scholar] [CrossRef]
  41. Livernois, A.M.; Graves, J.A.; Waters, P.D. The origin and evolution of vertebrate sex chromosomes and dosage compensation. Heredity 2012, 108, 50–58. [Google Scholar] [CrossRef] [PubMed]
  42. Bellott, D.W.; Hughes, J.F.; Skaletsky, H.; Brown, L.G.; Pyntikova, T.; Cho, T.J.; Koutseva, N.; Zaghlul, S.; Graves, T.; Rock, S.; et al. Mammalian Y chromosomes retain widely expressed dosage-sensitive regulators. Nature 2014, 508, 494–499. [Google Scholar] [CrossRef] [PubMed]
  43. Larney, C.; Bailey, T.L.; Koopman, P. Switching on sex: Transcriptional regulation of the testis-determining gene Sry. Development 2014, 141, 2195–2205. [Google Scholar] [CrossRef] [PubMed]
  44. Koopman, P.; Gubbay, J.; Vivian, N.; Goodfellow, P.; Lovell-Badge, R. Male development of chromosomally female mice transgenic for Sry. Nature 1991, 351, 117–121. [Google Scholar] [CrossRef]
  45. Koopman, P.; Munsterberg, A.; Capel, B.; Vivian, N.; Lovell-Badge, R. Expression of a candidate sex-determining gene during mouse testis differentiation. Nature 1990, 348, 450–452. [Google Scholar] [CrossRef]
  46. Sinclair, A.H.; Berta, P.; Palmer, M.S.; Hawkins, J.R.; Griffiths, B.L.; Smith, M.J.; Foster, J.W.; Frischauf, A.M.; Lovell-Badge, R.; Goodfellow, P.N. A gene from the human sex-determining region encodes a protein with homology to a conserved DNA-binding motif. Nature 1990, 346, 240–244. [Google Scholar] [CrossRef] [PubMed]
  47. Terao, M.; Ogawa, Y.; Takada, S.; Kajitani, R.; Okuno, M.; Mochimaru, Y.; Matsuoka, K.; Itoh, T.; Toyoda, A.; Kono, T.; et al. Turnover of mammal sex chromosomes in the Sry-deficient Amami spiny rat is due to male-specific upregulation of Sox9. Proc. Natl. Acad. Sci. USA 2022, 119, e2211574119. [Google Scholar] [CrossRef]
  48. Marshall Graves, J.A.; Shetty, S. Sex from W to Z: Evolution of vertebrate sex chromosomes and sex determining genes. J. Exp. Zool. 2001, 290, 449–462. [Google Scholar] [CrossRef]
  49. Davis, J.K.; Thomas, P.J.; Program, N.C.S.; Thomas, J.W. A W-linked palindrome and gene conversion in New World sparrows and blackbirds. Chromosome Res. 2010, 18, 543–553. [Google Scholar] [CrossRef]
  50. Moghadam, H.K.; Pointer, M.A.; Wright, A.E.; Berlin, S.; Mank, J.E. W chromosome expression responds to female-specific selection. Proc. Natl. Acad. Sci. USA 2012, 109, 8207–8211. [Google Scholar] [CrossRef]
  51. Smeds, L.; Warmuth, V.; Bolivar, P.; Uebbing, S.; Burri, R.; Suh, A.; Nater, A.; Bures, S.; Garamszegi, L.Z.; Hogner, S.; et al. Evolutionary analysis of the female-specific avian W chromosome. Nat. Commun. 2015, 6, 7330. [Google Scholar] [CrossRef] [PubMed]
  52. Zhou, R.; Macaya-Sanz, D.; Carlson, C.H.; Schmutz, J.; Jenkins, J.W.; Kudrna, D.; Sharma, A.; Sandor, L.; Shu, S.; Barry, K.; et al. A willow sex chromosome reveals convergent evolution of complex palindromic repeats. Genome Biol. 2020, 21, 38. [Google Scholar] [CrossRef] [PubMed]
  53. Smith, C.A. Sex determination in birds: HINTs from the W sex chromosome? Sex. Dev. 2007, 1, 279–285. [Google Scholar] [CrossRef] [PubMed]
  54. Otto, S.P.; Gerstein, A.C. Why have sex? The population genetics of sex and recombination. Biochem. Soc. Trans. 2006, 34, 519–522. [Google Scholar] [CrossRef]
  55. Furman, B.L.S.; Metzger, D.C.H.; Darolti, I.; Wright, A.E.; Sandkam, B.A.; Almeida, P.; Shu, J.J.; Mank, J.E. Sex Chromosome Evolution: So Many Exceptions to the Rules. Genome Biol. Evol. 2020, 12, 750–763. [Google Scholar] [CrossRef] [PubMed]
  56. Charlesworth, B. The evolution of sex chromosomes. Science 1991, 251, 1030–1033. [Google Scholar] [CrossRef] [PubMed]
  57. Bergero, R.; Charlesworth, D. The evolution of restricted recombination in sex chromosomes. Trends Ecol. Evol. 2009, 24, 94–102. [Google Scholar] [CrossRef]
  58. Handley, L.J.; Ceplitis, H.; Ellegren, H. Evolutionary strata on the chicken Z chromosome: Implications for sex chromosome evolution. Genetics 2004, 167, 367–376. [Google Scholar] [CrossRef]
  59. Graves, J.A. The epigenetic sole of sex and dosage compensation. Nat. Genet. 2014, 46, 215–217. [Google Scholar] [CrossRef]
  60. Smith, C.A.; Roeszler, K.N.; Sinclair, A.H. Genetic evidence against a role for W-linked histidine triad nucleotide binding protein (HINTW) in avian sex determination. Int. J. Dev. Biol. 2009, 53, 59–67. [Google Scholar] [CrossRef]
  61. Reed, K.J.; Sinclair, A.H. RETRACTED: FET-1: A novel W-linked, female specific gene up-regulated in the embryonic chicken ovary. Gene Expr. Patterns 2002, 2, 83–86. [Google Scholar] [CrossRef]
  62. Sun, C.; Jin, K.; Zhou, J.; Zuo, Q.; Song, J.; Yani, Z.; Chen, G.; Li, B. Role and function of the Hintw in early sex differentiation in chicken (Gallus gallus) embryo. Anim. Biotechnol. 2021, 34, 56–66. [Google Scholar] [CrossRef]
  63. Nagai, H.; Sezaki, M.; Bertocchini, F.; Fukuda, K.; Sheng, G. HINTW, a W-chromosome HINT gene in chick, is expressed ubiquitously and is a robust female cell marker applicable in intraspecific chimera studies. Genesis 2014, 52, 424–430. [Google Scholar] [CrossRef] [PubMed]
  64. Arit, D.; Bensch, S.; Hansson, B.; Hasselquist, D.; Westerdahl, H. Observation of a ZZW female in a natural population: Implications for avian sex determination. Proc. Biol. Sci. 2004, 271 (Suppl. 4), S249–S251. [Google Scholar] [CrossRef]
  65. Kupper, C.; Augustin, J.; Edwards, S.; Szekely, T.; Kosztolanyi, A.; Burke, T.; Janes, D.E. Triploid plover female provides support for a role of the W chromosome in avian sex determination. Biol. Lett. 2012, 8, 787–789. [Google Scholar] [CrossRef]
  66. Thorne, M.H.; Collins, R.K.; Sheldon, B.L. Triploidy and other chromosomal abnormalities in a selected line of chickens. Genet. Sel. Evol. 1991, 23, S212. [Google Scholar] [CrossRef]
  67. Borsani, G.; Tonlorenzi, R.; Simmler, M.C.; Dandolo, L.; Arnaud, D.; Capra, V.; Grompe, M.; Pizzuti, A.; Muzny, D.; Lawrence, C.; et al. Characterization of a murine gene expressed from the inactive X chromosome. Nature 1991, 351, 325–329. [Google Scholar] [CrossRef]
  68. Brockdorff, N.; Ashworth, A.; Kay, G.F.; Cooper, P.; Smith, S.; McCabe, V.M.; Norris, D.P.; Penny, G.D.; Patel, D.; Rastan, S. Conservation of position and exclusive expression of mouse Xist from the inactive X chromosome. Nature 1991, 351, 329–331. [Google Scholar] [CrossRef] [PubMed]
  69. Aguilar, R.; Spencer, K.B.; Kesner, B.; Rizvi, N.F.; Badmalia, M.D.; Mrozowich, T.; Mortison, J.D.; Rivera, C.; Smith, G.F.; Burchard, J.; et al. Targeting Xist with compounds that disrupt RNA structure and X inactivation. Nature 2022, 604, 160–166. [Google Scholar] [CrossRef] [PubMed]
  70. Wang, Q.; Mank, J.E.; Li, J.; Yang, N.; Qu, L. Allele-Specific Expression Analysis Does Not Support Sex Chromosome Inactivation on the Chicken Z Chromosome. Genome Biol. Evol. 2017, 9, 619–626. [Google Scholar] [CrossRef]
  71. McQueen, H.A.; McBride, D.; Miele, G.; Bird, A.P.; Clinton, M. Dosage compensation in birds. Curr. Biol. 2001, 11, 253–257. [Google Scholar] [CrossRef] [PubMed]
  72. Itoh, Y.; Melamed, E.; Yang, X.; Kampf, K.; Wang, S.; Yehya, N.; Van Nas, A.; Replogle, K.; Band, M.R.; Clayton, D.F.; et al. Dosage compensation is less effective in birds than in mammals. J. Biol. 2007, 6, 2. [Google Scholar] [CrossRef]
  73. Soler, L.; Alves, S.; Brionne, A.; Jacques, A.; Guerin, V.; Cherif-Feildel, M.; Combes-Soia, L.; Fouchecourt, S.; Thelie, A.; Blesbois, E.; et al. Protein expression reveals a molecular sexual identity of avian primordial germ cells at pre-gonadal stages. Sci. Rep. 2021, 11, 19236. [Google Scholar] [CrossRef] [PubMed]
  74. Clinton, M.; Zhao, D.; Nandi, S.; McBride, D. Evidence for avian cell autonomous sex identity (CASI) and implications for the sex-determination process? Chromosome Res. 2012, 20, 177–190. [Google Scholar] [CrossRef] [PubMed]
  75. Friedrich, S.R.; Nevue, A.A.; Andrade, A.L.P.; Velho, T.A.F.; Mello, C.V. Emergence of sex-specific transcriptomes in a sexually dimorphic brain nucleus. Cell. Rep. 2022, 40, 111152. [Google Scholar] [CrossRef] [PubMed]
  76. Agate, R.J.; Grisham, W.; Wade, J.; Mann, S.; Wingfield, J.; Schanen, C.; Palotie, A.; Arnold, A.P. Neural, not gonadal, origin of brain sex differences in a gynandromorphic finch. Proc. Natl. Acad. Sci. USA 2003, 100, 4873–4878. [Google Scholar] [CrossRef]
  77. Duan, X.; Jia, X.; Liang, K.; Huang, F.; Shan, J.; Chen, H.; Ruan, X.; Li, L.; Zhao, H.; Wang, Q. Liposome-Encapsulated Rec8 and Dmrt1 Plasmids Induce Red-Spotted Grouper (Epinephelus akaara) Testis Maturation. Mar. Biotechnol. 2022, 24, 345–353. [Google Scholar] [CrossRef]
  78. Kulibin, A.Y.; Malolina, E.A. Formation of the rete testis during mouse embryonic development. Dev. Dyn. 2020, 249, 1486–1499. [Google Scholar] [CrossRef]
  79. Panara, V.; Budd, G.E.; Janssen, R. Phylogenetic analysis and embryonic expression of panarthropod Dmrt genes. Front. Zool. 2019, 16, 23. [Google Scholar] [CrossRef]
  80. Wang, L.; Sun, F.; Wan, Z.Y.; Yang, Z.; Tay, Y.X.; Lee, M.; Ye, B.; Wen, Y.; Meng, Z.; Fan, B.; et al. Transposon-induced epigenetic silencing in the X chromosome as a novel form of dmrt1 expression regulation during sex determination in the fighting fish. BMC Biol. 2022, 20, 5. [Google Scholar] [CrossRef]
  81. Yoshimoto, S.; Okada, E.; Umemoto, H.; Tamura, K.; Uno, Y.; Nishida-Umehara, C.; Matsuda, Y.; Takamatsu, N.; Shiba, T.; Ito, M. A W-linked DM-domain gene, DM-W, participates in primary ovary development in Xenopus laevis. Proc. Natl. Acad. Sci. USA 2008, 105, 2469–2474. [Google Scholar] [CrossRef] [PubMed]
  82. Hayashi, S.; Suda, K.; Fujimura, F.; Fujikawa, M.; Tamura, K.; Tsukamoto, D.; Evans, B.J.; Takamatsu, N.; Ito, M. Neofunctionalization of a Noncoding Portion of a DNA Transposon in the Coding Region of the Chimerical Sex-Determining Gene dm-W in Xenopus Frogs. Mol. Biol. Evol. 2022, 39, msac138. [Google Scholar] [CrossRef] [PubMed]
  83. Yoshimoto, S.; Ikeda, N.; Izutsu, Y.; Shiba, T.; Takamatsu, N.; Ito, M. Opposite roles of DMRT1 and its W-linked paralogue, DM-W, in sexual dimorphism of Xenopus laevis: Implications of a ZZ/ZW-type sex-determining system. Development 2010, 137, 2519–2526. [Google Scholar] [CrossRef] [PubMed]
  84. Zhou, H.; Whitworth, C.; Pozmanter, C.; Neville, M.C.; Van Doren, M. Doublesex regulates fruitless expression to promote sexual dimorphism of the gonad stem cell niche. PLoS Genet. 2021, 17, e1009468. [Google Scholar] [CrossRef]
  85. Steinmann-Zwicky, M. Sex determination of the Drosophila germ line: Tra and dsx control somatic inductive signals. Development 1994, 120, 707–716. [Google Scholar] [CrossRef]
  86. Shen, M.M.; Hodgkin, J. mab-3, a gene required for sex-specific yolk protein expression and a male-specific lineage in C. elegans. Cell 1988, 54, 1019–1031. [Google Scholar] [CrossRef] [PubMed]
  87. Nanda, I.; Kondo, M.; Hornung, U.; Asakawa, S.; Winkler, C.; Shimizu, A.; Shan, Z.; Haaf, T.; Shimizu, N.; Shima, A.; et al. A duplicated copy of DMRT1 in the sex-determining region of the Y chromosome of the medaka, Oryzias latipes. Proc. Natl. Acad. Sci. USA 2002, 99, 11778–11783. [Google Scholar] [CrossRef] [PubMed]
  88. Matsuda, M.; Shinomiya, A.; Kinoshita, M.; Suzuki, A.; Kobayashi, T.; Paul-Prasanth, B.; Lau, E.L.; Hamaguchi, S.; Sakaizumi, M.; Nagahama, Y. DMY gene induces male development in genetically female (XX) medaka fish. Proc. Natl. Acad. Sci. USA 2007, 104, 3865–3870. [Google Scholar] [CrossRef] [PubMed]
  89. Matsuda, M.; Nagahama, Y.; Shinomiya, A.; Sato, T.; Matsuda, C.; Kobayashi, T.; Morrey, C.E.; Shibata, N.; Asakawa, S.; Shimizu, N.; et al. DMY is a Y-specific DM-domain gene required for male development in the medaka fish. Nature 2002, 417, 559–563. [Google Scholar] [CrossRef]
  90. Ogita, Y.; Mawaribuchi, S.; Nakasako, K.; Tamura, K.; Matsuda, M.; Katsumura, T.; Oota, H.; Watanabe, G.; Yoneda, S.; Takamatsu, N.; et al. Parallel Evolution of Two dmrt1-Derived Genes, dmy and dm-W, for Vertebrate Sex Determination. iScience 2020, 23, 100757. [Google Scholar] [CrossRef]
  91. Smith, C.A.; Katz, M.; Sinclair, A.H. DMRT1 is upregulated in the gonads during female-to-male sex reversal in ZW chicken embryos. Biol. Reprod. 2003, 68, 560–570. [Google Scholar] [CrossRef] [PubMed]
  92. Zhao, Y.; Lu, H.; Yu, H.; Cheng, H.; Zhou, R. Multiple alternative splicing in gonads of chicken DMRT1. Dev. Genes Evol. 2007, 217, 119–126. [Google Scholar] [CrossRef] [PubMed]
  93. Clinton, M.; Zhao, D. Avian sex determination: A chicken egg conundrum. Sex. Dev. 2023. [Google Scholar] [CrossRef] [PubMed]
  94. Yang, Y.; Gong, P.; Feng, Y.P.; Li, S.J.; Peng, X.L.; Ran, Z.P.; Qian, Y.G.; Gong, Y.Z. Temporospatial expression of Dmrt1 in chicken urogenital system (Gallus gallus) using whole mount in situ hybridization. Acta Biol. Hung. 2013, 64, 161–168. [Google Scholar] [CrossRef]
  95. Smith, C.A.; McClive, P.J.; Western, P.S.; Reed, K.J.; Sinclair, A.H. Conservation of a sex-determining gene. Nature 1999, 402, 601–602. [Google Scholar] [CrossRef]
  96. Yoshioka, H.; Ishimaru, Y.; Sugiyama, N.; Tsunekawa, N.; Noce, T.; Kasahara, M.; Morohashi, K. Mesonephric FGF signaling is associated with the development of sexually indifferent gonadal primordium in chick embryos. Dev. Biol. 2005, 280, 150–161. [Google Scholar] [CrossRef]
  97. Shan, Z.; Nanda, I.; Wang, Y.; Schmid, M.; Vortkamp, A.; Haaf, T. Sex-specific expression of an evolutionarily conserved male regulatory gene, DMRT1, in birds. Cytogenet. Cell. Genet. 2000, 89, 252–257. [Google Scholar] [CrossRef]
  98. Lambeth, L.S.; Raymond, C.S.; Roeszler, K.N.; Kuroiwa, A.; Nakata, T.; Zarkower, D.; Smith, C.A. Over-expression of DMRT1 induces the male pathway in embryonic chicken gonads. Dev. Biol. 2014, 389, 160–172. [Google Scholar] [CrossRef]
  99. Omotehara, T.; Smith, C.A.; Mantani, Y.; Kobayashi, Y.; Tatsumi, A.; Nagahara, D.; Hashimoto, R.; Hirano, T.; Umemura, Y.; Yokoyama, T.; et al. Spatiotemporal expression patterns of doublesex and mab-3 related transcription factor 1 in the chicken developing gonads and Mullerian ducts. Poult. Sci. 2014, 93, 953–958. [Google Scholar] [CrossRef]
  100. Lee, H.J.; Seo, M.; Choi, H.J.; Rengaraj, D.; Jung, K.M.; Park, J.S.; Lee, K.Y.; Kim, Y.M.; Park, K.J.; Han, S.T.; et al. DMRT1 gene disruption alone induces incomplete gonad feminization in chicken. FASEB J. 2021, 35, e21876. [Google Scholar] [CrossRef]
  101. Estermann, M.A.; Williams, S.; Hirst, C.E.; Roly, Z.Y.; Serralbo, O.; Adhikari, D.; Powell, D.; Major, A.T.; Smith, C.A. Insights into Gonadal Sex Differentiation Provided by Single-Cell Transcriptomics in the Chicken Embryo. Cell. Rep. 2020, 31, 107491. [Google Scholar] [CrossRef] [PubMed]
  102. Estermann, M.A.; Major, A.T.; Smith, C.A. Genetic Regulation of Avian Testis Development. Genes 2021, 12, 1459. [Google Scholar] [CrossRef] [PubMed]
  103. Kuroda, Y.; Arai, N.; Arita, M.; Teranishi, M.; Hori, T.; Harata, M.; Mizuno, S. Absence of Z-chromosome inactivation for five genes in male chickens. Chromosome Res. 2001, 9, 457–468. [Google Scholar] [CrossRef] [PubMed]
  104. Teranishi, M.; Shimada, Y.; Hori, T.; Nakabayashi, O.; Kikuchi, T.; Macleod, T.; Pym, R.; Sheldon, B.; Solovei, I.; Macgregor, H.; et al. Transcripts of the MHM region on the chicken Z chromosome accumulate as non-coding RNA in the nucleus of female cells adjacent to the DMRT1 locus. Chromosome Res. 2001, 9, 147–165. [Google Scholar] [CrossRef] [PubMed]
  105. Itoh, Y.; Replogle, K.; Kim, Y.H.; Wade, J.; Clayton, D.F.; Arnold, A.P. Sex bias and dosage compensation in the zebra finch versus chicken genomes: General and specialized patterns among birds. Genome Res. 2010, 20, 512–518. [Google Scholar] [CrossRef]
  106. Yang, X.; Deng, J.; Zheng, J.; Xia, L.; Yang, Z.; Qu, L.; Chen, S.; Xu, G.; Jiang, H.; Clinton, M.; et al. A Window of MHM Demethylation Correlates with Key Events in Gonadal Differentiation in the Chicken. Sex. Dev. 2016, 10, 152–158. [Google Scholar] [CrossRef]
  107. Yang, X.; Zheng, J.; Qu, L.; Chen, S.; Li, J.; Xu, G.; Yang, N. Methylation status of cMHM and expression of sex-specific genes in adult sex-reversed female chickens. Sex. Dev. 2011, 5, 147–154. [Google Scholar] [CrossRef]
  108. Yang, X.; Zheng, J.; Xu, G.; Qu, L.; Chen, S.; Li, J.; Yang, N. Exogenous cMHM regulates the expression of DMRT1 and ER alpha in avian testes. Mol. Biol. Rep. 2010, 37, 1841–1847. [Google Scholar] [CrossRef]
  109. Bisoni, L.; Batlle-Morera, L.; Bird, A.P.; Suzuki, M.; McQueen, H.A. Female-specific hyperacetylation of histone H4 in the chicken Z chromosome. Chromosome Res. 2005, 13, 205–214. [Google Scholar] [CrossRef]
  110. Briggs, S.F.; Reijo Pera, R.A. X chromosome inactivation: Recent advances and a look forward. Curr. Opin. Genet. Dev. 2014, 28, 78–82. [Google Scholar] [CrossRef]
  111. Mulvey, B.B.; Olcese, U.; Cabrera, J.R.; Horabin, J.I. An interactive network of long non-coding RNAs facilitates the Drosophila sex determination decision. Biochim. Biophys. Acta 2014, 1839, 773–784. [Google Scholar] [CrossRef]
  112. Roeszler, K.N.; Itman, C.; Sinclair, A.H.; Smith, C.A. The long non-coding RNA, MHM, plays a role in chicken embryonic development, including gonadogenesis. Dev. Biol. 2012, 366, 317–326. [Google Scholar] [CrossRef] [PubMed]
  113. Jiang, Y.; Peng, Z.; Man, Q.; Wang, S.; Huang, X.; Meng, L.; Wang, H.; Zhu, G. H3K27ac chromatin acetylation and gene expression analysis reveal sex- and situs-related differences in developing chicken gonads. Biol. Sex. Differ. 2022, 13, 6. [Google Scholar] [CrossRef] [PubMed]
  114. Graves, J.A. Evolution of vertebrate sex chromosomes and dosage compensation. Nat. Rev. Genet. 2016, 17, 33–46. [Google Scholar] [CrossRef] [PubMed]
  115. Estermann, M.A.; Smith, C.A. Applying Single-Cell Analysis to Gonadogenesis and DSDs (Disorders/Differences of Sex Development). Int. J. Mol. Sci. 2020, 21, 6614. [Google Scholar] [CrossRef]
  116. Hartady, T.; Syamsunarno, M.; Priosoeryanto, B.P.; Jasni, S.; Balia, R.L. Review of herbal medicine works in the avian species. Vet. World 2021, 14, 2889–2906. [Google Scholar] [CrossRef]
  117. Hamburger, V.; Hamilton, H.L. A series of normal stages in the development of the chick embryo. J. Morphol. 1951, 88, 49–92. [Google Scholar] [CrossRef]
  118. Ariza, L.; Carmona, R.; Canete, A.; Cano, E.; Munoz-Chapuli, R. Coelomic epithelium-derived cells in visceral morphogenesis. Dev. Dyn. 2016, 245, 307–322. [Google Scholar] [CrossRef]
  119. Guioli, S.; Nandi, S.; Zhao, D.; Burgess-Shannon, J.; Lovell-Badge, R.; Clinton, M. Gonadal asymmetry and sex determination in birds. Sex. Dev. 2014, 8, 227–242. [Google Scholar] [CrossRef]
  120. Yoshino, T.; Saito, D. Epithelial-to-mesenchymal transition-based morphogenesis of dorsal mesentery and gonad. Semin. Cell. Dev. Biol. 2019, 92, 105–112. [Google Scholar] [CrossRef]
  121. Stevant, I.; Neirijnck, Y.; Borel, C.; Escoffier, J.; Smith, L.B.; Antonarakis, S.E.; Dermitzakis, E.T.; Nef, S. Deciphering Cell Lineage Specification during Male Sex Determination with Single-Cell RNA Sequencing. Cell. Rep. 2018, 22, 1589–1599. [Google Scholar] [CrossRef] [PubMed]
  122. Nef, S.; Stevant, I.; Greenfield, A. Characterizing the bipotential mammalian gonad. Curr. Top. Dev. Biol. 2019, 134, 167–194. [Google Scholar] [CrossRef] [PubMed]
  123. Albrecht, K.H.; Eicher, E.M. Evidence that Sry is expressed in pre-Sertoli cells and Sertoli and granulosa cells have a common precursor. Dev. Biol. 2001, 240, 92–107. [Google Scholar] [CrossRef] [PubMed]
  124. Ginsburg, M.; Eyal-Giladi, H. Primordial germ cells of the young chick blastoderm originate from the central zone of the area pellucida irrespective of the embryo-forming process. Development 1987, 101, 209–219. [Google Scholar] [CrossRef]
  125. Smith, C.A.; Sinclair, A.H. Sex determination: Insights from the chicken. Bioessays 2004, 26, 120–132. [Google Scholar] [CrossRef]
  126. Swift, C.H. Origin of the sex-cords and definitive spermatogonia in the male chick. Am. J. Anat. 1916, 20, 375–410. [Google Scholar] [CrossRef]
  127. Smith, C.A.; Roeszler, K.N.; Hudson, Q.J.; Sinclair, A.H. Avian sex determination: What, when and where? Cytogenet. Genome Res. 2007, 117, 165–173. [Google Scholar] [CrossRef]
  128. Gonzalez-Moran, M.G. Histological and stereological changes in growing and regressing chicken ovaries during development. Anat. Rec. 2011, 294, 893–904. [Google Scholar] [CrossRef]
  129. Intarapat, S.; Stern, C.D. Sexually dimorphic and sex-independent left-right asymmetries in chicken embryonic gonads. PLoS ONE 2013, 8, e69893. [Google Scholar] [CrossRef]
  130. Carlon, N.; Stahl, A. Origin of the somatic components in chick embryonic gonads. Arch. D’anatomie Microsc. Morphol. Exp. 1985, 74, 52–59. [Google Scholar]
  131. de Melo Bernardo, A.; Heeren, A.M.; van Iperen, L.; Fernandes, M.G.; He, N.; Anjie, S.; Noce, T.; Ramos, E.S.; de Sousa Lopes, S.M. Meiotic wave adds extra asymmetry to the development of female chicken gonads. Mol. Reprod. Dev. 2015, 82, 774–786. [Google Scholar] [CrossRef] [PubMed]
  132. Gasc, J.M.; Stumpf, W.E. Sexual differentiation of the urogenital tract in the chicken embryo: Autoradiographic localization of sex-steroid target cells during development. J. Embryol. Exp. Morphol. 1981, 63, 207–223. [Google Scholar] [CrossRef] [PubMed]
  133. Roly, Z.Y.; Backhouse, B.; Cutting, A.; Tan, T.Y.; Sinclair, A.H.; Ayers, K.L.; Major, A.T.; Smith, C.A. The cell biology and molecular genetics of Mullerian duct development. Wiley Interdiscip. Rev. Dev. Biol. 2018, 7, e310. [Google Scholar] [CrossRef] [PubMed]
  134. Roly, Z.Y.; Godini, R.; Estermann, M.A.; Major, A.T.; Pocock, R.; Smith, C.A. Transcriptional landscape of the embryonic chicken Mullerian duct. BMC Genom. 2020, 21, 688. [Google Scholar] [CrossRef] [PubMed]
  135. Major, A.T.; Smith, C.A. Sex Reversal in Birds. Sex. Dev. 2016, 10, 288–300. [Google Scholar] [CrossRef]
  136. Hutson, J.; Ikawa, H.; Donahoe, P. The ontogeny of mullerian inhibiting substance in the gonads of the chicken. J. Pediatr. Surg. 1981, 16, 822–827. [Google Scholar] [CrossRef]
  137. Oreal, E.; Pieau, C.; Mattei, M.-G.; Josso, N.; Picard, J.-Y.; Carré-Eusèbe, D.; Magre, S. Early expression ofAMH in chicken embryonic gonads precedes testicularSOX9 expression. Dev. Dyn. 1998, 212, 522–532. [Google Scholar] [CrossRef]
  138. Josso, N.; Picard, J.Y. Anti-Mullerian hormone. Physiol. Rev. 1986, 66, 1038–1090. [Google Scholar] [CrossRef]
  139. Lambeth, L.S.; Smith, C.A. Disorders of sexual development in poultry. Sex. Dev. 2012, 6, 96–103. [Google Scholar] [CrossRef]
  140. Tran, D.; Josso, N. Relationship between avian and mammalian anti-Mulllerian hormones. Biol. Reprod. 1977, 16, 267–273. [Google Scholar] [CrossRef]
  141. Hutson, J.M.; Ikawa, H.; Donahoe, P.K. Estrogen inhibition of mullerian inhibiting substance in the chick embryo. J. Pediatr. Surg. 1982, 17, 953–959. [Google Scholar] [CrossRef] [PubMed]
  142. Dohr, G.; Tarmann, T. Contacts between Wolffian and Mullerian cells at the tip of the outgrowing Mullerian duct in rat embryos. Acta Anat. 1984, 120, 123–128. [Google Scholar] [CrossRef] [PubMed]
  143. Gruenwald, P. The relation of the growing müllerian duct to the wolffian duct and its importance for the genesis of malformations. Anat. Rec. 1941, 81, 1–19. [Google Scholar] [CrossRef]
  144. Atsuta, Y.; Takahashi, Y. Early formation of the Mullerian duct is regulated by sequential actions of BMP/Pax2 and FGF/Lim1 signaling. Development 2016, 143, 3549–3559. [Google Scholar] [CrossRef] [PubMed]
  145. Carroll, T.J.; Park, J.S.; Hayashi, S.; Majumdar, A.; McMahon, A.P. Wnt9b plays a central role in the regulation of mesenchymal to epithelial transitions underlying organogenesis of the mammalian urogenital system. Dev. Cell. 2005, 9, 283–292. [Google Scholar] [CrossRef]
  146. Chiga, M.; Ohmori, T.; Ohba, T.; Katabuchi, H.; Nishinakamura, R. Preformed Wolffian duct regulates Mullerian duct elongation independently of canonical Wnt signaling or Lhx1 expression. Int. J. Dev. Biol. 2014, 58, 663–668. [Google Scholar] [CrossRef]
  147. Lefebvre, V.; Smits, P. Transcriptional control of chondrocyte fate and differentiation. Birth Defects Res. C Embryo Today 2005, 75, 200–212. [Google Scholar] [CrossRef]
  148. Nishimura, R.; Hata, K.; Ikeda, F.; Ichida, F.; Shimoyama, A.; Matsubara, T.; Wada, M.; Amano, K.; Yoneda, T. Signal transduction and transcriptional regulation during mesenchymal cell differentiation. J. Bone Min. Metab. 2008, 26, 203–212. [Google Scholar] [CrossRef]
  149. Bridgewater, L.C.; Walker, M.D.; Miller, G.C.; Ellison, T.A.; Holsinger, L.D.; Potter, J.L.; Jackson, T.L.; Chen, R.K.; Winkel, V.L.; Zhang, Z.; et al. Adjacent DNA sequences modulate Sox9 transcriptional activation at paired Sox sites in three chondrocyte-specific enhancer elements. Nucleic Acids Res. 2003, 31, 1541–1553. [Google Scholar] [CrossRef]
  150. Takada, S.; Ota, J.; Kansaku, N.; Yamashita, H.; Izumi, T.; Ishikawa, M.; Wada, T.; Kaneda, R.; Choi, Y.L.; Koinuma, K.; et al. Nucleotide sequence and embryonic expression of quail and duck Sox9 genes. Gen. Comp. Endocrinol. 2006, 145, 208–213. [Google Scholar] [CrossRef]
  151. Kent, J.; Wheatley, S.C.; Andrews, J.E.; Sinclair, A.H.; Koopman, P. A male-specific role for SOX9 in vertebrate sex determination. Development 1996, 122, 2813–2822. [Google Scholar] [CrossRef] [PubMed]
  152. Morais da Silva, S.; Hacker, A.; Harley, V.; Goodfellow, P.; Swain, A.; Lovell-Badge, R. Sox9 expression during gonadal development implies a conserved role for the gene in testis differentiation in mammals and birds. Nat. Genet. 1996, 14, 62–68. [Google Scholar] [CrossRef] [PubMed]
  153. Vidal, V.P.; Chaboissier, M.C.; de Rooij, D.G.; Schedl, A. Sox9 induces testis development in XX transgenic mice. Nat. Genet. 2001, 28, 216–217. [Google Scholar] [CrossRef] [PubMed]
  154. Yamashita, S.; Kataoka, K.; Yamamoto, H.; Kato, T.; Hara, S.; Yamaguchi, K.; Renard-Guillet, C.; Katou, Y.; Shirahige, K.; Ochi, H.; et al. Comparative analysis demonstrates cell type-specific conservation of SOX9 targets between mouse and chicken. Sci. Rep. 2019, 9, 12560. [Google Scholar] [CrossRef]
  155. Smith, C.A.; Smith, M.J.; Sinclair, A.H. Gene expression during gonadogenesis in the chicken embryo. Gene 1999, 234, 395–402. [Google Scholar] [CrossRef]
  156. Lambeth, L.S.; Morris, K.; Ayers, K.L.; Wise, T.G.; O’Neil, T.; Wilson, S.; Cao, Y.; Sinclair, A.H.; Cutting, A.D.; Doran, T.J.; et al. Overexpression of Anti-Mullerian Hormone Disrupts Gonadal Sex Differentiation, Blocks Sex Hormone Synthesis, and Supports Cell Autonomous Sex Development in the Chicken. Endocrinology 2016, 157, 1258–1275. [Google Scholar] [CrossRef]
  157. Nakata, T.; Ishiguro, M.; Aduma, N.; Izumi, H.; Kuroiwa, A. Chicken hemogen homolog is involved in the chicken-specific sex-determining mechanism. Proc. Natl. Acad. Sci. USA 2013, 110, 3417–3422. [Google Scholar] [CrossRef]
  158. Yang, L.V.; Nicholson, R.H.; Kaplan, J.; Galy, A.; Li, L. Hemogen is a novel nuclear factor specifically expressed in mouse hematopoietic development and its human homologue EDAG maps to chromosome 9q22, a region containing breakpoints of hematological neoplasms. Mech. Dev. 2001, 104, 105–111. [Google Scholar] [CrossRef]
  159. Chen, C.; Zhou, S.; Lian, Z.; Jiang, J.; Gao, X.; Hu, C.; Zuo, Q.; Zhang, Y.; Chen, G.; Jin, K.; et al. Tle4z1 Facilitate the Male Sexual Differentiation of Chicken Embryos. Front. Physiol. 2022, 13, 856980. [Google Scholar] [CrossRef]
  160. Jiang, J.; Zhang, C.; Yuan, X.; Li, J.; Zhang, M.; Shi, X.; Jin, K.; Zhang, Y.; Zuo, Q.; Chen, G.; et al. Spin1z induces the male pathway in the chicken by down-regulating Tcf4. Gene 2021, 780, 145521. [Google Scholar] [CrossRef]
  161. Zhang, M.; Xu, P.; Sun, X.; Zhang, C.; Shi, X.; Li, J.; Jiang, J.; Chen, C.; Zhang, Y.; Chen, G.; et al. JUN promotes chicken female differentiation by inhibiting Smad2. Cytotechnology 2021, 73, 101–113. [Google Scholar] [CrossRef] [PubMed]
  162. Li, J.; Zhang, X.; Wang, X.; Sun, C.; Zheng, J.; Li, J.; Yi, G.; Yang, N. The m6A methylation regulates gonadal sex differentiation in chicken embryo. J. Anim. Sci. Biotechnol. 2022, 13, 52. [Google Scholar] [CrossRef] [PubMed]
  163. Zhang, X.; Li, J.; Wang, X.; Jie, Y.; Sun, C.; Zheng, J.; Li, J.; Yang, N.; Chen, S. ATAC-seq and RNA-seq analysis unravel the mechanism of sex differentiation and infertility in sex reversal chicken. Epigenetics Chromatin 2023, 16, 2. [Google Scholar] [CrossRef] [PubMed]
  164. Govoroun, M.S.; Pannetier, M.; Pailhoux, E.; Cocquet, J.; Brillard, J.P.; Couty, I.; Batellier, F.; Cotinot, C. Isolation of chicken homolog of the FOXL2 gene and comparison of its expression patterns with those of aromatase during ovarian development. Dev. Dyn. 2004, 231, 859–870. [Google Scholar] [CrossRef]
  165. Luo, W.; Gu, L.; Li, J.; Gong, Y. Transcriptome sequencing revealed that knocking down FOXL2 affected cell proliferation, the cell cycle, and DNA replication in chicken pre-ovulatory follicle cells. PLoS ONE 2020, 15, e0234795. [Google Scholar] [CrossRef] [PubMed]
  166. Tworoger, S.S.; Chubak, J.; Aiello, E.J.; Ulrich, C.M.; Atkinson, C.; Potter, J.D.; Yasui, Y.; Stapleton, P.L.; Lampe, J.W.; Farin, F.M.; et al. Association of CYP17, CYP19, CYP1B1, and COMT polymorphisms with serum and urinary sex hormone concentrations in postmenopausal women. Cancer Epidemiol. Biomark. Prev. 2004, 13, 94–101. [Google Scholar] [CrossRef] [PubMed]
  167. Smolarz, B.; Szyllo, K.; Romanowicz, H. The Genetic Background of Endometriosis: Can ESR2 and CYP19A1 Genes Be a Potential Risk Factor for Its Development? Int. J. Mol. Sci. 2020, 21, 8235. [Google Scholar] [CrossRef] [PubMed]
  168. Smith, C.A.; McClive, P.J.; Hudson, Q.; Sinclair, A.H. Male-specific cell migration into the developing gonad is a conserved process involving PDGF signalling. Dev. Biol. 2005, 284, 337–350. [Google Scholar] [CrossRef] [PubMed]
  169. Smith, C.A.; Andrews, J.E.; Sinclair, A.H. Gonadal sex differentiation in chicken embryos: Expression of estrogen receptor and aromatase genes. J. Steroid Biochem. Mol. Biol. 1997, 60, 295–302. [Google Scholar] [CrossRef]
  170. Nishikimi, H.; Kansaku, N.; Saito, N.; Usami, M.; Ohno, Y.; Shimada, K. Sex differentiation and mRNA expression of p450c17, p450arom and AMH in gonads of the chicken. Mol. Reprod. Dev. 2000, 55, 20–30. [Google Scholar] [CrossRef]
  171. Nakabayashi, O.; Kikuchi, H.; Kikuchi, T.; Mizuno, S. Differential expression of genes for aromatase and estrogen receptor during the gonadal development in chicken embryos. J. Mol. Endocrinol. 1998, 20, 193–202. [Google Scholar] [CrossRef] [PubMed]
  172. Jin, K.; Zuo, Q.; Song, J.; Zhang, Y.; Chen, G.; Li, B. CYP19A1 (aromatase) dominates female gonadal differentiation in chicken (Gallus gallus) embryos sexual differentiation. Biosci. Rep. 2020, 40, BSR20201576. [Google Scholar] [CrossRef] [PubMed]
  173. Bentsi-Barnes, I.K.; Kuo, F.T.; Barlow, G.M.; Pisarska, M.D. Human forkhead L2 represses key genes in granulosa cell differentiation including aromatase, P450scc, and cyclin D2. Fertil. Steril. 2010, 94, 353–356. [Google Scholar] [CrossRef] [PubMed]
  174. Pannetier, M.; Fabre, S.; Batista, F.; Kocer, A.; Renault, L.; Jolivet, G.; Mandon-Pepin, B.; Cotinot, C.; Veitia, R.; Pailhoux, E. FOXL2 activates P450 aromatase gene transcription: Towards a better characterization of the early steps of mammalian ovarian development. J. Mol. Endocrinol. 2006, 36, 399–413. [Google Scholar] [CrossRef]
  175. Guo, Y.; Cheng, L.; Li, X.; Tang, S.; Zhang, X.; Gong, Y. Transcriptional regulation of CYP19A1 expression in chickens: ESR1, ESR2 and NR5A2 form a functional network. Gen. Comp. Endocrinol. 2022, 315, 113939. [Google Scholar] [CrossRef]
  176. Hudson, Q.J.; Smith, C.A.; Sinclair, A.H. Aromatase inhibition reduces expression of FOXL2 in the embryonic chicken ovary. Dev. Dyn. 2005, 233, 1052–1055. [Google Scholar] [CrossRef]
  177. Ellis, H.L.; Shioda, K.; Rosenthal, N.F.; Coser, K.R.; Shioda, T. Masculine epigenetic sex marks of the CYP19A1/aromatase promoter in genetically male chicken embryonic gonads are resistant to estrogen-induced phenotypic sex conversion. Biol. Reprod. 2012, 87, 1–12. [Google Scholar] [CrossRef]
  178. Shioda, K.; Odajima, J.; Kobayashi, M.; Kobayashi, M.; Cordazzo, B.; Isselbacher, K.J.; Shioda, T. Transcriptomic and Epigenetic Preservation of Genetic Sex Identity in Estrogen-feminized Male Chicken Embryonic Gonads. Endocrinology 2021, 162, bqaa208. [Google Scholar] [CrossRef]
  179. Barske, L.A.; Capel, B. Estrogen represses SOX9 during sex determination in the red-eared slider turtle Trachemys scripta. Dev. Biol. 2010, 341, 305–314. [Google Scholar] [CrossRef]
  180. Dougherty, D.C.; Sanders, M.M. Estrogen action: Revitalization of the chick oviduct model. Trends Endocrinol. Metab. 2005, 16, 414–419. [Google Scholar] [CrossRef]
  181. Oka, T.; Schimke, R.T. Interaction of estrogen and progesterone in chick oviduct development. I. Antagonistic effect of progesterone on estrogen-induced proliferation and differentiation of tubular gland cells. J. Cell. Biol. 1969, 41, 816–831. [Google Scholar] [CrossRef] [PubMed]
  182. Palmiter, R.D.; Wrenn, J.T. Interaction of estrogen and progesterone in chick oviduct development. 3. Tubular gland cell cytodifferentiation. J. Cell. Biol. 1971, 50, 598–615. [Google Scholar] [CrossRef] [PubMed]
  183. Oka, T.; Schimke, R.T. Interaction of estrogen and progesterone in chick oviduct development. II. Effects of estrogen and progesterone on tubular gland cell function. J. Cell. Biol. 1969, 43, 123–137. [Google Scholar] [CrossRef] [PubMed]
  184. Wartenberg, H.; Lenz, E.; Schweikert, H.U. Sexual differentiation and the germ cell in sex reversed gonads after aromatase inhibition in the chicken embryo. Andrologia 1992, 24, 1–6. [Google Scholar] [CrossRef] [PubMed]
  185. Guioli, S.; Lovell-Badge, R. PITX2 controls asymmetric gonadal development in both sexes of the chick and can rescue the degeneration of the right ovary. Development 2007, 134, 4199–4208. [Google Scholar] [CrossRef] [PubMed]
  186. Mattsson, A.; Olsson, J.A.; Brunstrom, B. Activation of estrogen receptor alpha disrupts differentiation of the reproductive organs in chicken embryos. Gen. Comp. Endocrinol. 2011, 172, 251–259. [Google Scholar] [CrossRef]
  187. Biason-Lauber, A. WNT4, RSPO1, and FOXL2 in sex development. Semin. Reprod. Med. 2012, 30, 387–395. [Google Scholar] [CrossRef]
  188. Pellegrino, M.; Maiorino, R.; Schonauer, S. WNT4 signaling in female gonadal development. Endocr. Metab. Immune Disord. Drug. Targets 2010, 10, 168–174. [Google Scholar] [CrossRef]
  189. Biason-Lauber, A.; Konrad, D. WNT4 and sex development. Sex. Dev. 2008, 2, 210–218. [Google Scholar] [CrossRef]
  190. Liu, C.-F.; Liu, C.; Yao, H.H.C. Building Pathways for Ovary Organogenesis in the Mouse Embryo. In Organogenesis in Development; Academic Press: Cambridge, MA, USA, 2010; pp. 263–290. [Google Scholar]
  191. Wei, Q.; Yokota, C.; Semenov, M.V.; Doble, B.; Woodgett, J.; He, X. R-spondin1 is a high affinity ligand for LRP6 and induces LRP6 phosphorylation and beta-catenin signaling. J. Biol. Chem. 2007, 282, 15903–15911. [Google Scholar] [CrossRef]
  192. Smith, C.A.; Shoemaker, C.M.; Roeszler, K.N.; Queen, J.; Crews, D.; Sinclair, A.H. Cloning and expression of R-Spondin1 in different vertebrates suggests a conserved role in ovarian development. BMC Dev. Biol. 2008, 8, 72. [Google Scholar] [CrossRef] [PubMed]
  193. Ayers, K.L.; Sinclair, A.H.; Smith, C.A. The molecular genetics of ovarian differentiation in the avian model. Sex. Dev. 2013, 7, 80–94. [Google Scholar] [CrossRef] [PubMed]
  194. Mandel, H.; Shemer, R.; Borochowitz, Z.U.; Okopnik, M.; Knopf, C.; Indelman, M.; Drugan, A.; Tiosano, D.; Gershoni-Baruch, R.; Choder, M.; et al. SERKAL syndrome: An autosomal-recessive disorder caused by a loss-of-function mutation in WNT4. Am. J. Hum. Genet. 2008, 82, 39–47. [Google Scholar] [CrossRef] [PubMed]
  195. Vainio, S.; Heikkila, M.; Kispert, A.; Chin, N.; McMahon, A.P. Female development in mammals is regulated by Wnt-4 signalling. Nature 1999, 397, 405–409. [Google Scholar] [CrossRef]
  196. Mayere, C.; Regard, V.; Perea-Gomez, A.; Bunce, C.; Neirijnck, Y.; Djari, C.; Bellido-Carreras, N.; Sararols, P.; Reeves, R.; Greenaway, S.; et al. Origin, specification and differentiation of a rare supporting-like lineage in the developing mouse gonad. Sci. Adv. 2022, 8, eabm0972. [Google Scholar] [CrossRef]
  197. Chassot, A.A.; Ranc, F.; Gregoire, E.P.; Roepers-Gajadien, H.L.; Taketo, M.M.; Camerino, G.; de Rooij, D.G.; Schedl, A.; Chaboissier, M.C. Activation of beta-catenin signaling by Rspo1 controls differentiation of the mammalian ovary. Hum. Mol. Genet. 2008, 17, 1264–1277. [Google Scholar] [CrossRef]
  198. Lu, J.; Chuong, C.-M.; Widelitz, R.B. Isolation and characterization of chicken β-catenin. Gene 1997, 196, 201–207. [Google Scholar] [CrossRef]
  199. Estermann, M.A.; Hirst, C.E.; Major, A.T.; Smith, C.A. The homeobox gene TGIF1 is required for chicken ovarian cortical development and generation of the juxtacortical medulla. Development 2021, 148, dev199646. [Google Scholar] [CrossRef]
  200. Godwin, J. Social determination of sex in reef fishes. Semin. Cell. Dev. Biol. 2009, 20, 264–270. [Google Scholar] [CrossRef]
  201. Lamm, M.S.; Liu, H.; Gemmell, N.J.; Godwin, J.R. The Need for Speed: Neuroendocrine Regulation of Socially-controlled Sex Change. Integr. Comp. Biol. 2015, 55, 307–322. [Google Scholar] [CrossRef]
  202. Liu, H.; Todd, E.V.; Lokman, P.M.; Lamm, M.S.; Godwin, J.R.; Gemmell, N.J. Sexual plasticity: A fishy tale. Mol. Reprod. Dev. 2017, 84, 171–194. [Google Scholar] [CrossRef] [PubMed]
  203. Warner, R. Mating Behavior and Hermaphroditism in Coral Reef Fishes. Am. Sci. 1970, 72, 128–136. [Google Scholar]
  204. Warner, R.R.; Swearer, S.E. Social Control of Sex Change in the Bluehead Wrasse, Thalassoma bifasciatum (Pisces: Labridae). Biol. Bull. 1991, 181, 199–204. [Google Scholar] [CrossRef] [PubMed]
  205. Matsumine, H.; Herbst, M.A.; Ou, S.H.; Wilson, J.D.; McPhaul, M.J. Aromatase mRNA in the extragonadal tissues of chickens with the henny-feathering trait is derived from a distinctive promoter structure that contains a segment of a retroviral long terminal repeat. Functional organization of the Sebright, Leghorn, and Campine aromatase genes. J. Biol. Chem. 1991, 266, 19900–19907. [Google Scholar] [CrossRef] [PubMed]
  206. Balthazart, J.; Cornil, C.A.; Charlier, T.D.; Taziaux, M.; Ball, G.F. Estradiol, a key endocrine signal in the sexual differentiation and activation of reproductive behavior in quail. J. Exp. Zool. A Ecol. Genet. Physiol. 2009, 311, 323–345. [Google Scholar] [CrossRef]
  207. Pieau, C. Temperature variation and sex determination in reptiles. BioEssays 1996, 18, 19–26. [Google Scholar] [CrossRef]
  208. Jost, A. Hormonal factors in the sex differentiation of the mammalian foetus. Philos. Trans. R. Soc. Lond. B Biol. Sci. 1970, 259, 119–130. [Google Scholar] [CrossRef]
  209. Hu, M.C.; Hsu, N.C.; El Hadj, N.B.; Pai, C.I.; Chu, H.P.; Wang, C.K.; Chung, B.C. Steroid deficiency syndromes in mice with targeted disruption of Cyp11a1. Mol. Endocrinol. 2002, 16, 1943–1950. [Google Scholar] [CrossRef]
  210. Wolf, U. Reorganization of the sex-determining pathway with the evolution of placentation. Hum. Genet. 1999, 105, 288–292. [Google Scholar] [CrossRef]
  211. Maraud, R.; Rashedi, M.; Stoll, R. Influence of a temporary embryonic testis graft on the regression of Müllerian ducts in female chick embryo. Development 1982, 67, 81–87. [Google Scholar] [CrossRef]
  212. Maraud, R.; Vergnaud, O.; Rashedi, M. New insights on the mechanism of testis differentiation from the morphogenesis of experimentally induced testes in genetically female chick embryos. Am. J. Anat. 1990, 188, 429–437. [Google Scholar] [CrossRef]
  213. Halldin, K.; Berg, C.; Brandt, I.; Brunstrom, B. Sexual behavior in Japanese quail as a test end point for endocrine disruption: Effects of in ovo exposure to ethinylestradiol and diethylstilbestrol. Env. Health Perspect. 1999, 107, 861–866. [Google Scholar] [CrossRef] [PubMed]
  214. Widelitz, R.B.; Lin, G.W.; Lai, Y.C.; Mayer, J.A.; Tang, P.C.; Cheng, H.C.; Jiang, T.X.; Chen, C.F.; Chuong, C.M. Morpho-regulation in diverse chicken feather formation: Integrating branching modules and sex hormone-dependent morpho-regulatory modules. Dev. Growth Differ. 2019, 61, 124–138. [Google Scholar] [CrossRef]
  215. Ellem, S.J.; Risbridger, G.P. Aromatase and regulating the estrogen:androgen ratio in the prostate gland. J. Steroid Biochem. Mol. Biol. 2010, 118, 246–251. [Google Scholar] [CrossRef] [PubMed]
  216. Nelson, L.R.; Bulun, S.E. Estrogen production and action. J. Am. Acad. Derm. 2001, 45, S116–S124. [Google Scholar] [CrossRef] [PubMed]
  217. Bulun, S.E.; Lin, Z.; Imir, G.; Amin, S.; Demura, M.; Yilmaz, B.; Martin, R.; Utsunomiya, H.; Thung, S.; Gurates, B.; et al. Regulation of aromatase expression in estrogen-responsive breast and uterine disease: From bench to treatment. Pharm. Rev. 2005, 57, 359–383. [Google Scholar] [CrossRef]
  218. Hirst, C.E.; Major, A.T.; Ayers, K.L.; Brown, R.J.; Mariette, M.; Sackton, T.B.; Smith, C.A. Sex Reversal and Comparative Data Undermine the W Chromosome and Support Z-linked DMRT1 as the Regulator of Gonadal Sex Differentiation in Birds. Endocrinology 2017, 158, 2970–2987. [Google Scholar] [CrossRef]
  219. Testaz, S.; Duband, J.L. Central role of the alpha4beta1 integrin in the coordination of avian truncal neural crest cell adhesion, migration, and survival. Dev. Dyn. 2001, 222, 127–140. [Google Scholar] [CrossRef]
  220. Cutting, A.D.; Ayers, K.; Davidson, N.; Oshlack, A.; Doran, T.; Sinclair, A.H.; Tizard, M.; Smith, C.A. Identification, expression, and regulation of anti-Mullerian hormone type-II receptor in the embryonic chicken gonad. Biol. Reprod. 2014, 90, 106. [Google Scholar] [CrossRef]
  221. Burke, W.H.; Henry, M.H. Gonadal development and growth of chickens and turkeys hatched from eggs injected with an aromatase inhibitor. Poult. Sci. 1999, 78, 1019–1033. [Google Scholar] [CrossRef]
  222. Estermann, M.A.; Smith, C.A. Fadrozole-mediated sex reversal in the embryonic chicken gonad involves a PAX2 positive undifferentiated supporting cell state. bioRxiv 2022. [Google Scholar] [CrossRef] [PubMed]
  223. Yang, X.; Zheng, J.; Na, R.; Li, J.; Xu, G.; Qu, L.; Yang, N. Degree of sex differentiation of genetic female chicken treated with different doses of an aromatase inhibitor. Sex. Dev. 2008, 2, 309–315. [Google Scholar] [CrossRef]
  224. Sechman, A.; Rzasa, J.; Paczoska-Eliasiewicz, H. Effect of non-steroidal aromatase inhibitor on blood plasma ovarian steroid and thyroid hormones in laying hen (Gallus domesticus). J. Vet. Med. A Physiol. Pathol. Clin. Med. 2003, 50, 333–338. [Google Scholar] [CrossRef]
  225. Mohammadrezaei, M.; Toghyani, M.; Gheisari, A.; Toghyani, M.; Eghbalsaied, S. Synergistic effect of fadrozole and insulin-like growth factor-I on female-to-male sex reversal and body weight of broiler chicks. PLoS ONE 2014, 9, e103570. [Google Scholar] [CrossRef]
  226. Abinawanto; Zhang, C.; Saito, N.; Matsuda, Y.; Shimada, K. Identification of sperm-bearing female-specific chromosome in the sex-reversed chicken. J. Exp. Zool. 1998, 280, 65–72. [Google Scholar] [CrossRef]
  227. Vaillant, S.; Guemene, D.; Dorizzi, M.; Pieau, C.; Richard-Mercier, N.; Brillard, J.P. Degree of sex reversal as related to plasma steroid levels in genetic female chickens (Gallus domesticus) treated with Fadrozole. Mol. Reprod. Dev. 2003, 65, 420–428. [Google Scholar] [CrossRef]
  228. Takagi, S.; Tsukada, A.; Saito, N.; Shimada, K. Fertilizing ability of chicken sperm bearing the W chromosome. Poult. Sci. 2007, 86, 731–738. [Google Scholar] [CrossRef] [PubMed]
  229. Brunstrom, B.; Axelsson, J.; Mattsson, A.; Halldin, K. Effects of estrogens on sex differentiation in Japanese quail and chicken. Gen. Comp. Endocrinol. 2009, 163, 97–103. [Google Scholar] [CrossRef]
  230. Scholz, B.; Kultima, K.; Mattsson, A.; Axelsson, J.; Brunstrom, B.; Halldin, K.; Stigson, M.; Dencker, L. Sex-dependent gene expression in early brain development of chicken embryos. BMC Neurosci. 2006, 7, 12. [Google Scholar] [CrossRef] [PubMed]
  231. Ayers, K.L.; Davidson, N.M.; Demiyah, D.; Roeszler, K.N.; Grutzner, F.; Sinclair, A.H.; Oshlack, A.; Smith, C.A. RNA sequencing reveals sexually dimorphic gene expression before gonadal differentiation in chicken and allows comprehensive annotation of the W-chromosome. Genome Biol. 2013, 14, R26. [Google Scholar] [CrossRef]
  232. Morris, K.R.; Hirst, C.E.; Major, A.T.; Ezaz, T.; Ford, M.; Bibby, S.; Doran, T.J.; Smith, C.A. Gonadal and Endocrine Analysis of a Gynandromorphic Chicken. Endocrinology 2018, 159, 3492–3502. [Google Scholar] [CrossRef]
  233. Briganti, F.; Papeschi, A.; Mugnai, T.; Dessì-Fulgheri, F. Effect of testosterone on male traits and behaviour in juvenile pheasants. Ethol. Ecol. Evol. 1999, 11, 171–178. [Google Scholar] [CrossRef]
  234. Fennell, M.J.; Scanes, C.G. Inhibition of growth in chickens by testosterone, 5 alpha-dihydrotestosterone, and 19-nortestosterone. Poult. Sci. 1992, 71, 357–366. [Google Scholar] [CrossRef] [PubMed]
  235. Park, J.S.; Lee, K.Y.; Han, J.Y. Precise Genome Editing in Poultry and Its Application to Industries. Genes 2020, 11, 1182. [Google Scholar] [CrossRef] [PubMed]
  236. Gautron, J.; Rehault-Godbert, S.; Van de Braak, T.G.H.; Dunn, I.C. Review: What are the challenges facing the table egg industry in the next decades and what can be done to address them? Animal 2021, 15 (Suppl. 1), 100282. [Google Scholar] [CrossRef] [PubMed]
  237. Salgado Pardo, J.I.; Navas Gonzalez, F.J.; Gonzalez Ariza, A.; Arando Arbulu, A.; Leon Jurado, J.M.; Delgado Bermejo, J.V.; Camacho Vallejo, M.E. Traditional sexing methods and external egg characteristics combination allow highly accurate early sex determination in an endangered native turkey breed. Front. Vet. Sci. 2022, 9, 948502. [Google Scholar] [CrossRef]
  238. Raj, M. Welfare during stunning and slaughter of poultry. Poult. Sci. 1998, 77, 1815–1819. [Google Scholar] [CrossRef]
  239. Dwinger, R.; Lambooij, B. A brief summary of European legislation regarding animal welfare. Berl. Und Munch. Tierarztl. Wochenschr. 2012, 125, 297–304. [Google Scholar]
  240. Xi, J.F.; Wang, X.Z.; Zhang, Y.S.; Jia, B.; Li, C.C.; Wang, X.H.; Ying, R.W. Sex control by Zfy siRNA in the dairy cattle. Anim. Reprod. Sci. 2019, 200, 1–6. [Google Scholar] [CrossRef]
  241. Espinosa-Cervantes, R.; Cordova-Izquierdo, A. Sexing sperm of domestic animals. Trop. Anim. Health Prod. 2013, 45, 1–8. [Google Scholar] [CrossRef]
  242. Xiang, X.; Yu, Z.; Liu, Y.; Huang, Y.; Wang, J.; Chen, L.; Ma, M. Differential proteomics between unhatched male and female egg yolks reveal the molecular mechanisms of sex-allocation and sex-determination in chicken. Poult. Sci. 2022, 101, 101906. [Google Scholar] [CrossRef] [PubMed]
  243. Duncan, G.E. Determining the health benefits of poultry industry compliance measures: The case of campylobacteriosis regulation in New Zealand. New. Zealand Med. J. 2014, 127, 22–37. [Google Scholar] [PubMed]
  244. Gibson, T.J.; Jackson, E.L. The economics of animal welfare. Rev. Sci. Tech. (Int. Off. Epizoot.) 2017, 36, 125–135. [Google Scholar] [CrossRef] [PubMed]
  245. Montalcini, C.M.; Voelkl, B.; Gomez, Y.; Gantner, M.; Toscano, M.J. Evaluation of an Active LF Tracking System and Data Processing Methods for Livestock Precision Farming in the Poultry Sector. Sensors 2022, 22, 659. [Google Scholar] [CrossRef]
  246. Muir, W.M.; Cheng, H.W.; Croney, C. Methods to address poultry robustness and welfare issues through breeding and associated ethical considerations. Front. Genet. 2014, 5, 407. [Google Scholar] [CrossRef]
  247. Siqueira, T.S.; Borges, T.D.; Rocha, R.M.M.; Figueira, P.T.; Luciano, F.B.; Macedo, R.E.F. Effect of electrical stunning frequency and current waveform in poultry welfare and meat quality. Poult. Sci. 2017, 96, 2956–2964. [Google Scholar] [CrossRef]
  248. Bonafos, L.; Simonin, D.; Gavinelli, A. Animal welfare: European legislation and future perspectives. J. Vet. Med. Educ. 2010, 37, 26–29. [Google Scholar] [CrossRef]
  249. Ni, J.Q.; Erasmus, M.A.; Croney, C.C.; Li, C.; Li, Y. A critical review of advancement in scientific research on food animal welfare-related air pollution. J. Hazard. Mater. 2021, 408, 124468. [Google Scholar] [CrossRef]
  250. Eide, A.L.; Glover, J.C. Development of the longitudinal projection patterns of lumbar primary sensory afferents in the chicken embryo. J. Comp. Neurol. 1995, 353, 247–259. [Google Scholar] [CrossRef]
  251. Eide, A.L.; Glover, J.C. Developmental dynamics of functionally specific primary sensory afferent projections in the chicken embryo. Anat. Embryol. 1997, 195, 237–250. [Google Scholar] [CrossRef]
  252. Gohler, D.; Fischer, B.; Meissner, S. In-ovo sexing of 14-day-old chicken embryos by pattern analysis in hyperspectral images (VIS/NIR spectra): A non-destructive method for layer lines with gender-specific down feather color. Poult. Sci. 2017, 96, 1–4. [Google Scholar] [CrossRef] [PubMed]
  253. Steiner, G.; Bartels, T.; Stelling, A.; Krautwald-Junghanns, M.E.; Fuhrmann, H.; Sablinskas, V.; Koch, E. Gender determination of fertilized unincubated chicken eggs by infrared spectroscopic imaging. Anal. Bioanal. Chem. 2011, 400, 2775–2782. [Google Scholar] [CrossRef] [PubMed]
  254. Whittaker, D.J.; Soini, H.A.; Gerlach, N.M.; Posto, A.L.; Novotny, M.V.; Ketterson, E.D. Role of testosterone in stimulating seasonal changes in a potential avian chemosignal. J. Chem. Ecol. 2011, 37, 1349–1357. [Google Scholar] [CrossRef] [PubMed]
  255. Dodo, K.; Fujita, K.; Sodeoka, M. Raman Spectroscopy for Chemical Biology Research. J. Am. Chem. Soc. 2022, 144, 19651–19667. [Google Scholar] [CrossRef] [PubMed]
  256. Auner, G.W.; Koya, S.K.; Huang, C.; Broadbent, B.; Trexler, M.; Auner, Z.; Elias, A.; Mehne, K.C.; Brusatori, M.A. Applications of Raman spectroscopy in cancer diagnosis. Cancer Metastasis Rev. 2018, 37, 691–717. [Google Scholar] [CrossRef]
  257. Faur, C.I.; Falamas, A.; Chirila, M.; Roman, R.C.; Rotaru, H.; Moldovan, M.A.; Albu, S.; Baciut, M.; Robu, I.; Hedesiu, M. Raman spectroscopy in oral cavity and oropharyngeal cancer: A systematic review. Int. J. Oral. Maxillofac. Surg. 2022, 51, 1373–1381. [Google Scholar] [CrossRef]
  258. Harz, M.; Krause, M.; Bartels, T.; Cramer, K.; Rosch, P.; Popp, J. Minimal invasive gender determination of birds by means of UV-resonance Raman spectroscopy. Anal. Chem. 2008, 80, 1080–1086. [Google Scholar] [CrossRef]
  259. Galli, R.; Preusse, G.; Uckermann, O.; Bartels, T.; Krautwald-Junghanns, M.E.; Koch, E.; Steiner, G. In Ovo Sexing of Domestic Chicken Eggs by Raman Spectroscopy. Anal. Chem. 2016, 88, 8657–8663. [Google Scholar] [CrossRef]
  260. Fang, Q.; Papaioannou, T.; Jo, J.A.; Vaitha, R.; Shastry, K.; Marcu, L. Time-domain laser-induced fluorescence spectroscopy apparatus for clinical diagnostics. Rev. Sci. Instrum. 2004, 75, 151–162. [Google Scholar] [CrossRef]
  261. Perez Rubio, A.; Eiros, J.M. Cell culture-derived flu vaccine: Present and future. Hum. Vaccines Immunother. 2018, 14, 1874–1882. [Google Scholar] [CrossRef]
  262. Gouma, S.; Anderson, E.M.; Hensley, S.E. Challenges of Making Effective Influenza Vaccines. Annu. Rev. Virol. 2020, 7, 495–512. [Google Scholar] [CrossRef] [PubMed]
  263. Quansah, E.S.; Urwin, N.A.R.; Strappe, P.; Raidal, S. Progress towards generation of transgenic lines of chicken with a green fluorescent protein gene in the female specific (w) chromosome by sperm-mediated gene transfer. Adv. Genet. Eng. 2013, 2, 29. [Google Scholar]
  264. Bruijnis, M.R.N.; Blok, V.; Stassen, E.N.; Gremmen, H.G.J. Moral “Lock-In” in Responsible Innovation: The Ethical and Social Aspects of Killing Day-Old Chicks and Its Alternatives. J. Agr. Env. Ethics 2015, 28, 939–960. [Google Scholar] [CrossRef]
  265. DuRant, S.E.; Hopkins, W.A.; Carter, A.W.; Kirkpatrick, L.T.; Navara, K.J.; Hawley, D.M. Incubation temperature causes skewed sex ratios in a precocial bird. J. Exp. Biol. 2016, 219, 1961–1964. [Google Scholar] [CrossRef] [PubMed]
  266. Goerlich-Jansson, V.C.; Muller, M.S.; Groothuis, T.G. Manipulation of primary sex ratio in birds: Lessons from the homing pigeon (Columba livia domestica). Integr. Comp. Biol. 2013, 53, 902–912. [Google Scholar] [CrossRef]
  267. Kuroki, S.; Tachibana, M. Epigenetic regulation of mammalian sex determination. Mol. Cell. Endocrinol. 2018, 468, 31–38. [Google Scholar] [CrossRef] [PubMed]
  268. Ridnik, M.; Schoenfelder, S.; Gonen, N. Cis-Regulatory Control of Mammalian Sex Determination. Sex. Dev. 2021, 15, 317–334. [Google Scholar] [CrossRef] [PubMed]
  269. Rea, S.; Eisenhaber, F.; O’Carroll, D.; Strahl, B.D.; Sun, Z.W.; Schmid, M.; Opravil, S.; Mechtler, K.; Ponting, C.P.; Allis, C.D.; et al. Regulation of chromatin structure by site-specific histone H3 methyltransferases. Nature 2000, 406, 593–599. [Google Scholar] [CrossRef]
  270. Martin, C.; Zhang, Y. The diverse functions of histone lysine methylation. Nat. Rev. Mol. Cell Biol. 2005, 6, 838–849. [Google Scholar] [CrossRef]
  271. Nishino, K.; Hattori, N.; Tanaka, S.; Shiota, K. DNA methylation-mediated control of Sry gene expression in mouse gonadal development. J. Biol. Chem. 2004, 279, 22306–22313. [Google Scholar] [CrossRef]
  272. Dekker, J.; Rippe, K.; Dekker, M.; Kleckner, N. Capturing chromosome conformation. Science 2002, 295, 1306–1311. [Google Scholar] [CrossRef] [PubMed]
  273. Lieberman-Aiden, E.; van Berkum, N.L.; Williams, L.; Imakaev, M.; Ragoczy, T.; Telling, A.; Amit, I.; Lajoie, B.R.; Sabo, P.J.; Dorschner, M.O.; et al. Comprehensive mapping of long-range interactions reveals folding principles of the human genome. Science 2009, 326, 289–293. [Google Scholar] [CrossRef] [PubMed]
  274. Bain, M.M.; Fagan, A.J.; Mullin, J.M.; McNaught, I.; McLean, J.; Condon, B. Noninvasive monitoring of chick development in ovo using a 7T MRI system from day 12 of incubation through to hatching. J. Magn. Reson. Imaging 2007, 26, 198–201. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Overview of the derivation of chicken supporting cells and key genes promoting bipotential supporting cells to adopt either testicular or ovarian development during embryonic stages. The mesonephros gives rise to supporting cells which further form Sertoli and Leydig cells in males or granulose and theca cells in females. In embryonic ZZ gonads, two copies of DMRT1 trigger the differentiation program from supporting cells to pre-Sertoli cells, and several downstream male-determining genes further induce pre-Sertoli cells to form Sertoli and Leydig cells. In the embryonic ZW gonads, one copy of DMRT1 cannot activate the masculine development, and thus female supporting cells differentiate into pre-granulosa cells, which are further induced to form granulosa and theca cells under feminized signals. Modified from Estermann’s articles with permission [101,102].
Figure 1. Overview of the derivation of chicken supporting cells and key genes promoting bipotential supporting cells to adopt either testicular or ovarian development during embryonic stages. The mesonephros gives rise to supporting cells which further form Sertoli and Leydig cells in males or granulose and theca cells in females. In embryonic ZZ gonads, two copies of DMRT1 trigger the differentiation program from supporting cells to pre-Sertoli cells, and several downstream male-determining genes further induce pre-Sertoli cells to form Sertoli and Leydig cells. In the embryonic ZW gonads, one copy of DMRT1 cannot activate the masculine development, and thus female supporting cells differentiate into pre-granulosa cells, which are further induced to form granulosa and theca cells under feminized signals. Modified from Estermann’s articles with permission [101,102].
Ijms 24 08284 g001
Figure 2. Schematic view of key factors influencing the embryonic sex determination and differentiation in maternal and extra-maternal developmental patterns. In mouse, the process of embryonic sex development occurs in the maternal environment, and is mainly regulated by genetic and epigenetic factors. In chicken, the process of embryonic sex development occurs in the extra-maternal environment, and is governed by genetic and epigenetic factors and sex hormones. Likewise, the process of turtle embryonic sex development occurs in the extra-maternal environment, and is controlled by genetic and epigenetic factors, sex hormones, and temperature.
Figure 2. Schematic view of key factors influencing the embryonic sex determination and differentiation in maternal and extra-maternal developmental patterns. In mouse, the process of embryonic sex development occurs in the maternal environment, and is mainly regulated by genetic and epigenetic factors. In chicken, the process of embryonic sex development occurs in the extra-maternal environment, and is governed by genetic and epigenetic factors and sex hormones. Likewise, the process of turtle embryonic sex development occurs in the extra-maternal environment, and is controlled by genetic and epigenetic factors, sex hormones, and temperature.
Ijms 24 08284 g002
Figure 3. Overview of gonadal phenotypes and secondary sexual characteristics in embryonic and adult sex-reversed chickens created by treatment with either estrogens or aromatase inhibitors. The left panel: Injection of estrogens into chicken eggs can lead to male-to-female sex reversal, characterized by the proliferation of the left gonad and degeneration of the right gonad. During adult period, injection of estrogens into leg muscles of male chicken can induce the feminization of feathering patterns. The right panel: Injection of aromatase inhibitors into chicken eggs can lead to female-to-male sex reversal, characterized by the symmetrical proliferation and differentiation of bilateral gonads. During the adult period, the secondary sexual characteristics and gonadal appearance show different degrees of sex reversal in aromatase inhibitors-treated female chicken.
Figure 3. Overview of gonadal phenotypes and secondary sexual characteristics in embryonic and adult sex-reversed chickens created by treatment with either estrogens or aromatase inhibitors. The left panel: Injection of estrogens into chicken eggs can lead to male-to-female sex reversal, characterized by the proliferation of the left gonad and degeneration of the right gonad. During adult period, injection of estrogens into leg muscles of male chicken can induce the feminization of feathering patterns. The right panel: Injection of aromatase inhibitors into chicken eggs can lead to female-to-male sex reversal, characterized by the symmetrical proliferation and differentiation of bilateral gonads. During the adult period, the secondary sexual characteristics and gonadal appearance show different degrees of sex reversal in aromatase inhibitors-treated female chicken.
Ijms 24 08284 g003
Figure 4. Schematic view of early avian sex control technologies. Before incubation, previous studies attempted to clarify the association between the egg shape and the sex of hatchlings, as well as the correlation between the egg odor and the embryonic gender. During incubation, research focused on identifying the sex-specific biochemical composition by spectroscopic detection and investigating fluorescent biomarkers created by gene-editing systems to detect the gender of embryos.
Figure 4. Schematic view of early avian sex control technologies. Before incubation, previous studies attempted to clarify the association between the egg shape and the sex of hatchlings, as well as the correlation between the egg odor and the embryonic gender. During incubation, research focused on identifying the sex-specific biochemical composition by spectroscopic detection and investigating fluorescent biomarkers created by gene-editing systems to detect the gender of embryos.
Ijms 24 08284 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, X.; Li, J.; Chen, S.; Yang, N.; Zheng, J. Overview of Avian Sex Reversal. Int. J. Mol. Sci. 2023, 24, 8284. https://doi.org/10.3390/ijms24098284

AMA Style

Zhang X, Li J, Chen S, Yang N, Zheng J. Overview of Avian Sex Reversal. International Journal of Molecular Sciences. 2023; 24(9):8284. https://doi.org/10.3390/ijms24098284

Chicago/Turabian Style

Zhang, Xiuan, Jianbo Li, Sirui Chen, Ning Yang, and Jiangxia Zheng. 2023. "Overview of Avian Sex Reversal" International Journal of Molecular Sciences 24, no. 9: 8284. https://doi.org/10.3390/ijms24098284

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop