Abstract
Free full text
Compensation of gene dosage on the mammalian X
ABSTRACT
Changes in gene dosage can have tremendous evolutionary potential (e.g. whole-genome duplications), but without compensatory mechanisms, they can also lead to gene dysregulation and pathologies. Sex chromosomes are a paradigmatic example of naturally occurring gene dosage differences and their compensation. In species with chromosome-based sex determination, individuals within the same population necessarily show ‘natural’ differences in gene dosage for the sex chromosomes. In this Review, we focus on the mammalian X chromosome and discuss recent new insights into the dosage-compensation mechanisms that evolved along with the emergence of sex chromosomes, namely X-inactivation and X-upregulation. We also discuss the evolution of the genetic loci and molecular players involved, as well as the regulatory diversity and potentially different requirements for dosage compensation across mammalian species.
Introduction
Gene dosage denotes the number of gene copies present in a cell of an organism, which can be reflected in the amount of gene products, such as proteins and functional RNAs (Basilicata and Keller Valsecchi, 2021). Changes in gene dosage can, therefore, produce significant phenotypic consequences. For an individual, they often lead to harmful consequences (e.g. gene amplification of HER2 receptor associated with breast cancers; Seshadri et al., 1989) but, at an evolutionary scale, they contribute to adaptation and speciation (e.g. gene duplications; Kondrashov, 2012; Qian and Zhang, 2014). Gene-dosage changes can arise due to copy number variations (e.g. gene amplification, insertions or deletions) or changes in ploidy (i.e. the number of chromosomes sets in a cell via gain or loss of chromosomes or whole-genome duplications). The effects of having an extra chromosome are a lot more detrimental than having a whole extra set of chromosomes, as initially revealed by seminal experiments with the flowering plant Datura stramonium and with Drosophila melanogaster (Blakeslee, 1934; Blakeslee and Belling, 1924; Blakeslee et al., 1920; Bridges, 1921; reviewed by Birchler and Veitia, 2021). This led to the ‘gene-balance hypothesis’ (Birchler and Veitia, 2007, 2010, 2012), whereby maintaining a balanced gene dosage across the genome is crucial, especially for genes coding for products involved in functions where stoichiometry is important (e.g. members of multi-subunit complexes). Sex chromosomes (also known as ‘allosomes’) challenge such balanced gene dosage (Graves, 2006). In many animal and plant species where sex is determined by chromosomes (Box 1), individuals within the same population naturally exhibit differences in copy number for the genes within the sex chromosomes. In mammals, which have an XY sex-determination system and are the focus of our article, females have two copies of each gene residing on the X chromosome(s), whereas males have only one such copy, plus one copy of Y-linked genes, many of which are ‘unmatched’ in females (Box 2). According to GENCODE, the X chromosome in mouse has at least 932 protein-coding genes and 558 noncoding-RNA genes (Frankish et al., 2023); therefore, such asymmetry in gene dosage could lead to significant phenotypic consequences if left uncompensated. The X harbours genes involved in fundamental cell processes independent of sex-related functions; dozens of housekeeping genes are found on the X of both humans and mice: 99 and 91, respectively, according to a recent database (Hounkpe et al., 2021). This is consistent with mammalian sex chromosomes evolving from a precursor pair of autosomes, as we review below. The asymmetry in X-linked gene dosage in relation to autosomal gene dosage between the two sexes is thus thought to have favoured the emergence of sex-specific ‘dosage-compensation’ strategies.
The emergence of sex chromosomes in mammals and their asymmetries in gene dosage
Susumu Ohno, a pioneer in the study of sex chromosome evolution, proposed that sex chromosomes originated from a precursor pair of autosomes (called ‘proto-sex chromosomes’), which underwent key mutations generating sex-determining loci (Ohno, 1967). Comparative genomics has revealed that the sex chromosomes of living mammals, which include prototherians (monotremes, such as the platypus), metatherians (marsupials, such as the wombat) and eutherians (placental mammals, such as the mouse), have different evolutionary origins (Bellott et al., 2014; Bininda-Emonds et al., 2007; Cortez et al., 2014; Luo et al., 2011; Marshall Graves, 2008; Messer et al., 1998; Potrzebowski et al., 2010). The sex chromosomes of marsupials and placental mammals are homologous, positioning their emergence approximately 166 million years ago (preceding the divergence of the metatherian and eutherian lineages), whereas the sex chromosomes of monotremes emerged through a parallel path (Box 3).
Ohno's hypothesis expanded on the idea of Hermann J. Muller that the differentiation of sex chromosomes would follow the lack of recombination caused by the appearance of a sex-determining gene (Muller, 1914). In therian mammals, the first step in the evolution of the proto-sex chromosomes is considered to be the acquisition of the male-determining gene, sex-determining region Y (SRY), on one of the proto-sex chromosomes (Foster and Graves, 1994; Foster et al., 1992; Gubbay et al., 1990). SRY is thought to have evolved from a mutation in one allele of the proto SRY-related HMG box-containing gene 3 (SOX3) gene (Collignon et al., 1996; Stevanović et al., 1993; Sutton et al., 2011). In present-day therian chromosomes, SRY is located on the Y chromosome and SOX3 on the X chromosome. The emergence of SRY is thought to have been followed by a series of other events, the order of which remains under debate: the emergence/accumulation of other male-specific genes on the proto-Y, the suppression of meiotic recombination between the evolving proto-X and proto-Y chromosomes, and the progressive degradation of the proto-Y in terms of gene content due to lack of recombination (Bachtrog et al., 2011; Bergero and Charlesworth, 2009; Charlesworth et al., 2005; Chibalina and Filatov, 2011; Felsenstein, 1974; Rice, 1996; Wright et al., 2016). This progressive differentiation of the sex chromosomes meant that genes on the proto-sex chromosomes, once present in two copies and ‘in balance’ with genes across other autosomes, progressively became ‘haploid’ and (potentially) ‘unbalanced’. In other words, the heterogametic sex (XY) became a ‘natural aneuploid’ for X-linked genes (Disteche, 2016), with X-linked gene dosage reduced from two to one in XY individuals. This process is thought to have been accompanied by the emergence of dosage-compensation mechanisms to restore the balance between autosomal and allosomal gene expression, which we discuss below.
Dosage compensating the X chromosome: a two-step hypothesis
Ohno's influential hypothesis on the evolution of the sex chromosomes (Ohno, 1967) put forth two steps to account for the dosage compensation of X-linked gene expression: a first step entailing an increase of the activity of the X chromosome, aiming to balance the levels of gene products from the single X in males with those from the two sets of autosomes; and a second step required to counteract the effects of the first one in females, by deactivating one of their two X chromosomes, thereby bringing the levels of X-linked gene products down to the disomic levels from autosomal chromosomes. This second step was drawn from the insightful hypothesis proposed by the geneticist Mary Lyon of X-chromosome inactivation (XCI), a phenomenon that has since been confirmed and is well-established (Blewitt, 2024; Lyon, 1961). The first step, on the other hand, presupposes the existence of X-chromosome upregulation (XCU), which has remained controversial in mammals.
X-chromosome upregulation: hypotheses, observations and mechanisms
Longstanding controversies include whether or not XCU is present in mammals, and if so, whether it is global or affects only a subset of genes, and to which extent (whether it achieves complete dosage compensation or only partial). We have compiled a list of the studies that have investigated XCU in mammals and included their conclusions, approaches and data used for analysis (Table 1). One of the main reasons why different studies reached different conclusions is their approach when determining XCU. Many authors have compared the expression levels of X-linked genes with that of autosomal genes across several tissues in different mammals. Using this approach, the vast majority of studies have reported similar global levels of expression of X-linked and autosomal genes (based on expression ratios and/or distributions), thus concluding that upregulation of the single active X in mammals occurs. The two exceptions are a study using low-coverage proteomics data and a study in which non-expressed genes were not discarded from the analysis (Deng et al., 2011; He et al., 2011; Kharchenko et al., 2011). Concluding that XCU takes place from the fact that expression levels of the single active present-day X chromosome are similar to expression levels of autosomes assumes that, before sex-chromosome differentiation, expression levels of genes on the (ancestral) proto-sex chromosomes were similar to those on the ancestral autosomes. Such an assumption is not directly derived from Ohno's hypothesis, which did not postulate similar levels of expression but balanced levels of expression, which had to be preserved upon sex-chromosome differentiation. Thus, other authors have argued that comparing X-linked expression levels to autosomal expression levels is not a real test of Ohno's hypothesis (He et al., 2011; Julien et al., 2012; Lin et al., 2012). Instead, they should be compared to expression levels in the ancestral proto-sex chromosomes, for which these authors proposed to use, as a proxy, expression levels of the genes in monotremes and birds that are (autosomal) orthologs of the therian X-linked genes. Based on these comparative analyses, no upregulation was observed, leading the authors to refute Ohno's hypothesis. This approach makes assumptions too, and whether contemporary mammals can be directly compared to contemporary birds was initially questioned (Disteche, 2016). Meanwhile, consistent results have been achieved using other outgroups and various sets of autosomal genes (Julien et al., 2012; Marin et al., 2017; Wang et al., 2020), and although no XCU was found in placental mammals, full global XCU was demonstrated in marsupials (Julien et al., 2012). Current-to-ancestral comparisons may not always be feasible (if relevant data, including for outgroups, is not available; e.g. during specific developmental stages), so careful X-to-autosome comparisons can still be relevant and informative.
Table 1.
These considerations mirror the challenge of defining sex-chromosome dosage compensation. Some authors adopt a broader definition, such as ‘the regulatory mechanisms that balance gene expression between the autosomes and sex chromosomes in the heterogametic sex’ (Mank et al., 2011), whereas others explicitly include the notion of the evolutionary history of the chromosomes; for example, ‘the maintenance of ancestral expression levels of sex-linked genes relative to autosomal expression in the heterogametic sex’ (Gu and Walters, 2017).
So, is there XCU or not in placental mammals? XCU as predicted in Ohno's hypothesis, which refers to higher expression of the present-day X compared with that of a single proto-sex chromosome, cannot be directly tested. As we reviewed, different approaches to address this question have different assumptions and have reached different conclusions. The question is not only whether the X is upregulated or not, but to what extent. Based on current-to-ancestral comparisons in placentals, a twofold upregulation is not achieved at the mRNA level but upregulation takes place – at least for some genes, as proposed by many (Naik et al., 2022; reviewed by Gu and Walters, 2017; Mank et al., 2011; Pessia et al., 2014). Recent single-cell, single-allele RNA-sequencing has confirmed higher expression from genes on the active X chromosome (Lentini et al., 2022), as discussed further below. At the molecular level, the (active) X chromosome is enriched in features that are all consistent with a ‘hyperactive’ transcriptional state compared with autosomes (Fig. 1). Its gene promoters show higher transcriptional burst frequencies (Larsson et al., 2019; Talon et al., 2021) and are enriched in the initiation form of RNA polymerase II, active histone marks, including histone acetylation (H4K16ac), and the corresponding acetyltransferase (MOF) that mediates XCU in Drosophila (Deng et al., 2011, 2013; Yildirim et al., 2011). Concomitantly, the active X shows higher chromatin accessibility than autosomes, as profiled by single-cell ATAC-seq (Talon et al., 2021). This investigation has identified increased chromatin accessibility on the active X chromosomes in mouse XX fibroblasts and XY mouse embryonic stem cells (mESCs), but not on the active X chromosomes of XX induced pluripotent stem cells (iPSCs) or mESCs. Interestingly, these results match the observations that the X chromosomes in mouse XX fibroblasts and XY mESCs are upregulated, whereas X chromosomes in XX mESCs are not (Larsson et al., 2019; Lentini et al., 2022). Recently, the BRD4 protein (containing bromodomains, which recognise acetylated lysine residues such as those in histones) has been implicated in the transcriptional activation of X-linked genes showing upregulation (Lyu et al., 2022), but this has been contested (Lentini and Reinius, 2023; Lyu et al., 2023).
A seemingly absent full dosage compensation (at the transcriptional level) has led to alternative hypotheses for the origin of XCI and partial XCU, unrelated to dosage compensation (Chandra, 1985, 2022; Engelstädter and Haig, 2008; Gribnau and Grootegoed, 2012; Haig, 2006; Iwasa and Pomiankowski, 2001; Mank et al., 2011; Pessia et al., 2014). Recently, however, the Kaessmann lab has proposed a reconciliatory perspective. Translatome analysis of tissues from four therian species (but not from platypus, a monotreme) revealed ‘translation upregulation’, with higher ratios of current-to-ancestral expression for the translatome than for the transcriptome (Wang et al., 2020). This was also associated with higher translation efficiencies and protein abundance (Wang et al., 2020). Combined with transcriptional upregulation, translational upregulation appears to have largely restored ancestral expression levels and, thus, X-to-autosome balance (Wang et al., 2020). Accordingly, previous studies have reported that X-linked transcripts have significantly higher ribosome density (Faucillion and Larsson, 2015) and longer half-lives (Deng et al., 2013; Faucillion and Larsson, 2015; Rücklé et al., 2023) than autosomal transcripts. Recently, depletion of RNA-associated N6-methyladenosine (m6A) modification led to a reduction in the ratio of X:autosome expression in both mouse and human cells, mainly through an increase in the stability of autosomal transcripts (Rücklé et al., 2023). X-linked transcripts were mostly unaffected, which is explained by their low(er) levels of m6A (Rücklé et al., 2023). This appears to be an intrinsic feature of X-linked transcripts, which show a depletion of the GGACH sequence, the m6A consensus motif (Rücklé et al., 2023). This suggests that the higher stability of X-linked mRNAs is hard-wired in the X-chromosome DNA sequence; how this has evolved remains an intriguing open question. In summary, full dosage compensation in placental mammals appears to be happening through a combination of transcriptional and post-transcriptional upregulation of gene expression (Wang et al., 2020), and thus Ohno's hypothesis stands after all.
Unlike XCU in Drosophila and XCI in mammals, mammalian XCU occurs in both sexes and does not appear to rely on a chromosome-wide mechanism – two aspects that we believe have contributed to hinder our understanding of this enigmatic process. Despite an increasingly better molecular understanding of XCU, many questions remain unanswered: namely, which mechanisms confer specificity to the upregulation (i.e. how do they target X-linked genes), and whether all X-linked genes or a subset need to be upregulated? Of note, alternative mechanisms have been reported that have (potentially) allowed compensation the hemizygosity of X-linked genes during sex chromosome evolution; these include the downregulation of autosomal genes that are partners of X-linked genes (Julien et al., 2012), retention of a functional gene copy on the Y chromosome (Bellott et al., 2014; Cortez et al., 2014), duplication of genes on the X (Julien et al., 2012), and relocalisation of proto-Y genes to autosomes (Carelli et al., 2016; Hughes et al., 2015; Potrzebowski et al., 2008). Interestingly, in rodents that have lost the Y chromosome completely, some ‘Y-linked’ genes (presumably the dosage-sensitive ones) are found on the X or autosomes (Arakawa et al., 2002; Kuroiwa et al., 2010; Mulugeta et al., 2016).
X-chromosome inactivation: convergent evolution in therian mammals
Based on several genetic studies and observations in mammals, Mary Lyon put forward the idea of X-chromosome inactivation in 1961, by proposing that the dark-staining X chromosome in female somatic cells (Barr and Bertram, 1949; Ohno et al., 1959) was inactivated (Lyon, 1961). The 60th anniversary of her seminal proposal was celebrated recently (Moyano Rodriguez and Borensztein, 2023). Insightfully, Lyon also anticipated that this inactive X could be ‘either paternal or maternal in origin in different cells of the same animal’ (what is referred to as random XCI, true for placental mammals but not for marsupials) and that it occurred early in embryonic development (Lyon, 1961). A truly epigenetic process, XCI is heritable through mitosis and can be reversed (X-chromosome reactivation), which happens in specific developmental stages and pathological contexts (Panda et al., 2020; Spaziano and Cantone, 2021; Talon et al., 2019). How XCI is triggered specifically in XX individuals, how it affects only one X chromosome and the molecular mechanisms that are implicated in the transcriptional silencing of the X, which is accompanied by heterochromatinisation and chromosome refolding, have been recently reviewed elsewhere (Kanata et al., 2024; Keniry and Blewitt, 2023; Loda et al., 2022; Mutzel and Schulz, 2020; Schwämmle and Schulz, 2023). Here, we cover the evolutionary diversity observed in XCI across mammalian species and its implications for our understanding of dosage regulation and compensation.
According to Ohno's hypothesis, XCI was the ‘second step’ needed for X-linked dosage compensation upon differentiation of the mammalian sex chromosomes; XCI evolved in XX individuals to counteract the effects of XCU, which balanced X-linked gene expression to autosomes in XY individuals but created a problem for XX. Often in the field, XCI is mentioned as having evolved to ‘equilibrate X-linked gene expression between the sexes’, an oversight because of course selection does not work on the balance between the sexes, but on the individual (Vicoso and Bachtrog, 2009). Besides being imprecise, such formulation reinforces the idea that compensated X-linked gene expression is expected to be the same between the sexes, although this might not be the case. Instead, it just needs to be compatible with life and reproduction in each sex [‘incomplete but sufficient’ (Gu and Walters, 2017)]. This means, for example, that the X:autosome expression ratio (for each gene) does not have to be exactly the same in XX and XY individuals.
Remarkably, XCI is present in both marsupials and placental mammals (Fig. 2), but it appears to have evolved independently in these two lineages (reviewed by Shevchenko et al., 2013). Marsupial and placental XCI do share certain features: relying on the activity of long noncoding RNAs (lncRNAs), having the same functional outcome (silencing of X-linked genes), the inactive X being targeted by H3K27 trimethylation and to the perinucleolar compartment (Mahadevaiah et al., 2009), but their genetic origins are not homologous. At the forefront of orchestrating XCI in placental mammals stands the lncRNA, Xist, discovered more than 30 years ago (https://thenode.biologists.com/xist-discovery/discussion/). Xist is essential for XCI in mice (Marahrens et al., 1997; Penny et al., 1996), but no Xist gene has ever been found in marsupials (Davidow et al., 2007; Deakin et al., 2009; Hore et al., 2007; Okamoto and Heard, 2009; Shevchenko et al., 2007; Waters et al., 2005). The Xist gene is proposed to have emerged de novo in eutherians, exhibiting remnants traceable to mobile elements spanning diverse classes and to Lnx3, a protein-coding gene present in birds and marsupials but that no longer exists in eutherians (Duret et al., 2006; Elisaphenko et al., 2008). In marsupials, XCI is associated with a different lncRNA, Rsx. Although not being a sequence homologue of Xist, their RNAs share many functional attributes, such as the female-specific expression, the association and ‘coating’ of the inactive X, the activity in cis, and a similar protein interactome (Grant et al., 2012; Mahadevaiah et al., 2020; McIntyre et al., 2024). Formal genetic evidence of Rsx being essential for marsupial XCI is still lacking; however, expression of Rsx in mESCs from an autosomal transgene resulted in gene silencing in cis (Grant et al., 2012). It is noteworthy that marsupial XCI is imprinted (it is always the paternal X that is inactivated in XX somatic cells, whereas in placentals XCI is random) and comparatively more incomplete than the Xist-driven process (reviewed by Deakin et al., 2009).
Marsupial and placental XCI are thus a compelling illustration of convergent evolution, driven by two independently evolved lncRNAs that silence the activity of nearly an entire chromosome (McIntyre et al., 2024). Importantly, as mentioned before, the sex chromosomes in marsupials and placentals are homologous, whereas the dosage compensation mechanisms in XX individuals are not; it is unclear whether these were preceded by an older dosage compensation mechanism in the last common ancestor. Some authors have suggested an initial process potentially involving other noncoding RNAs, possibly acting gene-by-gene (as in monotremes, Box 3), which was replaced by chromosome-wide regulation by Xist and Rsx to facilitate more efficient silencing (Gribnau and Grootegoed, 2012; McIntyre et al., 2024). It is also possible that Rsx represents the ancestral regulator, whether acting gene-by-gene or chromosome-wide. Interestingly, a recent study in chicken has shown that a microRNA contributes to the downregulation of Z-linked genes in ZZ individuals, which are upregulated in ZW individuals (Fallahshahroudi et al., 2024 preprint).
Placental XCI: a diversity of roads leading to random XCI
Investigating XCI in placentals other than the mouse (e.g. human, macaque, rabbit, cow and pig) has revealed evolutionary flexibility in the regulation and dynamics of XCI (Fig. 2). Some of the truths we learnt about XCI with the mouse appear to be more of an exception than the rule in placental mammals, probably representing recently evolved characteristics rather than common ancestral traits – a likely reflection of the extensive evolutionary radiation of rodents (Fabre et al., 2012) and of the mouse genome evolving faster than that of larger mammals due to more generation cycles per unit of time (Svoboda, 2018).
When does XCI take place during development? In all species examined so far, the random XCI pattern observed in somatic cells is first detected in post-implantation embryos. This corresponds, for example, to embryonic day (E)6-E7 in the mouse and E15-E17 in the macaque, which raises interesting questions regarding the time and extent to which gene dosage compensation is needed (Heard and Rougeulle, 2021; Okamoto et al., 2021). However, it is also important to consider the differences between chronological time and developmental time (Dubansky, 2018; Garcia-Ojalvo and Bulut-Karslioglu, 2023). Even if macaque embryos go through more days without dosage compensation, in terms of ‘developmental time’ dosage compensation appears to be required at a similar stage as in the mouse (Fig. 3).
Not all species, however, reach random XCI in the same way. In the mouse, there is an earlier wave of XCI, which is paternally imprinted, similar to marsupials (another likely example of convergent evolution) and occurs during preimplantation stages (reviewed by Furlan and Galupa, 2022). Murine imprinted XCI is maintained in the extra-embryonic tissues and reverted in the inner-cell mass of the blastocyst, which gives rise to the epiblast cells, in which random XCI occurs some days later. In placental mammals, imprinted XCI is restricted to extra-embryonic tissues and appears to also be present in rat and bovine embryos (Dindot et al., 2004; Magaraki et al., 2019; Wake et al., 1976; Xue et al., 2002; Yu et al., 2020), but absent in human, macaque, rabbit, pig and horse, where extra-embryonic tissues show random XCI (Beckelmann et al., 2012; Goszczynski et al., 2021; Moreira de Mello et al., 2010; Okamoto et al., 2011, 2021; Ramos-Ibeas et al., 2019; Romagnano et al., 1987; Wang et al., 2012; Zou et al., 2019). Which species show imprinted XCI does not appear to be associated with the type of placental structure (Laudon et al., 2024). Instead, it has been linked to earlier zygotic genome activation (ZGA) and faster development (Migeon, 2002), but this does not appear to hold in bovine embryos (Svoboda, 2018). Evolutionary considerations about imprinted XCI have recently been reviewed by Furlan and Galupa (2022).
In the mouse, the expression of Xist is tightly coupled to XCI. Intriguingly, this is not the case for other mammalian species. During preimplantation development, XIST RNA is detected coating the X chromosome for several days without inducing gene silencing in human, monkey, rabbit and bovine embryos (Okamoto et al., 2011, 2021; Yu et al., 2020). Moreover, XIST RNA is detected coating both X chromosomes in XX embryos or even the X chromosome in XY human and macaque embryos for several days (Okamoto et al., 2011, 2021; Vallot et al., 2017). It is still unclear why and how XIST RNA accumulation is uncoupled from XCI, and what is the switch/trigger that allows these pre-XCI but XIST-associated states to eventually be resolved into random XCI in XX embryos and no XCI in XY embryos. Nevertheless, in human embryos, XIST presence might not be without consequences: it coincides with the downregulation of X-linked gene expression, which has been termed X-chromosome ‘dampening’ (XCD) (Petropoulos et al., 2016). Dampening does not appear to be present in other primates, such as the macaque or marmoset (Cidral et al., 2021; Okamoto et al., 2021) and remains contested in human (Moreira de Mello et al., 2017; Mandal et al., 2020). Recently, it has been shown in human ESCs that deletion of XIST leads to derepression of X-linked expression (suggesting that XIST is responsible for XCD) and that SPEN, the transcriptional repressor that is essential for initiating gene silencing during XCI (Dossin et al., 2020), is also involved (Alfeghaly et al., 2024; Dror et al., 2024). Again, what prevents the expression of XIST and recruitment of SPEN to lead to full XCI remains unknown.
New insights into the regulation of Xist across placental species
The regulation of Xist preceding random XCI also shows species-specific variations, at least, as evaluated by studies on cultured mouse and human cells. Xist is embedded in a regulatory landscape with many other noncoding loci (reviewed by Luchsinger-Morcelle et al., 2024), some of which evolved via pseudogenisation from protein-coding genes along with Xist (Chureau et al., 2002; Duret et al., 2006). Whether Tsix, which runs antisense to Xist and is essential for XCI regulation in mice, is as important in other species is an old debate (reviewed by Galupa and Heard, 2018). For now, genetic evidence to support or disprove such a role is still missing. More recently, Claire Rougeulle's lab has spearheaded functional analyses in primate ESCs of some of the noncoding neighbours of XIST, the JPX and FTX loci that, in mouse, are important positive regulators of Xist and random XCI (Furlan et al., 2018; Gjaltema et al., 2022; Sun et al., 2013; Tian et al., 2010). Interestingly, FTX functions are not conserved in human (Rosspopoff et al., 2023), and JPX is also a major regulator of XIST regulation in human but not in macaque or marmoset (Cazottes et al., 2023 preprint; Rosspopoff et al., 2023). Yet, between human and mouse there are differences in how Jpx/JPX regulates Xist/XIST. In mouse, the Jpx RNA mediates Xist regulation, at the post-transcriptional level; in human, the JPX RNA is dispensable for XIST regulation. Rather, it is JPX transcription that is required for proper XIST expression, in cis, probably via influencing RNA polymerase II recruitment to the XIST promoter (Rosspopoff et al., 2023). In macaque and marmoset ESCs, no indications have been found of ongoing transcription at syntenic positions of Tsix, Linx or Xite (also known as Rr18) (Cazottes et al., 2023 preprint), other noncoding loci recognised as significant Xist repressors in mice (Galupa et al., 2020; Hierholzer et al., 2022; Lee, 2000; Lee and Lu, 1999; Ogawa and Lee, 2003; Sado et al., 2001).
Unravelling the diverse mechanisms and regulatory strategies governing XCI across eutherian mammals sheds light on the intricate evolutionary dynamics of this dosage compensation process. As stated by Okamoto and colleagues, the existing diversity ‘probably reflects the fact that developmental processes are constantly changing during evolution and that the regulation of processes such as XCI have to display substantial plasticity to accommodate these changes’ (Okamoto et al., 2011).
A dance of upregulation and inactivation: evolutionary and developmental dynamics of XCU and XCI
Despite the sequential narrative of XCU and XCI in Ohno's hypothesis, they have had to evolve rather ‘simultaneously’ (and potentially influencing each other), as genes were lost from the proto-Y chromosome, creating a need for dosage-compensation mechanisms. It remains unclear how XCU and XCI evolved per se and in relation to each other, which is especially intriguing considering that XCU appears to operate on a gene-by-gene basis, whereas XCI is a chromosome-wide mechanism. The latter was perhaps not the case in the initial stages of sex-chromosome differentiation (Disteche, 2016; Gribnau and Grootegoed, 2012), in which dosage compensation in XX individuals might have happened on a gene-by-gene basis, rather resembling what is observed in present-day monotremes (Box 3).
Although the evolutionary dynamics remain enigmatic, progress has been made recently regarding the developmental dynamics of XCU and XCI (Fig. 1B). Based on allele-resolved single-cell RNA sequencing (scRNA-seq) of mouse embryos and embryonic stem cells, XCU has been proposed to occur on one of the two X chromosomes while the second one is undergoing XCI (Lentini et al., 2022). In the authors' words, ‘a flexible process that tunes RNA synthesis proportionally to the output of the second X allele across developmental states’ (Lentini et al., 2022). For this reason, the authors called it an ‘elastic’ process of dosage compensation, as opposed to the X chromosome(s) being constantly upregulated (before XCI) or upregulated as a single developmental event. In XY embryos, XCU is established upon ZGA and maintained throughout development, whereas in XX embryos XCU accompanies XCI: it occurs initially upon ZGA along with imprinted XCI, but is then reversed in embryonic lineages as the inactive X reactivates, and established again along with random XCI (Lentini et al., 2022) (Fig. 1B). Another recent study has also found that XCU is dynamically linked to random XCI (Naik et al., 2022). XCU thus happens in response to imbalanced X dosage, which has been further supported by reanalysis of allele-resolved scRNA-seq from Xist knockout embryos (Borensztein et al., 2017); in the absence of XCI, no XCU was initiated (Lentini et al., 2022). Overall, these findings (especially the timings at which XCU occurs) contrast with observations made in some previous studies (Table 1), probably because these previous studies did not take into account the allelic origin of X-linked gene expression, which can be confounded by processes such as XCI and ZGA (e.g. XCI and XCU on opposing alleles may cancel out if analysing only cumulative RNA level). Allele-resolved scRNA-seq analyses during macaque embryogenesis revealed similar findings as in mice, showing that in XX embryos XCU occurs along with or after XCI, whereas in XY embryos it takes place progressively from the first stages analysed (Okamoto et al., 2021).
Such elastic XCU implies a dosage-sensing mechanism coupling XCU and XCI in XX embryos. The same authors (Lentini et al., 2022) proposed that this could be achieved through a progressive shift of transcription factors to the active X from the inactive X territory, from which they are excluded as the inactive-X repressive compartment is formed (Chaumeil et al., 2006; Collombet et al., 2023). How XCU might be coupled to ZGA in XY embryos is less clear.
Which X-linked genes need to be dosage-compensated?
The assumption underlying the importance of dosage-compensation mechanisms is that their absence leads to detrimental phenotypes. So far it is not possible to manipulate XCU to test its importance, given that we still know so little about its mechanisms, but XCI instead can be abolished by knocking-out its major regulator. Failure to undergo XCI upon Xist deletion during early mouse development has revealed genome-wide changes in gene expression and embryonic lethality due to defects in extra-embryonic tissues (Borensztein et al., 2017; Marahrens et al., 1997; Mugford et al., 2012). Such a phenotype is likely an additive (or synergistic) result of many X-linked genes not being dosage-compensated. But throughout the evolution of mammalian sex chromosomes, as it is likely that genes on the proto-X were dosage-compensated gradually while genes on the proto-Y were being gradually lost, did all genes on the X have to be dosage-compensated? Has the evolution of dosage-compensation mechanisms been ‘driven predominantly by a need to equalise overall X-linked and autosomal expression levels’ or do ‘transcript levels of key individual genes exert the major selection pressure’ (Lin et al., 2007)? We know that, for many genes, heterozygous mutations are well tolerated. Other genes, on the contrary, show haploinsufficiency (deletion intolerance) or triplosensitivity (duplication intolerance) (Collins et al., 2022). It therefore appears to be reasonable that dosage-sensitive genes on proto-sex chromosomes have been the main drivers for the emergence of dosage-compensation mechanisms during sex-chromosome evolution. Supporting this notion, Pessia and colleagues have shown that, for X-linked genes presumably more dosage-sensitive (coding for members of large protein complexes, with ≥7 proteins), dosage compensation is more prevalent. Such genes showed higher expression ratios between the X and autosomes when compared with others coding for smaller protein complexes (Pessia et al., 2012), suggesting, therefore, that dosage compensation for such genes is more required.
Dosage-sensitive genes typically encode factors for which there are stoichiometry constraints, such as subunits of large complexes, transcription factors, members of signal transduction pathways or microRNAs (Basilicata and Keller Valsecchi, 2021; Birchler and Veitia, 2007; Desvignes et al., 2021; Meunier et al., 2013; Veitia and Birchler, 2022). Based on curated genomic data (Blake et al., 2021) and gene ontology analysis (Ashburner et al., 2000; The Gene Ontology Consortium, 2021), we have determined that the mouse X chromosome harbours 92 miRNA loci and 507 genes involved in processes related to signalling and/or transcription and/or that code for components of protein complexes. Which ones correspond de facto to dosage-sensitive genes remains to be determined; unfortunately X-linked genes (and Y-linked) are often excluded from genome-wide analyses, as was the case for the recent catalogue of human dosage-sensitive genes (Collins et al., 2022). Some authors have proposed that the X is depleted of dosage-sensitive genes (the ‘insensitive X hypothesis’), based on analyses that determine that X-linked genes are less dosage-sensitive than autosomal genes, and that dosage-sensitive housekeeping genes are preferentially located on autosomes (Chen et al., 2020; Lin et al., 2012; Yang and Chen, 2019). This could reflect the initial gene content of the proto-sex chromosomes; in fact, some chromosomes are thought to be better suited to become sex chromosomes, based on their gene content, and sex chromosomes do tend to originate from autosomes that are overall insensitive to dose changes (Bachtrog et al., 2011; Disteche, 2016; Livernois et al., 2012). Interestingly, the emergence of dosage-compensation mechanisms can in turn influence the evolution of the sex chromosomes themselves in terms of gene content (Box 4).
This does not exclude that particularly dosage-sensitive genes on the X have favoured the evolution of dosage-compensation mechanisms. For example, the X harbours the gene SMC1A, a subunit of the cohesin complex, which mediates sister chromatid cohesion, homologous recombination and DNA looping. Other cohesin subunits, such as Nipbl, are haploinsufficient (Mills et al., 2018). Heterozygous mutations in SMC1A itself have also been associated with haploinsufficiency underlying Cornelia de Lange syndrome (Deardorff et al., 2007; Musio et al., 2006). Another example of highly likely dosage-sensitive X-linked genes are MED14 and MED12, both subunits of the mediator complex, which play essential functions in eukaryotic transcription; higher levels of MED14 as a result of abnormal X-reactivation have been associated with impairment of mammary stem cell differentiation and increased tumorigenicity (Richart et al., 2022). Importantly, the mammalian X chromosome consists of a mix of ancestral genes and more recently acquired ones, and it has been suggested that these two gene groups might have different compensation requirements and potentially involve distinct regulatory mechanisms (Deng et al., 2013).
Conclusion
Understanding the regulation and compensation of mammalian gene dosage has clearly provided us with many new insights into development and evolution, which extends our understanding of physiology and pathology. For example, XIST transgenics has gained interest as a possible therapeutic tool for chromosome dosage disorders, such as Down syndrome (Gupta et al., 2024; Moyer et al., 2021). An aspect that we overlooked in this Review is that a subset of X-linked genes escapes XCI (Carrel et al., 1999), meaning they show biallelic expression in XX individuals. For some of these genes this translates to higher dosage compared with XY individuals, whereas for others this could be the means for dosage compensation, as they have homologues in the Y chromosome (Box 2). These escaping genes (including XIST) underlie sex-biassed susceptibility to certain diseases, such as autoimmune diseases (Dou et al., 2024; Forsyth et al., 2024; Hagen et al., 2020; Souyris et al., 2018; Youness et al., 2021). Interestingly, recent studies have shown how expression from the inactive X can modulate gene expression from the active X and autosomes (San Roman et al., 2023, 2024; Topa et al., 2024; Zhang et al., 2024).
Many important questions remain unanswered. As incomplete dosage compensation is well tolerated among some vertebrates, such as birds, what makes therian mammals (especially placental mammals) more sensitive to dosage differences, underlying the need for tight, chromosome-wide dosage-compensation mechanisms? Could this be related to constraints on the placenta and/or other extra-embryonic tissues? It is interesting to note that early embryonic lethality in mouse mutants is very often associated with severe placental malformations (Perez-Garcia et al., 2018). Another important open question is the extent to which the dosage sensitivity of a given gene will depend on the cell type or tissue where it is expressed.
New progress will certainly be achieved with the continuous improvement of technologies that allow us to quantify gene expression beyond transcription levels, such as quantitative proteomics (Schubert et al., 2017), or to quantitatively modulate gene expression (Ma et al., 2024; Naqvi et al., 2023; Noviello et al., 2023). Additionally, new embryonic systems in vitro hold promise to enable the exploration of a higher number of mammalian species (Handford et al., 2024; Lázaro et al., 2024), as well as to allow us to start functionally testing hypotheses about the dynamics and timing of dosage compensation. We look forward to the upcoming exciting times for dosage compensation research.
Acknowledgements
We thank Estelle Nicolas for critical feedback on the manuscript and Ikuhiro Okamoto for critical feedback on the figures. We apologise to authors whose work we overlooked or did not discuss or cite due to limits in the number of references. We thank the anonymous reviewers for pointing us to additional literature and for their constructive feedback. Figures were prepared with BioRender.com.
Footnotes
Funding
D.C. is supported by a fellowship from Ligue Contre le Cancer (LNCC_TAJT25850) and R.G. holds a tenured research position from the Centre National de la Recherche Scientifique (France). Research in the Galupa lab is supported by a grant from the Fondation pour la Recherche Médicale (AJE202305017142). Open Access funding provided by Fondation pour la Recherche Médicale. Deposited in PMC for immediate release.
References
- Alfeghaly C., Castel G., Cazottes E., Moscatelli M., Moinard E., Casanova M., Boni J., Mahadik K., Lammers J., Freour T.et al. (2024). XIST dampens X chromosome activity in a SPEN-dependent manner during early human development. Nat. Struct. Mol. Biol. [Epub ahead of print]. 10.1038/s41594-024-01325-3 [Abstract] [CrossRef] [Google Scholar]
- Arakawa Y., Nishida-Umehara C., Matsuda Y., Sutou S. and Suzuki H. (2002). X-chromosomal localization of mammalian Y-linked genes in two XO species of the Ryukyu spiny rat. Cytogenet. Genome Res. 99, 303-309. 10.1159/000071608 [Abstract] [CrossRef] [Google Scholar]
- Ashburner M., Ball C. A., Blake J. A., Botstein D., Butler H., Cherry J. M., Davis A. P., Dolinski K., Dwight S. S., Eppig J. T.et al. (2000). Gene Ontology: tool for the unification of biology. Nat. Genet. 25, 25-29. 10.1038/75556 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Bachtrog D., Kirkpatrick M., Mank J. E., McDaniel S. F., Pires J. C., Rice W. and Valenzuela N. (2011). Are all sex chromosomes created equal? Trends Genet. 27, 350-357. 10.1016/j.tig.2011.05.005 [Abstract] [CrossRef] [Google Scholar]
- Barr M. L. and Bertram E. G. (1949). A morphological distinction between neurones of the male and female, and the behaviour of the nucleolar satellite during accelerated nucleoprotein synthesis. Nature 163, 676. 10.1038/163676a0 [Abstract] [CrossRef] [Google Scholar]
- Basilicata M. F. and Keller Valsecchi C. I. (2021). The good, the bad, and the ugly: Evolutionary and pathological aspects of gene dosage alterations. PLoS Genet. 17, e1009906. 10.1371/journal.pgen.1009906 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Beckelmann J., Budik S., Bartel C. and Aurich C. (2012). Evaluation of Xist expression in preattachment equine embryos. Theriogenology 78, 1429-1436. 10.1016/j.theriogenology.2012.05.026 [Abstract] [CrossRef] [Google Scholar]
- Bellott D. W. and Page D. C. (2021). Dosage-sensitive functions in embryonic development drove the survival of genes on sex-specific chromosomes in snakes, birds, and mammals. Genome Res. 31, 198-210. 10.1101/gr.268516.120 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- 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. (2014). Mammalian Y chromosomes retain widely expressed dosage-sensitive regulators. Nature 508, 494-499. 10.1038/nature13206 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Bergero R. and Charlesworth D. (2009). The evolution of restricted recombination in sex chromosomes. Trends Ecol. Evol. 24, 94-102. 10.1016/j.tree.2008.09.010 [Abstract] [CrossRef] [Google Scholar]
- Bininda-Emonds O. R. P., Cardillo M., Jones K. E., MacPhee R. D. E., Beck R. M. D., Grenyer R., Price S. A., Vos R. A., Gittleman J. L. and Purvis A. (2007). The delayed rise of present-day mammals. Nature 446, 507-512. 10.1038/nature05634 [Abstract] [CrossRef] [Google Scholar]
- Birchler J. A. and Veitia R. A. (2007). The gene balance hypothesis: From classical genetics to modern genomics. Plant Cell 19, 395-402. 10.1105/tpc.106.049338 [Abstract] [CrossRef] [Google Scholar]
- Birchler J. A. and Veitia R. A. (2010). The gene balance hypothesis: implications for gene regulation, quantitative traits and evolution. New Phytol. 186, 54-62. 10.1111/j.1469-8137.2009.03087.x [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Birchler J. A. and Veitia R. A. (2012). Gene balance hypothesis: connecting issues of dosage sensitivity across biological disciplines. Proc. Natl. Acad. Sci. U.S.A. 109, 14746-14753. 10.1073/pnas.1207726109 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Birchler J. A. and Veitia R. A. (2021). One hundred years of gene balance: how stoichiometric issues affect gene expression, genome evolution, and quantitative traits. Cytogenet. Genome Res. 161, 529-550. 10.1159/000519592 [Abstract] [CrossRef] [Google Scholar]
- Blakeslee, A. F. (1934). New jimson weeds from old chromosomes. J. Hered 25, 81-108. 10.1093/oxfordjournals.jhered.a103898 [CrossRef] [Google Scholar]
- Blakeslee A. F. and Belling J. (1924). Chromosomal mutations in the jimson weed, datura stramonium. J. Hered 15, 195-206. 10.1093/oxfordjournals.jhered.a102447 [CrossRef] [Google Scholar]
- Blakeslee A. F., Belling J. and Farnham M. E. (1920). Chromosomal duplication and mendelian phenomena in datura mutants. Science 52, 388-390. 10.1126/science.52.1347.388 [Abstract] [CrossRef] [Google Scholar]
- Blake J. A., Baldarelli R., Kadin J. A., Richardson J. E., Smith C. L., Bult C. J.. and Mouse Genome Database Group (2021). Mouse Genome Database (MGD): Knowledgebase for mouse-human comparative biology. Nucleic Acids Res. 49, D981-D987. 10.1093/nar/gkaa1083 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Blewitt, M. E. (2024). Mary Lyon and the birth of X-inactivation research. Nat. Rev. Genet. 25, 6. 10.1038/s41576-023-00655-0 [Abstract] [CrossRef] [Google Scholar]
- Borensztein M., Syx L., Ancelin K., Diabangouaya P., Picard C., Liu T., Liang J.-B., Vassilev I., Galupa R., Servant N.et al. (2017). Xist-dependent imprinted X inactivation and the early developmental consequences of its failure. Nat. Struct. Mol. Biol. 24, 226-233. 10.1038/nsmb.3365 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Brashear W. A., Bredemeyer K. R. and Murphy W. J. (2021). Genomic architecture constrained placental mammal X Chromosome evolution. Genome Res. 31, 1353-1365. 10.1101/gr.275274.121 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Bridges, C. B. (1921). Triploid intersexes in drosophila melanogaster. Science 54, 252-254. 10.1126/science.54.1394.252 [Abstract] [CrossRef] [Google Scholar]
- Carelli F. N., Hayakawa T., Go Y., Imai H., Warnefors M. and Kaessmann H. (2016). The life history of retrocopies illuminates the evolution of new mammalian genes. Genome Res. 26, 301-314. 10.1101/gr.198473.115 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Carrel L., Cottle A. A., Goglin K. C. and Willard H. F. (1999). A first-generation X-inactivation profile of the human X chromosome. Proc. Natl. Acad. Sci. U.S.A. 96, 14440-14444. 10.1073/pnas.96.25.14440 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Cazottes E., Alfeghaly C., Rognard C., Loda A., Castel G., Villacorta L., Dong M., Heard E., Aksoy I., Savatier P.et al. (2023). Extensive remodelling of XIST regulatory networks during primate evolution. bioRxiv. [Google Scholar]
- Chandra, H. S. (1985). Is human X chromosome inactivation a sex-determining device? Proc. Natl. Acad. Sci. U.S.A. 82, 6947-6949. 10.1073/pnas.82.20.6947 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Chandra, H. S. (2022). Mammalian X-chromosome inactivation: proposed role in suppression of the male programme in genetic females. J. Genet. 101, 23. 10.1007/s12041-022-01363-0 [Abstract] [CrossRef] [Google Scholar]
- Charlesworth D., Charlesworth B. and Marais G. (2005). Steps in the evolution of heteromorphic sex chromosomes. Heredity 95, 118-128. 10.1038/sj.hdy.6800697 [Abstract] [CrossRef] [Google Scholar]
- Chaumeil J., Le Baccon P., Wutz A. and Heard E. (2006). A novel role for Xist RNA in the formation of a repressive nuclear compartment into which genes are recruited when silenced. Genes Dev. 20, 2223-2237. 10.1101/gad.380906 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Chen X. and Zhang J. (2015). No X-chromosome dosage compensation in human proteomes. Mol. Biol. Evol. 32, 1456-1460. 10.1093/molbev/msv036 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Chen X. and Zhang J. (2016). The X to autosome expression ratio in haploid and diploid human embryonic stem cells. Mol. Biol. Evol. 33, 3104-3107. 10.1093/molbev/msw187. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Chen J., Wang M., He X., Yang J.-R. and Chen X. (2020). The evolution of sex chromosome dosage compensation in animals. J. Genet. Genomics 47, 681-693. 10.1016/j.jgg.2020.10.005 [Abstract] [CrossRef] [Google Scholar]
- Chibalina M. V. and Filatov D. A. (2011). Plant Y chromosome degeneration is retarded by haploid purifying selection. Curr. Biol. 21, 1475-1479. 10.1016/j.cub.2011.07.045 [Abstract] [CrossRef] [Google Scholar]
- Chue J. and Smith C. A. (2011). Sex determination and sexual differentiation in the avian model. FEBS J. 278, 1027-1034. 10.1111/j.1742-4658.2011.08032.x [Abstract] [CrossRef] [Google Scholar]
- Chureau C., Prissette M., Bourdet A., Barbe V., Cattolico L., Jones L., Eggen A., Avner P. and Duret L. (2002). Comparative sequence analysis of the X-inactivation center region in mouse, human, and bovine. Genome Res. 12, 894-908. 10.1101/gr.152902 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Cidral A. L., de Mello J. C. M., Gribnau J. and Pereira L. V. (2021). Concurrent X chromosome inactivation and upregulation during non-human primate preimplantation development revealed by single-cell RNA-sequencing. Sci. Rep. 11, 9624. 10.1038/s41598-021-89175-7 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Collignon J., Sockanathan S., Hacker A., Cohen-Tannoudji M., Norris D., Rastan S., Stevanovic M., Goodfellow P. N. and Lovell-Badge R. (1996). A comparison of the properties of Sox-3 with Sry and two related genes, Sox-1 and Sox-2. Development 122, 509-520. 10.1242/dev.122.2.509 [Abstract] [CrossRef] [Google Scholar]
- Collins R. L., Glessner J. T., Porcu E., Lepamets M., Brandon R., Lauricella C., Han L., Morley T., Niestroj L.-M., Ulirsch J.et al. (2022). A cross-disorder dosage sensitivity map of the human genome. Cell 185, 3041-3055.e25. 10.1016/j.cell.2022.06.036 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Collombet S., Rall I., Dugast-Darzacq C., Heckert A., Halavatyi A., Le Saux A., Dailey G., Darzacq X. and Heard E. (2023). RNA polymerase II depletion from the inactive X chromosome territory is not mediated by physical compartmentalization. Nat. Struct. Mol. Biol. 30, 1216-1223. 10.1038/s41594-023-01008-5 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Copur Ö., Gorchakov A., Finkl K., Kuroda M. I. and Müller J. (2018). Sex-specific phenotypes of histone H4 point mutants establish dosage compensation as the critical function of H4K16 acetylation in Drosophila. Proc. Natl. Acad. Sci. U.S.A. 115, 13336-13341. 10.1073/pnas.1817274115 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Cortez D., Marin R., Toledo-Flores D., Froidevaux L., Liechti A., Waters P. D., Grützner F. and Kaessmann H. (2014). Origins and functional evolution of Y chromosomes across mammals. Nature 508, 488-493. 10.1038/nature13151 [Abstract] [CrossRef] [Google Scholar]
- Davidow L. S., Breen M., Duke S. E., Samollow P. B., McCarrey J. R. and Lee J. T. (2007). The search for a marsupial XIC reveals a break with vertebrate synteny. Chromosome Res. 15, 137-146. 10.1007/s10577-007-1121-6 [Abstract] [CrossRef] [Google Scholar]
- Deakin J. E., Hore T. A., Koina E. and Marshall Graves J. A. (2008). The status of dosage compensation in the multiple X chromosomes of the platypus. PLoS Genet. 4, e1000140. 10.1371/journal.pgen.1000140 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Deakin J. E., Chaumeil J., Hore T. A. and Marshall Graves J. A. (2009). Unravelling the evolutionary origins of X chromosome inactivation in mammals: insights from marsupials and monotremes. Chromosome Res. 17, 671-685. 10.1007/s10577-009-9058-6 [Abstract] [CrossRef] [Google Scholar]
- Deakin J. E., Delbridge M. L., Koina E., Harley N., Alsop A. E., Wang C., Patel V. S. and Graves J. A. M. (2013). Reconstruction of the ancestral marsupial karyotype from comparative gene maps. BMC Evol. Biol. 13, 258. 10.1186/1471-2148-13-258 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Deardorff M. A., Kaur M., Yaeger D., Rampuria A., Korolev S., Pie J., Gil-Rodríguez C., Arnedo M., Loeys B., Kline A. D.et al. (2007). Mutations in cohesin complex members SMC3 and SMC1A cause a mild variant of cornelia de Lange syndrome with predominant mental retardation. Am. J. Hum. Genet. 80, 485-494. 10.1086/511888 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- DeCasien A., Tsai K., Liu S., Thomas A. and Raznahan A. (2023). Linking X-Y Gametologue co-expression patterns to sex differences in disease. Biol. Psych. 93, S93. 10.1016/j.biopsych.2023.02.240 [CrossRef] [Google Scholar]
- Delgado C. L. R., Waters P. D., Gilbert C., Robinson T. J. and Graves J. A. M. (2009). Physical mapping of the elephant X chromosome: conservation of gene order over 105 million years. Chromosome Res. 17, 917-926. 10.1007/s10577-009-9079-1 [Abstract] [CrossRef] [Google Scholar]
- Deng X., Hiatt J. B., Nguyen D. K., Ercan S., Sturgill D., Hillier L. W., Schlesinger F., Davis C. A., Reinke V. J., Gingeras T. R.et al. (2011). Evidence for compensatory upregulation of expressed X-linked genes in mammals, Caenorhabditis elegans and Drosophila melanogaster. Nat. Genet. 43, 1179-1185. 10.1038/ng.948 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Deng X., Berletch J. B., Ma W., Nguyen D. K., Hiatt J. B., Noble W. S., Shendure J. and Disteche C. M. (2013). Mammalian X upregulation is associated with enhanced transcription initiation, RNA half-life, and MOF-mediated H4K16 acetylation. Dev. Cell 25, 55-68. 10.1016/j.devcel.2013.01.028 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Desvignes T., Sydes J., Montfort J., Bobe J. and Postlethwait J. H. (2021). Evolution after whole-genome duplication: teleost MicroRNAs. Mol. Biol. Evol. 38, 3308-3331. 10.1093/molbev/msab105 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Dindot S. V., Farin P. W., Farin C. E., Romano J., Walker S., Long C. and Piedrahita J. A. (2004). Epigenetic and genomic imprinting analysis in nuclear transfer derived Bos gaurus/Bos taurus hybrid fetuses. Biol. Reprod. 71, 470-478. 10.1095/biolreprod.103.025775 [Abstract] [CrossRef] [Google Scholar]
- Disteche, C. M. (2016). Dosage compensation of the sex chromosomes and autosomes. Semin. Cell Dev. Biol. 56, 9-18. 10.1016/j.semcdb.2016.04.013 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Disteche C. M., Filippova G. N. and Tsuchiya K. D. (2003). Escape com X inactivation. Cytogen. Cell Genet. 99, 36-43. 10.1159/000071572 [Abstract] [CrossRef] [Google Scholar]
- Dossin F., Pinheiro I., Żylicz J. J., Roensch J., Collombet S., Le Saux A., Chelmicki T., Attia M., Kapoor V., Zhan Y.et al. (2020). SPEN integrates transcriptional and epigenetic control of X-inactivation. Nature 578, 455-460. 10.1038/s41586-020-1974-9 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Dou D. R., Zhao Y., Belk J. A., Zhao Y., Casey K. M., Chen D. C., Li R., Yu B., Srinivasan S., Abe B. T.et al. (2024). Xist ribonucleoproteins promote female sex-biased autoimmunity. Cell 187, 733-749.e16. 10.1016/j.cell.2023.12.037 [Abstract] [CrossRef] [Google Scholar]
- Dror I., Chitiashvili T., Tan S. Y. X., Cano C. T., Sahakyan A., Markaki Y., Chronis C., Collier A. J., Deng W., Liang G.et al. (2024). XIST directly regulates X-linked and autosomal genes in naive human pluripotent cells. Cell 187, 110-129.e31. 10.1016/j.cell.2023.11.033 [Abstract] [CrossRef] [Google Scholar]
- Dubansky, B. (2018). The interaction of environment and chronological and developmental time. In Development and Environment (ed. Burggren W. and Dubansky B.), pp. 9-39. Cham: Springer International Publishing. [Google Scholar]
- Duret L., Chureau C., Samain S., Weissenbach J. and Avner P. (2006). The Xist RNA gene evolved in eutherians by pseudogenization of a protein-coding gene. Science 312, 1653-1655. 10.1126/science.1126316 [Abstract] [CrossRef] [Google Scholar]
- Elisaphenko E. A., Kolesnikov N. N., Shevchenko A. I., Rogozin I. B., Nesterova T. B., Brockdorff N. and Zakian S. M. (2008). A dual origin of the Xist gene from a protein-coding gene and a set of transposable elements. PLoS ONE 3, e2521. 10.1371/journal.pone.0002521 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Engelstädter J. and Haig D. (2008). Sexual antagonism and the evolution of X chromosome inactivation. Evolution 62, 2097-2104. 10.1111/j.1558-5646.2008.00431.x [Abstract] [CrossRef] [Google Scholar]
- Fabre P.-H., Hautier L., Dimitrov D. and Douzery E. J. P. (2012). A glimpse on the pattern of rodent diversification: a phylogenetic approach. BMC Evol. Biol. 12, 88. 10.1186/1471-2148-12-88 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Fallahshahroudi A., Rodriguez-Montes L., Yousefi Taemeh S., Trost N., Tellez M., Ballantyne M., Idoko-Akoh A., Taylor L., Sherman A., Sorato E.et al. (2024). A male-essential microRNA is key for avian sex chromosome dosage compensation. bioRxiv. [Google Scholar]
- Faucillion M.-L. and Larsson J. (2015). Increased expression of X-linked genes in mammals is associated with a higher stability of transcripts and an increased ribosome density. Genome Biol. Evol. 7, 1039-1052. 10.1093/gbe/evv054 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Felsenstein, J. (1974). The evolutionary advantage of recombination. Genetics 78, 737-756. 10.1093/genetics/78.2.737 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Forsyth K. S., Jiwrajka N., Lovell C. D., Toothacre N. E. and Anguera M. C. (2024). The conneXion between sex and immune responses. Nat. Rev. Immunol. 24, 487-502. 10.1038/s41577-024-00996-9 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Foster J. W. and Graves J. A. (1994). An SRY-related sequence on the marsupial X chromosome: implications for the evolution of the mammalian testis-determining gene. Proc. Natl. Acad. Sci. U.S.A. 91, 1927-1931. 10.1073/pnas.91.5.1927 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Foster J. W., Brennan F. E., Hampikian G. K., Goodfellow P. N., Sinclair A. H., Lovell-Badge R., Selwood L., Renfree M. B., Cooper D. W. and Graves J. A. (1992). Evolution of sex determination and the Y chromosome: SRY-related sequences in marsupials. Nature 359, 531-533. 10.1038/359531a0 [Abstract] [CrossRef] [Google Scholar]
- Frankish A., Carbonell-Sala S., Diekhans M., Jungreis I., Loveland J. E., Mudge J. M., Sisu C., Wright J. C., Arnan C., Barnes I.et al. (2023). GENCODE: reference annotation for the human and mouse genomes in 2023. Nucleic Acids Res. 51, 942-949. 10.1093/nar/gkac1071 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Furlan G. and Galupa R. (2022). Mechanisms of choice in X-chromosome inactivation. Cells 11, 535. 10.3390/cells11030535 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Furlan G., Gutierrez Hernandez N., Huret C., Galupa R., van Bemmel J. G., Romito A., Heard E., Morey C. and Rougeulle C. (2018). The ftx noncoding locus controls X chromosome inactivation independently of its RNA products. Mol. Cell 70, 462-472. 10.1016/j.molcel.2018.03.024 [Abstract] [CrossRef] [Google Scholar]
- Galupa R. and Heard E. (2018). X-Chromosome inactivation: a crossroads between chromosome architecture and gene regulation. Annu. Rev. Genet. 52, 535-566. 10.1146/annurev-genet-120116-024611 [Abstract] [CrossRef] [Google Scholar]
- Galupa R., Nora E. P., Worsley-Hunt R., Picard C., Gard C., van Bemmel J. G., Servant N., Zhan Y., El Marjou F., Johanneau C.et al. (2020). A conserved noncoding locus regulates random monoallelic Xist expression across a topological boundary. Mol. Cell 77, 352-367. 10.1016/j.molcel.2019.10.030 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Garcia-Ojalvo J. and Bulut-Karslioglu A. (2023). On time: developmental timing within and across species. Development 150, dev201045. 10.1242/dev.201045 [Abstract] [CrossRef] [Google Scholar]
- Gjaltema R. A. F., Schwämmle T., Kautz P., Robson M., Schöpflin R., Lustig L. R., Brandenburg L., Dunkel I., Vechiatto C., Ntini E.et al. (2022). Distal and proximal cis-regulatory elements sense X chromosome dosage and developmental state at the Xist locus. Mol. Cell 82, 190-208. 10.1016/j.molcel.2021.11.023 [Abstract] [CrossRef] [Google Scholar]
- Goszczynski D. E., Tinetti P. S., Choi Y. H., Ross P. J. and Hinrichs K. (2021). Allele-specific expression analysis reveals conserved and unique features of preimplantation development in equine ICSI embryos†. Biol. Reprod 105, 1416-1426. 10.1093/biolre/ioab174 [Abstract] [CrossRef] [Google Scholar]
- Grant J., Mahadevaiah S. K., Khil P., Sangrithi M. N., Royo H., Duckworth J., McCarrey J. R., VandeBerg J. L., Renfree M. B., Taylor W.et al. (2012). Rsx is a metatherian RNA with Xist-like properties in X-chromosome inactivation. Nature 487, 254-258. 10.1038/nature11171 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Graves, J. A. M. (2006). Sex chromosome specialization and degeneration in mammals. Cell 124, 901-914. 10.1016/j.cell.2006.02.024 [Abstract] [CrossRef] [Google Scholar]
- Graves J. A. M., Gécz J. and Hameister H. (2002). Evolution of the human X--a smart and sexy chromosome that controls speciation and development. Cytogenet. Genome Res. 99, 141-145. 10.1159/000071585 [Abstract] [CrossRef] [Google Scholar]
- Gribnau J. and Grootegoed J. A. (2012). Origin and evolution of X chromosome inactivation. Curr. Opin. Cell Biol. 24, 397-404. 10.1016/j.ceb.2012.02.004 [Abstract] [CrossRef] [Google Scholar]
- Grützner F., Rens W., Tsend-Ayush E., El-Mogharbel N., O'Brien P. C. M., Jones R. C., Ferguson-Smith M. A. and Marshall Graves J. A. (2004). In the platypus a meiotic chain of ten sex chromosomes shares genes with the bird Z and mammal X chromosomes. Nature 432, 913-917. 10.1038/nature03021 [Abstract] [CrossRef] [Google Scholar]
- Gubbay J., Collignon J., Koopman P., Capel B., Economou A., Münsterberg A., Vivian N., Goodfellow P. and Lovell-Badge R. (1990). A gene mapping to the sex-determining region of the mouse Y chromosome is a member of a novel family of embryonically expressed genes. Nature 346, 245-250. 10.1038/346245a0 [Abstract] [CrossRef] [Google Scholar]
- Gupta V., Parisi M., Sturgill D., Nuttall R., Doctolero M., Dudko O. K., Malley J. D., Eastman P.S. and Oliver B. (2006). Global analysis of X-chromosome dosage compensation. J. Biol. 5, 3. 10.1186/jbiol30 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Gupta K., Czerminski J. T. and Lawrence J. B. (2024). Trisomy silencing by XIST: translational prospects and challenges. Hum. Genet. 10.1007/s00439-024-02651-8 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Gu L. and Walters J. R. (2017). Evolution of sex chromosome dosage compensation in animals: a beautiful theory, undermined by facts and bedeviled by details. Genome Biol. Evol. 9, 2461-2476. 10.1093/gbe/evx154 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Gu L., Reilly P. F., Lewis J. J., Reed R. D., Andolfatto P. and Walters J. R. (2019). Dichotomy of dosage compensation along the Neo Z chromosome of the monarch butterfly. Curr. Biol. 29, 4071-4077. 10.1016/j.cub.2019.09.056 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Hagen S. H., Henseling F., Hennesen J., Savel H., Delahaye S., Richert L., Ziegler S. M. and Altfeld M. (2020). Heterogeneous escape from X chromosome inactivation results in sex differences in type I IFN responses at the single human pDC Level. Cell Rep. 33, 108485. 10.1016/j.celrep.2020.108485 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Haig, D. (2006). Self-imposed silence: parental antagonism and the evolution of X-chromosome inactivation. Evolution 60, 440-447. [Abstract] [Google Scholar]
- Handford C. E., Junyent S., Jorgensen V. and Zernicka-Goetz M. (2024). Topical section: embryonic models (2023) for current opinion in genetics & development. Curr. Opin. Genet. Dev. 84, 102134. 10.1016/j.gde.2023.102134 [Abstract] [CrossRef] [Google Scholar]
- Heard E. and Rougeulle C. (2021). Digging into X chromosome inactivation. Science 374, 942-943. 10.1126/science.abm1857 [Abstract] [CrossRef] [Google Scholar]
- He X., Chen X., Xiong Y., Chen Z., Wang X., Shi S., Wang X. and Zhang J. (2011). He et al. reply. Nat. Genet. 43, 1171-1172. 10.1038/ng.1010 [CrossRef] [Google Scholar]
- Hierholzer A., Chureau C., Liverziani A., Ruiz N. B., Cattanach B. M., Young A. N., Kumar M., Cerase A. and Avner P. (2022). A long noncoding RNA influences the choice of the X chromosome to be inactivated. Proc. Natl. Acad. Sci. U.S.A. 119, e2118182119. 10.1073/pnas.2118182119 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Hore T. A., Koina E., Wakefield M. J. and Marshall Graves J. A. (2007). The region homologous to the X-chromosome inactivation centre has been disrupted in marsupial and monotreme mammals. Chromosome Res. 15, 147-161. 10.1007/s10577-007-1119-0 [Abstract] [CrossRef] [Google Scholar]
- Hounkpe B. W., Chenou F., de Lima F. and De Paula E. V. (2021). HRT Atlas v1.0 database: redefining human and mouse housekeeping genes and candidate reference transcripts by mining massive RNA-seq datasets. Nucleic Acids Res. 49, D947-D955. 10.1093/nar/gkaa609 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Hughes J. F., Skaletsky H., Koutseva N., Pyntikova T. and Page D. C. (2015). Sex chromosome-to-autosome transposition events counter Y-chromosome gene loss in mammals. Genome Biol. 16, 104. 10.1186/s13059-015-0667-4 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Huylmans A. K., Macon A. and Vicoso B. (2017). Global dosage compensation is ubiquitous in lepidoptera, but counteracted by the masculinization of the Z chromosome. Mol. Biol. Evol. 34, 2637-2649. 10.1093/molbev/msx190 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Ioannidis J., Taylor G., Zhao D., Liu L., Idoko-Akoh A., Gong D., Lovell-Badge R., Guioli S., McGrew M. J. and Clinton M. (2021). Primary sex determination in birds depends on DMRT1 dosage, but gonadal sex does not determine adult secondary sex characteristics. Proc. Natl. Acad. Sci. U.S.A. 118, e2020909118. 10.1073/pnas.2020909118 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Iwasa Y. and Pomiankowski A. (2001). The evolution of X-linked genomic imprinting. Genetics 158, 1801-1809. 10.1093/genetics/158.4.1801 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Jordan W., Rieder L. E. and Larschan E. (2019). Diverse genome topologies characterize dosage compensation across species. Trends Genet. 35, 308-315. 10.1016/j.tig.2019.02.001 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Julien P., Brawand D., Soumillon M., Necsulea A., Liechti A., Schütz F., Daish T., Grützner F. and Kaessmann H. (2012). Mechanisms and evolutionary patterns of mammalian and avian dosage compensation. PLoS Biol. 10, e1001328. 10.1371/journal.pbio.1001328 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Kalita A. I., Marois E., Kozielska M., Weissing F. J., Jaouen E., Möckel M. M., Rühle F., Butter F., Basilicata M. F. and Valsecchi C. I. K. (2023). The sex-specific factor SOA controls dosage compensation in Anopheles mosquitoes. Nature 623, 175-182. 10.1038/s41586-023-06641-0 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Kanata E., Duffié R. and Schulz E. G. (2024). Establishment and maintenance of random monoallelic expression. Development 151, dev201741. 10.1242/dev.201741 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Kemkemer C., Kohn M., Kehrer-Sawatzki H., Fundele R. H. and Hameister H. (2009). Enrichment of brain-related genes on the mammalian X chromosome is ancient and predates the divergence of synapsid and sauropsid lineages. Chromosome Res. 17, 811-820. 10.1007/s10577-009-9072-8 [Abstract] [CrossRef] [Google Scholar]
- Keniry A. and Blewitt M. E. (2023). Chromatin-mediated silencing on the inactive X chromosome. Development 150, dev201742. 10.1242/dev.201742 [Abstract] [CrossRef] [Google Scholar]
- Kharchenko P. V., Xi R. and Park P. J. (2011). Evidence for dosage compensation between the X chromosome and autosomes in mammals. Nat. Genet. 43, 1167-1169; author reply 1171. 10.1038/ng.991 [Abstract] [CrossRef] [Google Scholar]
- Kim J., Farré M., Auvil L., Capitanu B., Larkin D. M., Ma J. and Lewin H. A. (2017). Reconstruction and evolutionary history of eutherian chromosomes. Proc. Natl. Acad. Sci. U.S.A. 114, E5379-E5388. [Europe PMC free article] [Abstract] [Google Scholar]
- Kondrashov, F. A. (2012). Gene duplication as a mechanism of genomic adaptation to a changing environment. Proc. Biol. Sci. 279, 5048-5057. [Europe PMC free article] [Abstract] [Google Scholar]
- Kuroiwa A., Ishiguchi Y., Yamada F., Shintaro A. and Matsuda Y. (2010). The process of a Y-loss event in an XO/XO mammal, the Ryukyu spiny rat. Chromosoma 119, 519-526. 10.1007/s00412-010-0275-8 [Abstract] [CrossRef] [Google Scholar]
- Larsson A. J. M., Coucoravas C., Sandberg R. and Reinius B. (2019). X-chromosome upregulation is driven by increased burst frequency. Nat. Struct. Mol. Biol. 26, 963-969. 10.1038/s41594-019-0306-y [Abstract] [CrossRef] [Google Scholar]
- Lau A. C., Zhu K. P., Brouhard E. A., Davis M. B. and Csankovszki G. (2016). An H4K16 histone acetyltransferase mediates decondensation of the X chromosome in C. elegans males. Epigenetics Chromatin 9, 44. 10.1186/s13072-016-0097-x [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Laudon D., Gostling N. J., Sengers B. G., Chavatte-Palmer P. and Lewis R. M. (2024). Placental evolution from a three-dimensional and multiscale structural perspective. Evolution 78, 13-25. 10.1093/evolut/qpad209 [Abstract] [CrossRef] [Google Scholar]
- Lázaro J., Sochacki J. and Ebisuya M. (2024). The stem cell zoo for comparative studies of developmental tempo. Curr. Opin. Genet. Dev. 84, 102149. 10.1016/j.gde.2023.102149 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Lee, J. T. (2000). Disruption of imprinted X inactivation by parent-of-origin effects at Tsix. Cell 103, 17-27. 10.1016/S0092-8674(00)00101-X [Abstract] [CrossRef] [Google Scholar]
- Lee J. T. and Lu N. (1999). Targeted mutagenesis of Tsix leads to nonrandom X inactivation. Cell 99, 47-57. 10.1016/S0092-8674(00)80061-6 [Abstract] [CrossRef] [Google Scholar]
- Leitão E., Schröder C., Parenti I., Dalle C., Rastetter A., Kühnel T., Kuechler A., Kaya S., Gérard B., Schaefer E.et al. (2022). Systematic analysis and prediction of genes associated with monogenic disorders on human chromosome X. Nat. Commun. 13, 6570. 10.1038/s41467-022-34264-y [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Lentini A. and Reinius B. (2023). Limitations of X:autosome ratio as a measurement of X-chromosome upregulation. Curr. Biol. 33, R395-R396. 10.1016/j.cub.2023.03.059 [Abstract] [CrossRef] [Google Scholar]
- Lentini A., Cheng H., Noble J. C., Papanicolaou N., Coucoravas C., Andrews N., Deng Q., Enge M. and Reinius B. (2022). Elastic dosage compensation by X-chromosome upregulation. Nat. Commun. 13, 1854. 10.1038/s41467-022-29414-1 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Lercher M. J., Urrutia A. O. and Hurst L. D. (2003). Evidence that the human X chromosome is enriched for male-specific but not female-specific genes. Mol. Biol. Evol. 20, 1113-1116. 10.1093/molbev/msg131 [Abstract] [CrossRef] [Google Scholar]
- Lewis S. E., Searle S. M. J., Harris N., Gibson M., Lyer V., Richter J., Wiel C., Bayraktaroglu L., Birney E., Crosby M. A.et al. (2002). Apollo: a sequence annotation editor. Genome Biol. 3, RESEARCH0082. 10.1186/gb-2002-3-12-research0082 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Li X., Hu Z., Yu X., Zhang C., Ma B., He L., Wei C. and Wu J. (2017). Dosage compensation in the process of inactivation/reactivation during both germ cell development and early embryogenesis in mouse. Sci. Rep. 7, 3729. 10.1038/s41598-017-03829-z [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Lin H., Gupta V., Vermilyea M. D., Falciani F., Lee J. T., O'Neill L. P. and Turner B. M. (2007). Dosage compensation in the mouse balances up-regulation and silencing of X-linked genes. PLoS Biol. 5, e326. 10.1371/journal.pbio.0050326 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Lin H., Halsall J. A., Antczak P., O'Neill L. P., Falciani F. and Turner B. M. (2011). Relative overexpression of X-linked genes in mouse embryonic stem cells is consistent with Ohno's hypothesis. Nat. Genet. 43, 1169-1170. 10.1038/ng.992 [Abstract] [CrossRef] [Google Scholar]
- Lin F., Xing K., Zhang J. and He X. (2012). Expression reduction in mammalian X chromosome evolution refutes Ohno's hypothesis of dosage compensation. Proc. Natl. Acad. Sci. U.S.A. 109, 11752-11757. 10.1073/pnas.1201816109 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Lister N. C., Milton A. M., Patel H. R., Waters S., Hanrahan B. J., McIntyre K. L., Livernois A. M., Wee L. K., Ringel A. R., Mundlos S.et al. (2023). Incomplete transcriptional dosage compensation of vertebrate sex chromosomes is balanced by post-transcriptional compensation. bioRxiv. [Europe PMC free article] [Abstract] [Google Scholar]
- Livernois A. M., Graves J. A. M. and Waters P. D. (2012). The origin and evolution of vertebrate sex chromosomes and dosage compensation. Heredity 108, 50-58. 10.1038/hdy.2011.106 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Li G., Figueiró H. V., Eizirik E. and Murphy W. J. (2019). Recombination-aware phylogenomics reveals the structured genomic landscape of hybridizing cat species. Mol. Biol. Evol. 36, 2111-2126. 10.1093/molbev/msz139 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Loda A., Collombet S. and Heard E. (2022). Gene regulation in time and space during X-chromosome inactivation. Nat. Rev. Mol. Cell Biol. 23, 231-249. 10.1038/s41580-021-00438-7 [Abstract] [CrossRef] [Google Scholar]
- Lorenzo J. L. R., Hubinsky M., Vyskot B. and Hobza R. (2020). Histone post-translational modifications in Silene latifolia X and Y chromosomes suggest a mammal-like dosage compensation system. Plant Science 229, e110528. 10.1016/j.plantsci.2020.110528 [Abstract] [CrossRef] [Google Scholar]
- Luchsinger-Morcelle S. J., Gribnau J. and Mira-Bontenbal H. (2024). Orchestrating asymmetric expression: mechanisms behind Xist regulation. Epigenomes 8, 6. 10.3390/epigenomes8010006 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Luo Z.-X., Yuan C.-X., Meng Q.-J. and Ji Q. (2011). A Jurassic eutherian mammal and divergence of marsupials and placentals. Nature 476, 442-445. 10.1038/nature10291 [Abstract] [CrossRef] [Google Scholar]
- Lyon, M. F. (1961). Gene action in the X-chromosome of the mouse (Mus musculus L.). Nature 190, 372-373. 10.1038/190372a0 [Abstract] [CrossRef] [Google Scholar]
- Lyu Q., Yang Q., Hao J., Yue Y., Wang X., Tian J. and An L. (2022). A small proportion of X-linked genes contribute to X chromosome upregulation in early embryos via BRD4-mediated transcriptional activation. Curr. Biol. 32, 4397-4410.e5. 10.1016/j.cub.2022.08.059 [Abstract] [CrossRef] [Google Scholar]
- Lyu Q., Yang Q., Tian J. and An L. (2023). Response to Lentini and Reinius. Curr. Biol. 33, R397. 10.1016/j.cub.2023.03.052 [Abstract] [CrossRef] [Google Scholar]
- Magaraki A., Loda A., Gontan C., Merzouk S., Sleddens-Linkels E., Meek S., Baarends W. M., Burdon T. and Gribnau J. (2019). A novel approach to differentiate rat embryonic stem cells in vitro reveals a role for RNF12 in activation of X chromosome inactivation. Sci. Rep. 9, 6068. 10.1038/s41598-019-42246-2 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Mahadevaiah S. K., Royo H., VandeBerg J. L., McCarrey J. R., Mackay S. and Turner J. M. A. (2009). Key features of the X inactivation process are conserved between marsupials and eutherians. Curr. Biol. 19, 1478-1484. 10.1016/j.cub.2009.07.041 [Abstract] [CrossRef] [Google Scholar]
- Mahadevaiah S. K., Sangrithi M. N., Hirota T. and Turner J. M. A. (2020). A single-cell transcriptome atlas of marsupial embryogenesis and X inactivation. Nature 586, 612-617. 10.1038/s41586-020-2629-6 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Mandal S., Chandel D., Kaur H., Majumdar S., Arava M. and Gayen S. (2020). Single-cell analysis reveals partial reactivation of X chromosome instead of chromosome-wide dampening in naive human pluripotent stem cells. Stem Cell Rep. 14, 745-754. 10.1016/j.stemcr.2020.03.027 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Mank J. E., Hosken D. J. and Wedell N. (2011). Some inconvenient truths about sex chromosome dosage compensation and the potential role of sexual conflict. Evolution 65, 2133-2144. 10.1111/j.1558-5646.2011.01316.x [Abstract] [CrossRef] [Google Scholar]
- Marahrens Y., Panning B., Dausman J., Strauss W. and Jaenisch R. (1997). Xist-deficient mice are defective in dosage compensation but not spermatogenesis. Genes Dev. 11, 156-166. 10.1101/gad.11.2.156 [Abstract] [CrossRef] [Google Scholar]
- Marin R., Cortez D., Lamanna F., Pradeepa M. M., Leushkin E., Julien P., Liechti A., Halbert J., Brüning T., Mössinger K.et al. (2017). Convergent origination of a Drosophila-like dosage compensation mechanism in a reptile lineage. Genome Res. 27, 1974-1987. 10.1101/gr.223727.117 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Martínez-Pacheco M., Tenorio M., Almonte L., Fajardo V., Godínez A., Fernández D., Cornejo-Páramo P., Díaz-Barba K., Halbert J.et al. (2020). Expression evolution of ancestral XY gametologs across all major groups of placental mammals. Genome Biol. Evol. 12, 2015-2028. 10.1093/gbe/evaa173 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Marshall Graves, J. A. (2008). Weird animal genomes and the evolution of vertebrate sex and sex chromosomes. Annu. Rev. Genet. 42, 565-586. 10.1146/annurev.genet.42.110807.091714 [Abstract] [CrossRef] [Google Scholar]
- Ma X., Yin J., Qiao L., Wan H., Liu X., Zhou Y., Wu J., Niu L., Wu M., Wang X.et al. (2024). A programmable targeted protein-degradation platform for versatile applications in mammalian cells and mice. Mol. Cell 84, 1585-1600.e7. 10.1016/j.molcel.2024.02.019 [Abstract] [CrossRef] [Google Scholar]
- McIntyre K. L., Waters S. A., Zhong L., Hart-Smith G., Raftery M., Chew Z. A., Patel H. R., Marshall Graves J. A. and Waters P. D. (2024). Identification of the RSX interactome in a marsupial shows functional coherence with the Xist interactome during X inactivation. Genome Biol. 25, 134. 10.1186/s13059-024-03280-0 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Messer M., Weiss A., Shaw D. and Westerman M. (1998). Evolution of the monotremes: phylogenetic relationship to marsupials and eutherians, and estimation of divergence dates based on α-Lactalbumin amino acid sequences semantic scholar. J. Mamm. Evol 5, 95-105. 10.1023/A:1020523120739 [CrossRef] [Google Scholar]
- Meunier J., Lemoine F., Soumillon M., Liechti A., Weier M., Guschanski K., Hu H., Khaitovich P. and Kaessmann H. (2013). Birth and expression evolution of mammalian microRNA genes. Genome Res. 23, 34-45. 10.1101/gr.140269.112 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Meyer, B. J. (2022). The X chromosome in C. elegans sex determination and dosage compensation. Curr. Opin. Genet. Dev. 74, 101912. 10.1016/j.gde.2022.101912 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Migeon, B. R. (2002). X chromosome inactivation: theme and variations. Cytogenet. Genome Res. 99, 8-16. 10.1159/000071568 [Abstract] [CrossRef] [Google Scholar]
- Mills J. A., Herrera P. S., Kaur M., Leo L., McEldrew D., Tintos-Hernandez J. A., Rajagopalan R., Gagne A., Zhang Z., Ortiz-Gonzalez X. R.et al. (2018). NIPBL+/- haploinsufficiency reveals a constellation of transcriptome disruptions in the pluripotent and cardiac states. Sci. Rep. 8, 1056. 10.1038/s41598-018-19173-9 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Moreira de Mello J. C., de Araújo E. S. S., Stabellini R., Fraga A. M., de Souza J. E. S., Sumita D. R., Camargo A. A. and Pereira L. V. (2010). Random X inactivation and extensive mosaicism in human placenta revealed by analysis of allele-specific gene expression along the X chromosome. PLoS ONE 5, e10947. 10.1371/journal.pone.0010947 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Moreira de Mello J. C., Fernandes G. R., Vibranovski M. D. and Pereira L. V. (2017). Early X chromosome inactivation during human preimplantation development revealed by single-cell RNA-sequencing. Sci. Rep. 7, 10794. 10.1038/s41598-017-11044-z [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Moyano Rodriguez Y. and Borensztein M. (2023). X-chromosome inactivation: a historic topic that's still hot. Development 150, dev202072. 10.1242/dev.202072 [Abstract] [CrossRef] [Google Scholar]
- Moyer A. J., Gardiner K. and Reeves R. H. (2021). All creatures great and small: new approaches for understanding down syndrome genetics. Trends Genet. 37, 444-459. 10.1016/j.tig.2020.09.017 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Mugford J. W., Yee D. and Magnuson T. (2012). Failure of extra-embryonic progenitor maintenance in the absence of dosage compensation. Development 139, 2130-2138. 10.1242/dev.076497 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Muller, H. J. (1914). A gene for the fourth chromosome of Drosophila. J. Exp. Zool. 17, 325-336. 10.1002/jez.1400170303 [CrossRef] [Google Scholar]
- Mulugeta E., Wassenaar E., Sleddens-Linkels E., van IJcken W. F. J., Heard E., Grootegoed J. A., Just W., Gribnau J. and Baarends W. M. (2016). Genomes of Ellobius species provide insight into the evolutionary dynamics of mammalian sex chromosomes. Genome Res. 26, 1202-1210. 10.1101/gr.201665.115 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Murphy W. J., Larkin D. M., Everts-van der Wind A., Bourque G., Tesler G., Auvil L., Beever J. E., Chowdhary B. P., Galibert F., Gatzke L.et al. (2005). Dynamics of mammalian chromosome evolution inferred from multispecies comparative maps. Science 309, 613-617. 10.1126/science.1111387 [Abstract] [CrossRef] [Google Scholar]
- Musio A., Selicorni A., Focarelli M. L., Gervasini C., Milani D., Russo S., Vezzoni P. and Larizza L. (2006). X-linked Cornelia de Lange syndrome owing to SMC1L1 mutations. Nat. Genet. 38, 528-530. 10.1038/ng1779 [Abstract] [CrossRef] [Google Scholar]
- Mutzel V. and Schulz E. G. (2020). Dosage sensing, threshold responses, and epigenetic memory: a systems biology perspective on random X-chromosome inactivation. BioEssays 42, e1900163. 10.1002/bies.201900163 [Abstract] [CrossRef] [Google Scholar]
- Muyle A., Marais G. A. B., Bačovský V., Hobza R. and Lenormand T. (2022). Dosage compensation evolution in plants: theories, controversies and mechanisms. Philos. Trans. R. Soc. Lond. B Biol. Sci. 377, 20210222. 10.1098/rstb.2021.0222 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Nadeau, J. H. (1989). Maps of linkage and synteny homologies between mouse and man. Trends Genet. 5, 82-86. 10.1016/0168-9525(89)90031-0 [Abstract] [CrossRef] [Google Scholar]
- Naik H. C., Hari K., Chandel D., Jolly M. K. and Gayen S. (2022). Single-cell analysis reveals X upregulation is not global in pre-gastrulation embryos. iScience 25, 104465. 10.1016/j.isci.2022.104465 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Nakamura T., Fujiwara K., Saitou M. and Tsukiyama T. (2021). Non-human primates as a model for human development. Stem Cell Reports 16, 1093-1103. 10.1016/j.stemcr.2021.03.021 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Naqvi S., Bellott D. W., Lin K. S. and Page D. C. (2018). Conserved microRNA targeting reveals preexisting gene dosage sensitivities that shaped amniote sex chromosome evolution. Genome Res. 28, 474-483. 10.1101/gr.230433.117 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Naqvi S., Kim S., Hoskens H., Matthews H. S., Spritz R. A., Klein O. D., Hallgrímsson B., Swigut T., Claes P., Pritchard J. K.et al. (2023). Precise modulation of transcription factor levels identifies features underlying dosage sensitivity. Nat. Genet. 55, 841-851. 10.1038/s41588-023-01366-2 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Navarro-Cobos M. J., Balaton B. P. and Brown C. J. (2020). Genes that escape from X-chromosome inactivation: Potential contributors to Klinefelter syndrome. Am. J. Med. Genet. C Semin. Med. Genet. 184, 226-238. 10.1002/ajmg.c.31800 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Nguyen D. K. and Disteche C. M. (2006). Dosage compensation of the active X chromosome in mammals. Nat. Genet. 38, 47-53. 10.1038/ng1705. [Abstract] [CrossRef] [Google Scholar]
- Nguyen T. A., Wu K., Pandey S., Lehr A. W., Li Y., Bemben M. A., Badger J. D., Lauzon J. L., Wang T., Zaghloul K. A.et al. (2020). A cluster of autism-associated variants on X-linked NLGN4X functionally resemble NLGN4Y. Neuron 106, 759-768. 10.1016/j.neuron.2020.03.008 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Noviello G., Gjaltema R. A. F. and Schulz E. G. (2023). CasTuner is a degron and CRISPR/Cas-based toolkit for analog tuning of endogenous gene expression. Nat. Commun. 14, 3225. 10.1038/s41467-023-38909-4 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Ogawa Y. and Lee J. T. (2003). Xite, X-inactivation intergenic transcription elements that regulate the probability of choice. Mol. Cell 11, 731-743. 10.1016/S1097-2765(03)00063-7 [Abstract] [CrossRef] [Google Scholar]
- Ohno, S. (1967). Sex Chromosomes and Sex-Linked Genes. Berlin, Heidelberg: Springer Berlin Heidelberg. [Google Scholar]
- Ohno S., Kaplan W. D. and Kinosita R. (1959). Formation of the sex chromatin by a single X-chromosome in liver cells of Rattus norvegicus. Exp. Cell Res. 18, 415-418. 10.1016/0014-4827(59)90031-X [Abstract] [CrossRef] [Google Scholar]
- Okamoto I. and Heard E. (2009). Lessons from comparative analysis of X-chromosome inactivation in mammals. Chromosome Res. 17, 659-669. 10.1007/s10577-009-9057-7 [Abstract] [CrossRef] [Google Scholar]
- Okamoto I., Patrat C., Thépot D., Peynot N., Fauque P., Daniel N., Diabangouaya P., Wolf J.-P., Renard J.-P., Duranthon V.et al. (2011). Eutherian mammals use diverse strategies to initiate X-chromosome inactivation during development. Nature 472, 370-374. 10.1038/nature09872 [Abstract] [CrossRef] [Google Scholar]
- Okamoto I., Nakamura T., Sasaki K., Yabuta Y., Iwatani C., Tsuchiya H., Nakamura S.-I., Ema M., Yamamoto T. and Saitou M. (2021). The X chromosome dosage compensation program during the development of cynomolgus monkeys. Science 374, eabd8887. 10.1126/science.abd8887 [Abstract] [CrossRef] [Google Scholar]
- Panda A., Zylicz J. J. and Pasque V. (2020). New insights into X-chromosome reactivation during reprogramming to pluripotency. Cells 9, 2706. 10.3390/cells9122706 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Park C., Carrel L. and Makova K. D. (2010). Strong purifying selection at genes escaping X chromosome inactivation. Mol. Biol. Evol. 27, 2446-2450. 10.1093/molbev/msq143 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Penny G. D., Kay G. F., Sheardown S. A., Rastan S. and Brockdorff N. (1996). Requirement for Xist in X chromosome inactivation. Nature 379, 131-137. 10.1038/379131a0 [Abstract] [CrossRef] [Google Scholar]
- Perez-Garcia V., Fineberg E., Wilson R., Murray A., Mazzeo C. I., Tudor C., Sienerth A., White J. K., Tuck E., Ryder E. J.et al. (2018). Placentation defects are highly prevalent in embryonic lethal mouse mutants. Nature 555, 463-468. 10.1038/nature26002 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Pessia E., Makino T., Bailly-Bechet M., McLysaght A. and Marais G. A. B. (2012). Mammalian X chromosome inactivation evolved as a dosage-compensation mechanism for dosage-sensitive genes on the X chromosome. Proc. Natl. Acad. Sci. U.S.A. 109, 5346-5351. 10.1073/pnas.1116763109 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Pessia E., Engelstädter J. and Marais G. A. B. (2014). The evolution of X chromosome inactivation in mammals: the demise of Ohno's hypothesis? Cell. Mol. Life Sci. 71, 1383-1394. 10.1007/s00018-013-1499-6 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Petropoulos S., Edsgärd D., Reinius B., Deng Q., Panula S. P., Codeluppi S., Plaza Reyes A., Linnarsson S., Sandberg R. and Lanner F. (2016). Single-cell RNA-Seq reveals lineage and X chromosome dynamics in human preimplantation embryos. Cell 165, 1012-1026. 10.1016/j.cell.2016.03.023 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Potrzebowski L., Vinckenbosch N., Marques A. C., Chalmel F., Jégou B. and Kaessmann H. (2008). Chromosomal gene movements reflect the recent origin and biology of therian sex chromosomes. PLoS Biol. 6, e80. 10.1371/journal.pbio.0060080 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Potrzebowski L., Vinckenbosch N. and Kaessmann H. (2010). The emergence of new genes on the young therian X. Trends Genet. 26, 1-4. 10.1016/j.tig.2009.11.001 [Abstract] [CrossRef] [Google Scholar]
- Qian W. and Zhang J. (2014). Genomic evidence for adaptation by gene duplication. Genome Res. 24, 1356-1362. 10.1101/gr.172098.114 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Ramos-Ibeas P., Sang F., Zhu Q., Tang W. W. C., Withey S., Klisch D., Wood L., Loose M., Surani M. A. and Alberio R. (2019). Pluripotency and X chromosome dynamics revealed in pig pre-gastrulating embryos by single cell analysis. Nat. Commun. 10, 500. 10.1038/s41467-019-08387-8 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Rens W., Grützner F., O'brien P. C. M., Fairclough H., Graves J. A. M. and Ferguson-Smith M. A. (2004). Resolution and evolution of the duck-billed platypus karyotype with an X1Y1X2Y2X3Y3X4Y4X5Y5 male sex chromosome constitution. Proc. Natl. Acad. Sci. U.S.A. 101, 16257-16261. 10.1073/pnas.0405702101 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Rice, W. R. (1996). Evolution of the Y sex chromosome in animals. Bioscience 46, 331-343. 10.2307/1312947 [CrossRef] [Google Scholar]
- Richart L., Picod-Chedotel M.-L., Wassef M., Macario M., Aflaki S., Salvador M. A., Héry T., Dauphin A., Wicinski J., Chevrier V.et al. (2022). XIST loss impairs mammary stem cell differentiation and increases tumorigenicity through Mediator hyperactivation. Cell 185, 2164-2183.e25. 10.1016/j.cell.2022.04.034 [Abstract] [CrossRef] [Google Scholar]
- Romagnano A., Richer C. L., King W. A. and Betteridge K. J. (1987). Analysis of X-chromosome inactivation in horse embryos. J. Reprod. Fertil. Suppl. 35, 353-361. [Abstract] [Google Scholar]
- Rosin L. F., Chen D., Chen Y. and Lei E. P. (2022). Dosage compensation in Bombyx mori is achieved by partial repression of both Z chromosomes in males. Proc. Natl. Acad. Sci. U.S.A. 119, e2113374119. 10.1073/pnas.2113374119 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Rosspopoff O., Cazottes E., Huret C., Loda A., Collier A. J., Casanova M., Rugg-Gunn P. J., Heard E., Ouimette J.-F. and Rougeulle C. (2023). Species-specific regulation of XIST by the JPX/FTX orthologs. Nucleic Acids Res. 51, 2177-2194. 10.1093/nar/gkad029 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Rücklé C., Körtel N., Basilicata M. F., Busch A., Zhou Y., Hoch-Kraft P., Tretow K., Kielisch F., Bertin M., Pradhan M.et al. (2023). RNA stability controlled by m6A methylation contributes to X-to-autosome dosage compensation in mammals. Nat. Struct. Mol. Biol. 30, 1207-1215. 10.1038/s41594-023-00997-7 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Sado T., Wang Z., Sasaki H. and Li E. (2001). Regulation of imprinted X-chromosome inactivation in mice by Tsix. Development 128, 1275-1286. 10.1242/dev.128.8.1275 [Abstract] [CrossRef] [Google Scholar]
- Saiba R., Arava M. and Gayen S. (2018). Dosage compensation in human pre-implantation embryos: X-chromosome inactivation or dampening? EMBO Rep. 19, e46294. 10.15252/embr.201846294 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- San Roman A. K., Godfrey A. K., Skaletsky H., Bellott D. W., Groff A. F., Harris H. L., Blanton L. V., Hughes J. F., Brown L., Phou S.et al. (2023). The human inactive X chromosome modulates expression of the active X chromosome. Cell Genomics 3, 100259. 10.1016/j.xgen.2023.100259 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- San Roman A. K., Skaletsky H., Godfrey A. K., Bokil N. V., Teits L., Singh I., Blanton L. V., Bellott D. W., Pyntikova T.et al. (2024). The human Y and inactive X chromosomes similarly modulate autosomal gene expression. Cell Genomics 4, 100462. 10.1016/j.xgen.2023.100462 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Schubert O. T., Röst H. L., Collins B. C., Rosenberger G. and Aebersold R. (2017). Quantitative proteomics: challenges and opportunities in basic and applied research. Nat. Protoc. 12, 1289-1294. 10.1038/nprot.2017.040 [Abstract] [CrossRef] [Google Scholar]
- Schwämmle T. and Schulz E. G. (2023). Regulatory principles and mechanisms governing the onset of random X-chromosome inactivation. Curr. Opin. Genet. Dev. 81, 102063. 10.1016/j.gde.2023.102063 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Seshadri R., Matthews C., Dobrovic A. and Horsfall D. J. (1989). The significance of oncogene amplification in primary breast cancer. Int. J. Cancer 43, 270-272. 10.1002/ijc.2910430218 [Abstract] [CrossRef] [Google Scholar]
- Shen H., Yanas A., Owens M. C., Zhang C., Fritsch C., Fare C. M., Copley K. E., Shorter J., Goldman Y. E. and Liu K. F. (2022). Sexually dimorphic RNA helicases DDX3X and DDX3Y differentially regulate RNA metabolism through phase separation. Mol. Cell 82, 2588-2603. 10.1016/j.molcel.2022.04.022 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Shevchenko A. I., Zakharova I. S., Elisaphenko E. A., Kolesnikov N. N., Whitehead S., Bird C., Ross M., Weidman J. R., Jirtle R. L., Karamysheva T. V.et al. (2007). Genes flanking Xist in mouse and human are separated on the X chromosome in American marsupials. Chromosome Res. 15, 127-136. 10.1007/s10577-006-1115-9 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Shevchenko A. I., Zakharova I. S. and Zakian S. M. (2013). The evolutionary pathway of x chromosome inactivation in mammals. Acta Naturae 5, 40-53. 10.32607/20758251-2013-5-2-40-53 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Shevchenko A. I., Dementyeva E. V., Zakharova I. S. and Zakian S. M. (2019). Diverse developmental strategies of X chromosome dosage compensation in eutherian mammals. Int. J. Dev. Biol. 63, 223-233. 10.1387/ijdb.180376as [Abstract] [CrossRef] [Google Scholar]
- Sinha A. U. and Meller J. (2007). Cinteny: flexible analysis and visualization of synteny and genome rearrangements in multiple organisms. BMC Bioinformatics 8, 82. 10.1186/1471-2105-8-82 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Souyris M., Cenac C., Azar P., Daviaud D., Canivet A., Grunenwald S., Pienkowski C., Chaumeil J., Mejía J. E. and Guéry J.-C. (2018). TLR7 escapes X chromosome inactivation in immune cells. Sci. Immunol. 3, eaap8855. 10.1126/sciimmunol.aap8855 [Abstract] [CrossRef] [Google Scholar]
- Spaziano A. and Cantone I. (2021). X-chromosome reactivation: a concise review. Biochem. Soc. Trans. 49, 2797-2805. 10.1042/BST20210777 [Abstract] [CrossRef] [Google Scholar]
- Stevanović M., Lovell-Badge R., Collignon J. and Goodfellow P. N. (1993). SOX3 is an X-linked gene related to SRY. Hum. Mol. Genet. 2, 2013-2018. 10.1093/hmg/2.12.2013 [Abstract] [CrossRef] [Google Scholar]
- Sun S., Del Rosario B. C., Szanto A., Ogawa Y., Jeon Y. and Lee J. T. (2013). Jpx RNA activates Xist by evicting CTCF. Cell 153, 1537-1551. 10.1016/j.cell.2013.05.028 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Sutton E., Hughes J., White S., Sekido R., Tan J., Arboleda V., Rogers N., Knower K., Rowley L., Eyre H.et al. (2011). Identification of SOX3 as an XX male sex reversal gene in mice and humans. J. Clin. Invest 121, 328-341. 10.1172/JCI42580 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Svoboda, P. (2018). Mammalian zygotic genome activation. Semin. Cell Dev. Biol. 84, 118-126. 10.1016/j.semcdb.2017.12.006 [Abstract] [CrossRef] [Google Scholar]
- Talon I., Janiszewski A., Chappell J., Vanheer L. and Pasque V. (2019). Recent advances in understanding the reversal of gene silencing during X chromosome reactivation. Front. Cell Dev. Biol. 7, 169. 10.3389/fcell.2019.00169 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Talon I., Janiszewski A., Theeuwes B., Lefevre T., Song J., Bervoets G., Vanheer L., De Geest N., Poovathingal S., Allsop R.et al. (2021). Enhanced chromatin accessibility contributes to X chromosome dosage compensation in mammals. Genome Biol. 22, 302. 10.1186/s13059-021-02518-5 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- The Gene Ontology Consortium (2021). The Gene Ontology resource: enriching a GOld mine. Nucleic Acids Res. 49, D325-D334. 10.1093/nar/gkaa1113 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Tian D., Sun S. and Lee J. T. (2010). The long noncoding RNA, Jpx, is a molecular switch for X chromosome inactivation. Cell 143, 390-403. 10.1016/j.cell.2010.09.049 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Tomihara K., Kawamoto M., Suzuki Y., Katsuma S. and Kiuchi T. (2022). Masculinizer-induced dosage compensation is achieved by transcriptional downregulation of both copies of Z-linked genes in the silkworm, Bombyx mori. Biol. Lett. 18, 20220116. 10.1098/rsbl.2022.0116 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Topa H., Benoit-Pilven C., Tukiainen T. and Pietiläinen O. (2024). X-chromosome inactivation in human iPSCs provides insight into X-regulated gene expression in autosomes. Genome Biol. 25, 144. 10.1186/s13059-024-03286-8 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Vallot C., Patrat C., Collier A. J., Huret C., Casanova M., Liyakat Ali T. M., Tosolini M., Frydman N., Heard E., Rugg-Gunn P. J.et al. (2017). XACT noncoding RNA competes with XIST in the control of X chromosome activity during human early development. Cell Stem Cell 20, 102-111. 10.1016/j.stem.2016.10.014 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Veitia R. A. and Birchler J. A. (2022). Gene-dosage issues: a recurrent theme in whole genome duplication events. Trends Genet. 38, 1-3. 10.1016/j.tig.2021.06.006 [Abstract] [CrossRef] [Google Scholar]
- Vicoso B. and Bachtrog D. (2009). Progress and prospects toward our understanding of the evolution of dosage compensation. Chromosome Res. 17, 585-602. 10.1007/s10577-009-9053-y [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Wake N., Takagi N. and Sasaki M. (1976). Non-random inactivation of X chromosome in the rat yolk sac. Nature 262, 580-581. 10.1038/262580a0 [Abstract] [CrossRef] [Google Scholar]
- Wang X., Miller D. C., Clark A. G. and Antczak D. F. (2012). Random X inactivation in the mule and horse placenta. Genome Res. 22, 1855-1863. 10.1101/gr.138487.112 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Wang F., Shin J., Shea J. M., Yu J., Bošković A., Byron M., Zhu X., Shalek A. K., Regev A., Lawrence J. B. et al. (2016). Regulation of X-linked gene expression during early mouse development by Rlim. eLife 5, e19127. 10.7554/eLife.19127 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Wang M., Lin F., Xing K. and Liu L. (2017). Random X-chromosome inactivation dynamics in vivo by single-cell RNA sequencing. BMC Genomics 18, 90. 10.1186/s12864-016-3466-8 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Wang Z. Y., Leushkin E., Liechti A., Ovchinnikova S., Mößinger K., Brüning T., Rummel C., Grützner F., Cardoso-Moreira M., Janich P. et al. (2020). Transcriptome and translatome co-evolution in mammals. Nature 588, 642-647. 10.1038/s41586-020-2899-z [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Warren W. C., Hillier L. W., Marshall Graves J. A., Birney E., Ponting C. P., Grützner F., Belov K., Miller W., Clarke L., Chinwalla A. T.et al. (2008). Genome analysis of the platypus reveals unique signatures of evolution. Nature 453, 175-183. 10.1038/nature06936 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Waters P. D., Delbridge M. L., Deakin J. E., El-Mogharbel N., Kirby P. J., Carvalho-Silva D. R. and Graves J. A. M. (2005). Autosomal location of genes from the conserved mammalian X in the platypus (Ornithorhynchus anatinus): implications for mammalian sex chromosome evolution. Chromosome Res. 13, 401-410. 10.1007/s10577-005-0978-5 [Abstract] [CrossRef] [Google Scholar]
- Wright A. E., Dean R., Zimmer F. and Mank J. E. (2016). How to make a sex chromosome. Nat. Commun. 7, 12087. 10.1038/ncomms12087 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Xiong Y., Chen X., Chen Z., Wang X., Shi S., Wang X., Zhang J. and He X. (2010). RNA sequencing shows no dosage compensation of the active X-chromosome. Nat. Genet. 42, 1043-1047. 10.1038/ng.711 [Abstract] [CrossRef] [Google Scholar]
- Xue F., Tian X. C., Du F., Kubota C., Taneja M., Dinnyes A., Dai Y., Levine H., Pereira L. V. and Yang X. (2002). Aberrant patterns of X chromosome inactivation in bovine clones. Nat. Genet. 31, 216-220. 10.1038/ng900 [Abstract] [CrossRef] [Google Scholar]
- Yang J.-R. and Chen X. (2019). Dosage sensitivity of X-linked genes in human embryonic single cells. BMC Genomics 20, 42. 10.1186/s12864-019-5432-8 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Yildirim E., Sadreyev R. I., Pinter S. F. and Lee J. T. (2011). X-chromosome hyperactivation in mammals via nonlinear relationships between chromatin states and transcription. Nat. Struct. Mol. Biol. 19, 56-61. 10.1038/nsmb.2195 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Yoshido A. and Marec F. (2023). Deviations in the Z:A ratio disrupt sexual development in the eri silkmoth, Samia cynthia ricini. Genetics 224, iyad023. 10.1093/genetics/iyad023 [Abstract] [CrossRef] [Google Scholar]
- Youness A., Miquel C.-H. and Guéry J.-C. (2021). Escape from X Chromosome inactivation and the female predominance in autoimmune diseases. Int. J. Mol. Sci. 22, 1114. 10.3390/ijms22031114 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Yu B., van Tol H. T. A., Stout T. A. E. and Roelen B. A. J. (2020). Initiation of X Chromosome inactivation during bovine embryo development. Cells 9, 1016. [Europe PMC free article] [Abstract] [Google Scholar]
- Zhai J., Xiao Z., Wang Y. and Wang H. (2022). Human embryonic development: from peri-implantation to gastrulation. Trends Cell Biol. 32, 18-29. 10.1016/j.tcb.2021.07.008 [Abstract] [CrossRef] [Google Scholar]
- Zhang S., Wang R., Zhang L., Birchler J. A. and Sun L. (2024). Inverse and proportional trans modulation of gene expression in human aneuploidies. Genes 15, 637. 10.3390/genes15050637 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Zhou Y., Shearwin-Whyatt L., Li J., Song Z., Hayakawa T., Stevens D., Fenelon J. C., Peel E., Cheng Y., Pajpach F.et al. (2021). Platypus and echidna genomes reveal mammalian biology and evolution. Nature 592, 756-762. 10.1038/s41586-020-03039-0 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Zou H., Yu D., Du X., Wang J., Chen L., Wang Y., Xu H., Zhao Y., Zhao S., Pang Y.et al. (2019). No imprinted XIST expression in pigs: biallelic XIST expression in early embryos and random X inactivation in placentas. Cell. Mol. Life Sci. 76, 4525-4538. 10.1007/s00018-019-03123-3 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
Articles from Development (Cambridge, England) are provided here courtesy of Company of Biologists
Citations & impact
Impact metrics
Alternative metrics
Discover the attention surrounding your research
https://www.altmetric.com/details/166346883
Article citations
Sex Chromosome Evolution: Hallmarks and Question Marks.
Mol Biol Evol, 41(11):msae218, 01 Nov 2024
Cited by: 0 articles | PMID: 39417444 | PMCID: PMC11542634
Review Free full text in Europe PMC
Sex as a biological variable in ageing: insights and perspectives on the molecular and cellular hallmarks.
Open Biol, 14(10):240177, 30 Oct 2024
Cited by: 0 articles | PMID: 39471841 | PMCID: PMC11521605
Review Free full text in Europe PMC
Similar Articles
To arrive at the top five similar articles we use a word-weighted algorithm to compare words from the Title and Abstract of each citation.
Mammalian X Chromosome Dosage Compensation: Perspectives From the Germ Line.
Bioessays, 40(6):e1800024, 14 May 2018
Cited by: 13 articles | PMID: 29756331
Review
Role and control of X chromosome dosage in mammalian development.
Curr Opin Genet Dev, 23(2):109-115, 04 Mar 2013
Cited by: 57 articles | PMID: 23465885
Review
X chromosome dosage compensation: how mammals keep the balance.
Annu Rev Genet, 42:733-772, 01 Jan 2008
Cited by: 327 articles | PMID: 18729722
Review
Dosage compensation of the active X chromosome in mammals.
Nat Genet, 38(1):47-53, 11 Dec 2005
Cited by: 335 articles | PMID: 16341221
Funding
Funders who supported this work.
Centre National de la Recherche Scientifique
Fondation pour la Recherche Médicale (1)
Grant ID: AJE202305017142
Ligue Contre le Cancer (1)
Grant ID: LNCC_TAJT25850