X-inactivation in female mammals is triggered by the association of non-coding Xist RNA in cis with the X chromosome. Although it has been suggested that the A-repeat located in the proximal part of the Xist RNA is required for chromosomal silencing in ES cells, its role in mouse has not yet been addressed. Here, we deleted the A-repeat in mouse and studied its effects on X-inactivation during embryogenesis. The deletion,when paternally transmitted, caused a failure of imprinted X-inactivation in the extraembryonic tissues, demonstrating the essential role of the A-repeat in X-inactivation in the mouse embryo. Unexpectedly, the failure of X-inactivation was caused by a lack of Xist RNA rather than by a defect in the silencing function of the mutated RNA, which we expected to be expressed from the mutated X. Interestingly, the normally silent paternal copy of Tsix, which is an antisense negative regulator of Xist,was ectopically activated in the preimplantation embryo. Furthermore, CpG sites in the promoter region of paternal Xist, which are essentially unmethylated in the extraembryonic tissues of the wild-type female embryo,acquire a significant level of methylation on the mutated paternal X. These findings demonstrate that the DNA sequence deleted on the mutated X, most probably the A-repeat, is essential as a genomic element for the appropriate transcriptional regulation of the Xist/Tsix loci and subsequent X-inactivation in the mouse embryo.

During the early development of female mammals, one of the two X chromosomes undergoes transcriptional silencing along almost the entire region of the chromosome to achieve dosage equivalence with males of the X-linked genes, which females have twice as many of as males do (X-inactivation)(Lyon, 1961). Several lines of evidence indicate that the paternal X chromosome is preferentially inactivated in the early preimplantation stage embryo(Huynh and Lee, 2003; Mak et al., 2004; Okamoto et al., 2004). This imprinted X-inactivation is maintained throughout development of the extraembryonic tissues, such as the placenta and part of the extraembryonic membranes (Takagi and Sasaki,1975), which originate from the trophectoderm and primitive endoderm of the blastocyst. In the epiblast lineage, a derivative of the inner cell mass of the blastocyst, the previously inactivated paternal X transiently restores transcriptional activity, and subsequently either the maternal or paternal X undergoes inactivation in a basically random fashion as cells differentiate (random X-inactivation).

It is known that the Xist (X-inactive specific transcript) gene located in the X inactivation center (Xic), a cytogenetically identified chromosomal region essential for X-inactivation to occur in cis, plays a crucial role in both imprinted and random X-inactivation. Targeted disruption of Xist renders the mutated X incompetent to undergo inactivation(Marahrens et al., 1997; Penny et al., 1996), and therefore paternal transmission of Xist deficiency results in the failure of imprinted paternal X-inactivation in the extraembryonic tissues and, consequently, a selective loss of female embryos soon after implantation owing to the extremely poor development of the extraembryonic tissues.

Xist encodes long non-coding transcripts as long as 17 kb in length, which are subjected to splicing and polyadenylation like common protein-coding RNAs (Brockdorff et al.,1992; Brown et al.,1992). The Xist RNA is peculiar in that it stays in the nucleus and associates with the X chromosome, from which it is transcribed(Brown et al., 1992; Clemson et al., 1996), where it eventually induces chromosomal silencing by unknown mechanisms. Although the overall structure of the Xist gene is relatively conserved among eutherian mammals, its nucleotide sequence diverges greatly(Chureau et al., 2002; Nesterova et al., 2001), which is consistent with the presumed role of the Xist gene product as a functional RNA. It is known, however, that several regions consisting of a series of repeats are conserved between mouse and human(Brockdorff et al., 1992; Brown et al., 1992). Wutz et al. previously showed that one of these repeats, known as the A-repeat, which is located in the proximal part of the Xist RNA, is crucial for the silencing function of the RNA (Wutz et al., 2002). They demonstrated that the RNA transcribed from a single copy Xist cDNA lacking the A-repeat driven by an inducible promoter can accumulate on the X chromosome in male ES cells upon induction but fails to initiate chromosomal silencing. The A-repeat contains 7.5 copies of a conserved direct repeat unit, which harbors two short inverted repeats that might fold into a secondary structure comprising two stem loops. These findings imply that the stem loop structures might be the modules that interact with the putative protein factors that are involved in chromosome silencing.

Although the above inducible expression assay in ES cells identified for the first time a likely functional domain in Xist RNA responsible for chromosomal silencing, its significance for X-inactivation taking place in the developing embryo has not yet been addressed. In this study, we introduced a mutant Xist allele lacking the A-repeat into the mouse and examined its effects on X-inactivation in the embryo. Our results clearly demonstrate that deletion of the A-repeat rendered the mutated X incompetent to undergo inactivation in embryos, which was consistent with the previous ES cell assay. However, the incompetence of the mutated X to undergo inactivation was apparently due to the lack of Xist expression on the mutated X chromosome. This finding suggests an unexpected essential role of the A-repeat as a genomic element for the appropriate regulation of Xist and subsequent X-inactivation in the mouse embryo.

Targeted deletion of the A-repeat

Targeting vector pXBΔA was constructed so that a floxed HSV-tk and PGK-neo fragment from pflox (a gift from En Li. location?) was flanked by genomic fragments derived from Bac clone 333J22 containing the Xistgene as shown in Fig. 1A (the 5′ arm, nucleotides 97, 704-106, 412 in AJ421479; the 3′ arm,nucleotides 107, 228-113, 284 in AJ421479). J1 ES cells(Li et al., 1992) were electroporated with pXBΔA as previously described(Sado et al., 2005), and selection was started 24 hours later in the presence of 250 μg/ml G418. Of 254 selected colonies, four harbored the expected homologous recombination(XistΔA2lox). Chimeric males were generated and crossed with females heterozygous for a Tsixdeficiency (Sado et al., 2001)to facilitate germ-line transmission in the same manner as previously described (Sado et al., 2005). Females heterozygous for XistΔA2lox were crossed with CAG-cre transgenic males to derive pups carrying XistΔA. Excision of the selection marker in ES cells was performed by transient expression of Cre recombinase using pBS185 (Life Technology).

Histology

Deciduas dissected out from the uterus were fixed in Bouin's fixative. Following dehydration, deciduas were embedded in Technovit 7100 (Kulzer),sectioned at 2 μm, and stained with Hematoxylin and Eosin.

Genotyping blastocysts

Blastocysts were flushed from the uterus at E3.5 and the zona pellucida was removed by acid tyroid treatment. Each blastocyst was transferred to 10 μl of water and heated for 3 minutes at 95°C. Five microliters of this solution was used for two-round PCR with semi-nested primer sets for genotyping and sexing. The primers used for the first round amplification were R700P2 (wild-type-specific), dA1F(XistΔA-specific), Xist1395R, Zfy1 and Zfy2. Subsequently, the wild-type Xist,XistΔA and Zfy sequences were individually amplified in a second round PCR using R700P2/F1063AS for wild-type Xist, dA1F/F1063AS for XistΔA and Zfy1/Zfy4 for Zfy. Primer sequences used in this study are shown in Table 1.Total RNA was extracted from the remaining 5 μl using Trizol (Invitrogen) in the presence of 10μg of E. coli tRNA.

Table 1.

Primer sequences used for PCR

PrimerSequence (5′ → 3′)
dA1F GCAGGTCGAGGGACCTAATA 
F1063AS GCACAACCCCGCAAATGCTA 
G6pdF4 CGCCATTTTGTCCTATGCTG 
G6pdR4 AGGTTGACCATGGAGTATGG 
GapdF ATGGCCTTCCGTGTTCCTAC 
GapdR TGTGAGGGAGATGCTCAGTG 
GapdR2 ATAGGGCCTCTCTTGCTCAG 
HprtF4 GTTGAATCTGCAAATACGAGG 
HprtR3 CAACGATTTACTGAAAGTGGG 
P422R TCAAGATGCGTGGATATCTCGG 
R700P2 CGGGGCTTGGTGGATGGAAAT 
Tsix2F CAATCTCGCAAGATCCGGTGA 
Tsix2R GATGCCAACGACACGTCTGA 
Tsix4F TGGGTCATTGGCATCTTAGTC 
Tsix4R TCAGCGTGAATCAACGAGAC 
Tsix4R2 CCCAGGGTGTCTGATCTCTT 
Xist1395R AAGCTGACATGTGACACACAAA 
Xist2281F GATGCCAACGACACGTCTGA 
Xist2424R AAGGACTCCAAAGTAACAATTCA 
Xist2688R AGAGCATTACAATTCAAGGCTC 
XistEx7F31 GGCATAGTCTCTACAAAATTTTCATT 
XistEx7R20 CTCCAATTTCTGGGCTCAAG 
Zfy1 GATAAGCTTACATAATCACATGGA 
Zfy2 CCTATGAAATCCTTTGCTGC 
Zfy4 TGCTTTTTGAGTGCTGATGG 
PrimerSequence (5′ → 3′)
dA1F GCAGGTCGAGGGACCTAATA 
F1063AS GCACAACCCCGCAAATGCTA 
G6pdF4 CGCCATTTTGTCCTATGCTG 
G6pdR4 AGGTTGACCATGGAGTATGG 
GapdF ATGGCCTTCCGTGTTCCTAC 
GapdR TGTGAGGGAGATGCTCAGTG 
GapdR2 ATAGGGCCTCTCTTGCTCAG 
HprtF4 GTTGAATCTGCAAATACGAGG 
HprtR3 CAACGATTTACTGAAAGTGGG 
P422R TCAAGATGCGTGGATATCTCGG 
R700P2 CGGGGCTTGGTGGATGGAAAT 
Tsix2F CAATCTCGCAAGATCCGGTGA 
Tsix2R GATGCCAACGACACGTCTGA 
Tsix4F TGGGTCATTGGCATCTTAGTC 
Tsix4R TCAGCGTGAATCAACGAGAC 
Tsix4R2 CCCAGGGTGTCTGATCTCTT 
Xist1395R AAGCTGACATGTGACACACAAA 
Xist2281F GATGCCAACGACACGTCTGA 
Xist2424R AAGGACTCCAAAGTAACAATTCA 
Xist2688R AGAGCATTACAATTCAAGGCTC 
XistEx7F31 GGCATAGTCTCTACAAAATTTTCATT 
XistEx7R20 CTCCAATTTCTGGGCTCAAG 
Zfy1 GATAAGCTTACATAATCACATGGA 
Zfy2 CCTATGAAATCCTTTGCTGC 
Zfy4 TGCTTTTTGAGTGCTGATGG 

RT-PCR

For quantitative RT-PCR analysis of undifferentiated ES cells, cDNA was prepared at 60°C using Thermoscript (Invitrogen) with Xist 2688R, Tsix2R and GapdR as primers. Real-time PCR was carried out as previously described (Sado et al.,2006), using Xist2281F and Xist 2424R as primers for Xist, Tsix2F and P422R for Tsix, and GapdF and GapdR2 for Gapd. The expression levels of Xist and Tsixwere normalized to Gapd levels as previously described(Sado et al., 2006).

For allelic expression analysis of X-linked genes in the trophoblast and in E7.5 embryos, cDNA was synthesized from 1 μg of total RNA using an oligo-dT primer, and PCR was carried out using G6pdF4 and G6pdR4 as primers for G6pd, and HprtF4 and HprtR3 for Hprt. The amplified products of G6pd and Hprt were subsequently digested with DraI and HinfI, respectively(Sugimoto and Abe, 2007).

For real-time PCR of individual F1 blastocysts isolated from XJF1XJF1 females crossed with either XΔAY or wild-type XY males, each embryo was lysed in 10 μl of distilled water by heating for 3 minutes at 95°C, and RNA was prepared using Trizol (Invitrogen) in the presence of 10 μg of E. colitRNA. Following DNase I treatment, cDNA was synthesized using an oligo-dT primer using the whole RNA sample in a reaction of 20 μl. One microliter of cDNA was subjected to real-time PCR amplification of a region containing a SacI polymorphism between JF1 and the laboratory strain in exon 7 of Xist with XistEx7F31 and XistEx7R20. Because Xist and Tsix are reciprocally imprinted in the blastocyst, SacI digestion of the product allowed the confirmation of specific amplification of Xist and not Tsix.

For allelic expression analysis of Tsix in individual genotyped F1 blastocysts, two-round PCR was carried out using cDNA synthesized in a strand-specific manner with Tsix4R and GapdR. The first round PCR was carried out on whole cDNA produced using Tsix4F, Tsix4R, GapdF and GapdR (25 cycles). The second round PCR was carried out on one-twentieth of the first round reaction using Tsix4F and Tsix4R2 (33 cycles), or GapdF and GapdR2 (25 cycles). The amplified product of Tsix was subsequently digested with BsmAI.

RNA-FISH

An RNA probe was prepared by in vitro transcription with Cy3-UTP (Amersham Pharmacia) and a plasmid (pBE1.5) containing a Xist cDNA fragment(nucleotides 9829-11,335 in NR_001463 in GenBank) using T7 RNA polymerase. Cytological preparations of blastocysts were made according to Okamoto et al.(Okamoto et al., 2000).

Following examination of RNA-FISH, X- and Y-chromosome painting was carried out according to the manufacturer's instructions (Cambio) to determine the sex chromosome constitution.

RNA half-life assay

About 2.5×104 undifferentiated ES cells were seeded onto each dish without feeder cells. The medium of each dish was replaced with that containing 25 μg/ml of DRB (5,6-Dichloro-1-β-D-ribofuranosyl benzimidazole, Calbiochem) on the following day and cells were collected every three hours. One microgram of total RNA isolated at each time point was converted into cDNA using an oligo-dT primer, and the level of Xistand Tsix RNA was quantitated by real-time PCR using the primer sets XistEx7F31/XistEx7R20 and Tsix4F/Tsix4R2, respectively. While XistExF31/XistEx7R20 are located about 20-kb downstream from the 5′ end of Xist, Tsix4F/Tsix4R2 are located upstream of the transcription start site of Xist. It was confirmed that there was no TsixcDNA in the reaction, which was long enough to serve as a template for XistExF31/XistEx7R20. The abundance of the respective RNAs at each time point was shown as the value relative to the abundance at 3 hours after the addition of DRB.

Bisulfite sequencing

Bisulfite treatment of genomic DNA prepared from the trophoblast at E6.5 and sperm was carried out using a Bisulfast kit (TOYOBO), and DNA was purified using an EZ kit (Zymo Research). Two-round PCR was carried out using semi-nested primers. First round PCR was carried out with 25 cycles on whole DNA prepared from the trophoblast or 300 ng of sperm DNA, and the second round PCR was done with 30 cycles on one-twentieth of the reaction from the first round PCR. Primer sequences will be provided upon request.

The mutated Xist allele lacking the A-repeat was introduced into the mouse

To study the importance of the A-repeat in X-inactivation taking place in the mouse embryo, we created a new mutant allele of Xist lacking the A-repeat by a gene targeting strategy (Fig. 1). Following verification by Southern blot analysis, ES cells harboring the expected homologous recombination were serially injected into blastocysts to generate chimeras. We postulated that the presence of a floxed selection marker in the targeted allele(XistΔA2lox) would functionally disrupt the Xist gene, and that the mutated allele, when paternally inherited, would result in a selective loss of female embryos soon after implantation because of the failure of imprinted X-inactivation in the extraembryonic lineages. We previously demonstrated, however, that this female-specific lethality is sometimes rescued by the simultaneous presence of a Tsix deficiency on the maternal X(Ohhata et al., 2008; Sado et al., 2005; Sado et al., 2006). Accordingly, the male chimeras were crossed with females heterozygous for the Tsix deficient allele (ΔTsix)(Sado et al., 2001) to facilitate transmission of the XistΔA2lox allele from fathers to live female pups, as previously described. Consequently, females carrying the ΔTsix allele and XistΔA2lox allele on the maternal and paternal X, respectively, were successfully recovered. They were subsequently crossed with males expressing Cre recombinase ubiquitously to derive XistΔA/+ females and XistΔA/Y males. These animals were apparently normal and fertile.

Fig. 1.

Targeted deletion of the A-repeat in the Xist gene.(A) Scheme for generating the XistΔA allele. Genomic structure of the Xist gene and the targeting vector are shown. The A-repeat region is shown in gray. The XistΔA allele lacks an 812 bp fragment containing the A-repeat (nucleotides 101-912 in GenBank L04961). Positions of the probes used for Southern blotting in B and C are also indicated. B, BamHI; P, PstI; R, EcoRI; Xb, XbaI; Xh, XhoI. (B) Homologous recombination was confirmed by Southern blotting. ΔA2lox211 is one of the four ES lines harboring the correct targeting event. Genomic DNA digested with XbaI(left) and PstI (right) was probed with HS0.7(Sado et al., 2005) and XXh0.7(Sado et al., 2006),respectively. (C) The presence of the respective mutation in the mouse was confirmed by Southern blotting. Tail DNA digested with XbaI was probed with BE0.6 (Sado et al.,2005).

Fig. 1.

Targeted deletion of the A-repeat in the Xist gene.(A) Scheme for generating the XistΔA allele. Genomic structure of the Xist gene and the targeting vector are shown. The A-repeat region is shown in gray. The XistΔA allele lacks an 812 bp fragment containing the A-repeat (nucleotides 101-912 in GenBank L04961). Positions of the probes used for Southern blotting in B and C are also indicated. B, BamHI; P, PstI; R, EcoRI; Xb, XbaI; Xh, XhoI. (B) Homologous recombination was confirmed by Southern blotting. ΔA2lox211 is one of the four ES lines harboring the correct targeting event. Genomic DNA digested with XbaI(left) and PstI (right) was probed with HS0.7(Sado et al., 2005) and XXh0.7(Sado et al., 2006),respectively. (C) The presence of the respective mutation in the mouse was confirmed by Southern blotting. Tail DNA digested with XbaI was probed with BE0.6 (Sado et al.,2005).

Basal transcription of Xist was not affected by deletion of the A-repeat in undifferentiated male ES cells

The targeting strategy was designed to delete the A-repeat without disrupting the endogenous Xist promoter. To confirm that the promoter remained functional in the XistΔA allele, we took advantage of XistΔA/Y ES cells established by transiently expressing Cre recombinase in XistΔA2lox/Y ES cells (data not shown). It is known that the Xist locus is transcribed at a very low level in undifferentiated male ES cells, although this basal transcription is eventually downregulated after differentiation. We examined the basal transcription of Xist to determine whether it was affected by the deletion of the A-repeat in undifferentiated XistΔA/Y ES cells. Quantitative RT-PCR with strand-specifically prepared cDNA revealed that, in XistΔA/Y ES cells, the mutated XistΔA RNA was expressed at a level comparable to wild-type Xist RNA in the parental male ES cells(Fig. 2A). In addition, the XistΔA allele was downregulated in the same manner as the wild-type allele after the induction of differentiation (Fig. 2A). These results demonstrated that neither the function of the Xistpromoter per se nor the mechanism for downregulating Xist on the future active X was affected by the deletion.

We further analyzed the stability of XistΔA RNA by treating cells with DRB, an inhibitor of RNA polymerase II. Real-time PCR on cDNA prepared from a series of DRB-treated cells demonstrated that there was no significant difference in stability between wild-type and mutant Xist RNA(Fig. 2B), indicating that the deletion of the A-repeat did not impair the stability of the RNA.

Intriguingly, the expression level of Tsix was increased in the mutant male ES cells. The stability of Tsix RNA in the mutant ES cells was, however, comparable to that in wild-type ES cells(Fig. 2B). It seemed likely therefore that the higher expression of Tsix in the mutant was mediated not by an increased stability of the RNA(Fig. 2A), but by a higher level of transcription. The expression level of Tsix, however,declined once ES cells were induced to differentiate. This suggests that although the genetic alterations we introduced at the Xist locus somehow facilitated the transcription of Tsix on the mutated X, they did not affect the mechanism for downregulating Tsix upon differentiation.

Paternal transmission of XistΔA results in a selective loss of female embryos

The functional significance of the A-repeat in embryonic development was first addressed by examining whether or not the mutated XistΔA allele could be transmitted to female pups from the father. Of 218 pups born to wild-type females crossed with XistΔA/Y males, 216 were male and 2 were female. One of the two females turned out to be XO, where the X chromosome was maternal in origin, and the other female inherited the XistΔA allele(Fig. 3A), suggesting that most female embryos had been lost in utero. In reciprocal crosses, the XistΔA allele was transmitted to apparently healthy female pups from the mothers at the expected ratio(Fig. 3A). Thus, the selective loss of females upon paternal transmission of the mutation was probably due to defects in the imprinted X-inactivation in the extraembryonic lineages. When embryos were dissected out at embryonic day (E) 6.5,+/XistΔA (the maternal allele precedes the paternal one by convention) female embryos, although found in a reasonable number, were all stunted with an abnormal morphology that was indistinguishable from that of female embryos carrying a dysfunctional Xist allele derived from the father(Fig. 3B). Histological analysis revealed that the extraembryonic ectoderm was severely affected in the morphologically abnormal embryos, which were most probably females carrying the paternal XistΔAallele (Fig. 3C). These embryos were reminiscent of those carrying an extra-copy of the maternal X(Goto and Takagi, 1998; Tada et al., 1993) or the paternally derived Xist-deficient X(Marahrens et al., 1997). These results strongly suggest that the deletion of the A-repeat severely compromised the function of the Xist gene to initiate X-inactivation.

Fig. 2.

Expression of Xist and Tsix in XistΔA/Y ES cells.(A) The stable level of Xist and Tsix RNA was quantitated by real-time RT-PCR. Xist expression in XistΔA/Y ES cells was comparable to that in the parental wild-type male ES cells before and after differentiation (upper panel). There was a striking difference in the expression level of Tsix in XistΔA/Y and wild-type XY ES in the undifferentiated state (lower panel). The former was about 5-fold higher than the latter, although both were downregulated upon differentiation.(B) The stability of Xist and Tsix RNA was analyzed by treating undifferentiated ES cells with DRB. There was no significant difference in the stability of either RNA between wild-type and XistΔA/Y ES cells.

Fig. 2.

Expression of Xist and Tsix in XistΔA/Y ES cells.(A) The stable level of Xist and Tsix RNA was quantitated by real-time RT-PCR. Xist expression in XistΔA/Y ES cells was comparable to that in the parental wild-type male ES cells before and after differentiation (upper panel). There was a striking difference in the expression level of Tsix in XistΔA/Y and wild-type XY ES in the undifferentiated state (lower panel). The former was about 5-fold higher than the latter, although both were downregulated upon differentiation.(B) The stability of Xist and Tsix RNA was analyzed by treating undifferentiated ES cells with DRB. There was no significant difference in the stability of either RNA between wild-type and XistΔA/Y ES cells.

Fig. 3.

XistΔA fails to inactivate the mutated X chromosome. (A) The paternal transmission of XistΔA results in an extreme bias in the sex ratio of live pups born to wild-type females crossed with XΔAY males in favor of males (upper panel). One of the two females turned out to be XO; therefore, the mutated allele was transmitted to only one female (0.46%). By contrast, the mutated allele was transmitted to both male and female pups at the expected ratio from the mothers (lower panel). (B) Gross morphology of an embryo typical of those that inherited the paternal XΔA (XXΔA). Scale bar: 200 μm. (C) Histological sections of the presumptive XXΔA embryos are shown in comparison with a presumptive male litter mate at E6.5. epc, ectoplacental cone; exe, extraembryonic ectoderm;ee, embryonic ectoderm (D) RT-PCR analysis of allelic expression of X-linked G6pd and Hprt in the trophoblast recovered from E6.5 embryos. XJF1 and XΔA/Xlab are maternal and paternal in origin, respectively. Expression of the paternal copy was evident in XJF1XΔA in both cases. (E)Expression of the GFP transgene on the paternal X in female embryos at E7.5. In contrast to wild-type female embryos (XXGFP), GFP fluorescence was uniformly negative in the embryonic tissues in XΔAXGFP embryos, suggesting that XΔA failed to undergo inactivation even in the embryonic tissues. (F) RT-PCR analysis of the expression of X-linked G6pd and Hprt in the embryonic tissues at E7.5. The maternal copies of both genes on XJF1 were not expressed in XΔAXJF1 embryos, suggesting that XΔA failed to undergo inactivation.

Fig. 3.

XistΔA fails to inactivate the mutated X chromosome. (A) The paternal transmission of XistΔA results in an extreme bias in the sex ratio of live pups born to wild-type females crossed with XΔAY males in favor of males (upper panel). One of the two females turned out to be XO; therefore, the mutated allele was transmitted to only one female (0.46%). By contrast, the mutated allele was transmitted to both male and female pups at the expected ratio from the mothers (lower panel). (B) Gross morphology of an embryo typical of those that inherited the paternal XΔA (XXΔA). Scale bar: 200 μm. (C) Histological sections of the presumptive XXΔA embryos are shown in comparison with a presumptive male litter mate at E6.5. epc, ectoplacental cone; exe, extraembryonic ectoderm;ee, embryonic ectoderm (D) RT-PCR analysis of allelic expression of X-linked G6pd and Hprt in the trophoblast recovered from E6.5 embryos. XJF1 and XΔA/Xlab are maternal and paternal in origin, respectively. Expression of the paternal copy was evident in XJF1XΔA in both cases. (E)Expression of the GFP transgene on the paternal X in female embryos at E7.5. In contrast to wild-type female embryos (XXGFP), GFP fluorescence was uniformly negative in the embryonic tissues in XΔAXGFP embryos, suggesting that XΔA failed to undergo inactivation even in the embryonic tissues. (F) RT-PCR analysis of the expression of X-linked G6pd and Hprt in the embryonic tissues at E7.5. The maternal copies of both genes on XJF1 were not expressed in XΔAXJF1 embryos, suggesting that XΔA failed to undergo inactivation.

Genes on the mutated X are not repressed in either the embryonic or extraembryonic tissues

We went on to study the transcriptional status of the mutated paternal X(XΔA) in the extraembryonic tissues. X-inactivation is imprinted in the trophoblast, a derivative of the trophectoderm in the blastocyst, in favor of the paternal X. The expression of X-linked genes was analyzed using trophoblasts isolated from E6.5 embryos. Embryos were prepared by crossing females carrying an X chromosome derived from JF1 (Mus m. molossinus) with XΔAY males, so that the parental origin of the X-linked gene transcripts could be addressed by the presence or absence of restriction site polymorphisms between JF1 and the laboratory strains used in this study (C57Bl/6 and 129). The regions harboring a polymorphism in the X-linked G6pd and Hprt genes(Sugimoto and Abe, 2007) were amplified by RT-PCR and the products were subsequently digested with DraI and HinfI, respectively. In the trophoblast of wild-type female embryos (XJF1Xlab), the expression of G6pd and Hprt was confined to the maternal alleles,consistent with the imprinted paternal X-inactivation in this tissue(Fig. 3D). By contrast,expression of the paternal copy of these genes was evident in XJF1XΔA females(Fig. 3D). This result demonstrated that genes on the paternally derived XΔA were,at least in part, misexpressed in the trophoblast. It seemed likely therefore that the paternal XΔA failed to undergo inactivation in the extraembryonic lineages, where the paternal X is programmed to be inactivated.

Fig. 4.

Expression of XistΔA RNA in preimplantation embryos. RNA-FISH was carried out in XXΔAand XX embryos at the eight-cell and blastocyst stages. At the eight-cell stage, although the accumulation of Xist RNA was evident in XX embryos, expression of XistΔARNA in XXΔA was very faint (arrowheads) and subsequently became undetectable by the blastocyst stage. Inset in the panel of the 8-cell XXΔA embryo is an enlargement of the nucleus showing a faint Xist signal (marked with an asterisk).

Fig. 4.

Expression of XistΔA RNA in preimplantation embryos. RNA-FISH was carried out in XXΔAand XX embryos at the eight-cell and blastocyst stages. At the eight-cell stage, although the accumulation of Xist RNA was evident in XX embryos, expression of XistΔARNA in XXΔA was very faint (arrowheads) and subsequently became undetectable by the blastocyst stage. Inset in the panel of the 8-cell XXΔA embryo is an enlargement of the nucleus showing a faint Xist signal (marked with an asterisk).

Subsequently, we examined whether the maternal XΔA could undergo inactivation in the embryonic lineage, where the X chromosome imprint is no longer effective, by using wild-type males carrying EGFP transgenes on the single X (XGFP) (Nakanishi et al., 2002) as the father. It has been shown that the expression of the transgene reflects the activity of XGFP(Ohhata et al., 2004). In wild-type female embryos recovered at E7.5, GFP fluorescence was observed in the embryonic tissue but not in the extraembryonic tissue, as expected(Fig. 3E). This substantiates that the transgene used here serves as a good reporter for addressing the activity of the X bearing it at this stage of embryo development. Female embryos heterozygous for XistΔA,which were morphologically indistinguishable from their wild-type littermates,were essentially negative for GFP throughout the embryo, suggesting that XGFP was uniformly inactivated in the embryonic lineage, which is normally subject to random X-inactivation(Fig. 3E). Furthermore, allelic expression analysis of G6pd and Hprt revealed that both genes were expressed exclusively from XΔA, and that the transcripts from XJF1 were barely detectable in XΔAXJF1 heterozygotes(Fig. 3F). These results indicate that, as is the case with the X carrying the dysfunctional Xist allele, XΔA is incompetent to undergo inactivation in the embryonic lineage, as well as in the extraembryonic lineages.

Expression of Xist is diminished on the mutated X in the preimplantation embryo

The above finding demonstrates that the A-repeat plays an essential role in X-inactivation during mouse development. This is consistent with the previous report by Wutz et al. that Xist RNAs lacking the A-repeat fail to initiate X-inactivation in transgenic ES cells(Wutz et al., 2002). In particular, one of the mutated Xist RNAs tested by Wutz et al.(ΔSX), which lacks almost the same region as the XistΔA RNA expressed from XΔA, is defective in silencing despite its accumulation on the X chromosome. This observation predicts that XistΔA RNA coats XΔA but fails to induce chromosomal silencing at the onset of X-inactivation in the embryo. To examine whether this was the case, RNA-FISH was carried out using an Xist-specific RNA probe in the preimplantation embryo, in which only the paternal copy of Xist is expressed and accumulated in cis (Kay et al., 1993; Sheardown et al.,1997). Embryos were recovered from wild-type females crossed with XistΔA/Y males at the eight-cell and blastocyst stages. All the female embryos should inherit the XistΔA allele in this cross. The sex of each embryo was identified by painting with X- and Y-specific probes afterwards (data not shown). RNA-FISH demonstrated that, although the expression of XistΔA was detected in XXΔA eight-cell embryos, the hybridization signal was very faint, essentially like a pinpoint(Fig. 4), and eventually disappeared at the blastocyst stage (Fig. 4). In control female embryos, accumulation of Xist RNA was detected as an intense signal at the eight-cell and blastocyst stages(Fig. 4). In agreement with these observations, real-time RT-PCR on total RNA of individual blastocysts demonstrated that the level of Xist RNA in XXΔA was much lower than that in XX. The level in XXΔA was, in fact,almost the same as that in XY, which was nearly below the detection limit(Fig. 5A). This excluded the possibility that XistΔA RNA,although expressed in the blastocyst, failed to coat the mutated paternal X chromosome, and is in contrast to human XIST RNA lacking the A-repeat expressed in tumor cells by an inducible promoter, which does not coat the X chromosome (Chow et al., 2007). These results raised the unexpected possibility that the failure of XΔA to undergo inactivation was primarily due to the lack of XistΔA RNA coating the mutated X.

Tsix is ectopically activated on the paternal XΔA in the blastocyst

Available evidence suggests that Tsix, the expression of which becomes detectable as early as the eight-cell to morula stage (Y.H. and T.S.,unpublished) and is confined to the maternal allele in the preimplantation embryo, prevents the upregulation of Xist on the maternal X during the process of imprinted X-inactivation(Lee, 2000; Sado et al., 2001). Given this negative effect of Tsix on Xist expression, it was of interest to explore whether the expression of Tsix was affected on the paternal XΔA at the blastocyst stage. Accordingly, RNA fractions of single genotyped blastocysts recovered from XJF1XJF1 females crossed with XΔAY males were individually converted into cDNA using gene-specific primers and subjected to two-round PCR. The parental origin of the Tsixtranscripts was subsequently addressed by restriction digestion with BsmAI, the recognition site of which is present only on the laboratory strain-derived X chromosome(Sugimoto and Abe, 2007). Intriguingly, the paternal copy of Tsix, which is normally silent at the blastocyst stage, was ectopically expressed from the mutated paternal XΔA in XXΔA. Such ectopic activation of Tsix was not observed in blastocysts that inherited either the wild-type X or the X chromosome carrying another Xist mutant allele, Xist1lox (Sado et al.,2005), from the father (Fig. 5B; data not shown). This finding raised an interesting possibility: that the unexpected silencing of Xist in XXΔA embryos during preimplantation development might be ascribed to the ectopic activation of the normally silent paternal copy of Tsix on XΔA. Given the proposed function of Tsix in the establishment of repressive chromatin in the Xist promoter region (Navarro et al., 2005; Sado et al.,2005; Sun et al.,2006), it is reasonable to assume that Tsix ectopically activated on XΔA attracts repressive chromatin modifications,which are not normally associated with the paternal allele, onto the Xist promoter region in cis in the tissues where X-inactivation is imprinted. Accordingly, the methylation status of 19 CpG sites in the Xist promoter (McDonald et al.,1998) was examined on the paternal XΔA in the trophoblast isolated from XJF1XΔA embryos at E6.5. The methylation profile of the maternal JF1-type allele could not be addressed here because the trophoblast samples inevitably contained maternal tissues with two XJF1 chromosomes. Bisulfite sequencing revealed that whereas the paternal allele in XJF1XB6 embryos was essentially unmethylated, as previously described(McDonald et al., 1998),significant methylation was evident on the paternal XΔA in XJF1XΔA (Fig. 5C). It seems possible therefore that the ectopic expression of Tsix facilitates CpG methylation at the Xist promoter on the paternal XΔA in the trophoblast.

Fig. 5.

Xist and Tsix are aberrantly regulated in XXΔA embryos. (A) Quantitative RT-PCR analysis of individual F1 blastocysts recovered from XJF1XJF1 females crossed with either XY or XΔAY males. The level of XistΔA RNA in XJF1XΔA was much lower than that of wild-type Xist in XX blastocysts. (B) Normally, a silent copy of paternal Tsix was ectopically expressed in XJF1XΔA blastocysts upon paternal transmission of the XistΔA allele. Following two-round PCR on cDNA prepared from single F1 blastocysts, the amplified products were digested with BsmAI, the recognition site of which is present only in the laboratory strain. (C) Methylation profile of the Xist promoter region on the paternal X in the trophoblast isolated from XX and XXΔA embryos at E6.5 was analyzed by bisulfite sequencing. It was evident that the methylation level was significantly higher in XXΔA than that in XX, suggesting that the expression of normally silent Tsix on the paternal XΔA induces aberrant methylation of the Xist promoter region in cis. An arrow indicates the position of the transcription start site of Xist.

Fig. 5.

Xist and Tsix are aberrantly regulated in XXΔA embryos. (A) Quantitative RT-PCR analysis of individual F1 blastocysts recovered from XJF1XJF1 females crossed with either XY or XΔAY males. The level of XistΔA RNA in XJF1XΔA was much lower than that of wild-type Xist in XX blastocysts. (B) Normally, a silent copy of paternal Tsix was ectopically expressed in XJF1XΔA blastocysts upon paternal transmission of the XistΔA allele. Following two-round PCR on cDNA prepared from single F1 blastocysts, the amplified products were digested with BsmAI, the recognition site of which is present only in the laboratory strain. (C) Methylation profile of the Xist promoter region on the paternal X in the trophoblast isolated from XX and XXΔA embryos at E6.5 was analyzed by bisulfite sequencing. It was evident that the methylation level was significantly higher in XXΔA than that in XX, suggesting that the expression of normally silent Tsix on the paternal XΔA induces aberrant methylation of the Xist promoter region in cis. An arrow indicates the position of the transcription start site of Xist.

CpG sites in the Xist promoter on XΔA are not methylated in sperm

The methylation profile of the Xist promoter was further analyzed in sperm to see whether the aberrant methylation of the paternal allele in the trophoblast was derived from sperm or acquired after fertilization. It has been shown previously that the methylation level of the Xist promoter is relatively low in oocytes, sperm and the preimplantation embryo(McDonald et al., 1998). Bisulfite sequencing revealed that the Xist promoter in sperm exhibited even lower methylation in mutant XΔAY males than in wild-type males (Fig. 6A). The significance of this difference in the methylation levels is not known at present, but this result indicates that the aberrant methylation found in the Xist promoter on the paternal XΔA in the trophoblast arises during embryogenesis.

Fig. 6.

Methylation profile of the Xist promoter region and Xite HS6 in sperm recovered from XY and XΔAY males. (A) Although the Xist promoter region was partially methylated in sperm recovered from XY as previously reported (McDonald et al., 1998), it exhibited even lower methylation in the sperm of XΔAY males. (B) The Xite HS6 region, the methylation of which in sperm was suggested to be an imprint responsible for the repression of paternal Tsix in the tissues or cells showing imprinted X-inactivation, was heavily methylated in sperm of both XY and XΔAY males.

Fig. 6.

Methylation profile of the Xist promoter region and Xite HS6 in sperm recovered from XY and XΔAY males. (A) Although the Xist promoter region was partially methylated in sperm recovered from XY as previously reported (McDonald et al., 1998), it exhibited even lower methylation in the sperm of XΔAY males. (B) The Xite HS6 region, the methylation of which in sperm was suggested to be an imprint responsible for the repression of paternal Tsix in the tissues or cells showing imprinted X-inactivation, was heavily methylated in sperm of both XY and XΔAY males.

It has been proposed that the differentially methylated domains in Tsix and Xite found between oocytes and sperm might constitute primary marks for the imprinted expression of Tsix in the zygote (Boumil et al., 2006). One of these domains, HS6, in Xite has been relatively well characterized by bisulfite sequencing and has been shown to be heavily methylated in sperm. It was postulated that if the methylation of this region in sperm was causally involved in the repression of paternal Tsix in the preimplantation embryo, it might be abolished in sperm isolated from XΔAY males. We therefore assessed the methylation status of this region in sperm of the mutant males. As shown in Fig. 6B, this region was heavily methylated on XΔA as on the wild-type X, suggesting that the methylation status of HS6 in the Xite region was not directly involved in the ectopic activation of Tsix on the paternal XΔA chromosome.

By taking advantage of an inducible expression system of a single copy Xist cDNA with various deletions from the single X chromosome in male ES cells, Wutz et al. showed that the silencing function of Xist RNA is dramatically compromised if the A-repeat is deleted, even though the mutant forms of the RNA accumulate in cis on the X chromosome(Wutz et al., 2002). This finding was further corroborated by the functional assay of Xist RNA lacking the A-repeat expressed from the endogenous locus by the inducible promoter in male ES cells (Wutz et al.,2002). Although that study highlighted for the first time the possible functional domain of Xist RNA that is crucial for inducing chromosome-wide silencing, the importance of the A-repeat for X-inactivation occurring in the mouse embryo had not been addressed until this study. The XistΔA allele was created so that it would produce a transcript nearly the same as the one lacking the A-repeat expressed from the endogenous Xist locus upon induction in the study by Wutz et al. (Wutz et al.,2002). Genetic and molecular analyses of female embryos heterozygous for XistΔA clearly demonstrated that the deletion of the A-repeat rendered the mutated X incompetent to undergo inactivation, indicating the crucial role of the A-repeat in X-inactivation during mouse development. The presence or absence of the A-repeat, however, does not seem to be directly involved in the primary choice of X-inactivation in the embryonic lineage, as different deletions in the Xist gene, although they retain the A-repeat, have resulted in primary non-random X-inactivation similar to that observed in this study(Marahrens et al., 1998; Sado et al., 2005).

We initially postulated that the failure of X-inactivation could be ascribed to the defect in XistΔARNA, which should be capable of coating the mutated X based on the inducible expression assay in ES cells (Wutz et al.,2002). Intriguingly, it was found that the level of XistΔA RNA was greatly reduced in XXΔA preimplantation embryos compared with that of wild-type Xist RNA in XX embryos. This could be due to a reduction either in the stability of the mutated RNA or in the expression level per se. Quantitative RT-PCR demonstrated, however, that the stability of XistRNA detected in undifferentiated male ES cells was comparable regardless of the presence or absence of the A-repeat, suggesting that the latter possibility was more favorable. It is likely therefore that the reduction in the level of Xist RNA in preimplantation embryos is primarily due to the transcriptional silencing of Xist. Our result demonstrates that the region encoding the A-repeat plays a crucial role as a regulatory element in the appropriate regulation of Xist in vivo.

Intriguingly, the lack of Xist RNA in the blastocyst was accompanied by an ectopic activation of the normally silent paternal copy of Tsix on the same X chromosome. Furthermore, the Xistpromoter on the mutated paternal X in the trophoblast at E6.5 was aberrantly methylated at CpG sites that are normally unmethylated on the paternal X. These findings raised an interesting possibility that the transcriptional silencing of Xist in preimplantation embryos is triggered by ectopically expressed Tsix, which subsequently promotes CpG methylation in the Xist promoter region on the mutated paternal X. In this scenario, the deleted region in the XistΔA allele, most probably the A-repeat, is crucial for the appropriate repression of Tsix on the paternal X at the onset of imprinted X-inactivation. Given the fact that the transcription of Tsix is initiated 40 kb downstream from the A-repeat, it is tempting to speculate that the A-repeat exerts its effect on Tsix through a long-range chromatin conformation. However, the opposite scenario is also possible: the upregulation of Tsix was somehow caused by the primary silencing of Xist that resulted from the loss of some crucial regulatory element located within the deleted region. Because the aberrant methylation of the Xist promoter appears to be established during embryogenesis, it might be expected that the mutated XistΔA RNA would be transcribed from the paternal XΔA in the early preimplantation embryo. This was not the case, however, and the expression of paternal XistΔA was diminished from the very early stages. This observation may favor the later scenario that the silencing of Xist is the primary event. These two possibilities cannot be distinguished between on the basis of current data and the further experimentation is certainly required. The simplest way to address this issue is to terminate Tsix on XΔA and see whether XistΔA is expressed or not. We are currently trying to produce mice carrying a Tsix deficiency on XΔA through second gene targeting in ΔA2lox211 ES cells.

Although the targeted deletion of the A-repeat did not allow us to address the functional significance of the A-repeat as an element in the XistRNA because of the unexpected lack of expression from the mutated X, this study clearly demonstrates that the region encoding the A-repeat is essential as a genomic element for X-inactivation in the mouse embryo. Further attempts to identify the factors that interact with the DNA sequence harboring the A-repeat should provide further insight into the molecular mechanisms of Xist/Tsix regulation and the random choice of X-inactivation.

This work was supported by grants from PRESTO, Japan,Science and Technology Agency and from the Ministry of Education, Science, Sports and Culture of Japan to T.S.

Boumil, R. M., Ogawa, Y., Sun, B. K., Huynh, K. D. and Lee, J. T. (
2006
). Differential methylation of Xite and CTCF sites in Tsix mirrors the pattern of X-inactivation choice in mice.
Mol. Cell. Biol.
26
,
2109
-2117.
Brockdorff, N., Ashworth, A., Kay, G. F., McCabe, V. M., Norris,D. P., Cooper, P. J., Swift, S. and Rastan, S. (
1992
). The product of the mouse Xist gene is a 15 kb inactive X-specific transcript containing no conserved ORF and located in the nucleus.
Cell
71
,
515
-526.
Brown, C. J., Hendrich, B. D., Rupert, J. L., Lafreniere, R. G.,Xing, Y., Lawrence, J. and Willard, H. F. (
1992
). The human XIST gene: analysis of a 17 kb inactive X-specific RNA that contains conserved repeats and is highly localized within the nucleus.
Cell
71
,
527
-542.
Chow, J. C., Hall, L. L., Baldry, S. E., Thorogood, N. P.,Lawrence, J. B. and Brown, C. J. (
2007
). Inducible XIST-dependent X-chromosome inactivation in human somatic cells is reversible.
Proc. Natl. Acad. Sci. USA
104
,
10104
-10109.
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.
Clemson, C. M., McNeil, J. A., Willard, H. F. and Lawrence, J. B. (
1996
). XIST RNA paints the inactive X chromosome at interphase: evidence for a novel RNA involved in nuclear/chromosome structure.
J. Cell Biol.
132
,
259
-275.
Goto, Y. and Takagi, N. (
1998
). Tetraploid embryos rescue embryonic lethality caused by an additional maternally inherited X chromosome in the mouse.
Development
125
,
3353
-3363.
Huynh, K. D. and Lee, J. T. (
2003
). Inheritance of a pre-inactivated paternal X chromosome in early mouse embryos.
Nature
426
,
857
-862.
Kay, G. F., Penny, G. D., Patel, D., Ashworth, A., Brockdorff,N. and Rastan, S. (
1993
). Expression of Xist during mouse development suggests a role in the initiation of X chromosome inactivation.
Cell
72
,
171
-182.
Lee, J. T. (
2000
). Disruption of imprinted X inactivation by parent-of-origin effects at Tsix.
Cell
103
,
17
-27.
Li, E., Bestor, T. H. and Jaenisch, R. (
1992
). Targeted mutation of the DNA methyltransferase gene results in embryonic lethality.
Cell
69
,
915
-926.
Lyon, M. (
1961
). Gene action in the X-chromosome of the mouse (Mus musculus L).
Nature
190
,
372
-373.
Mak, W., Nesterova, T. B., de Napoles, M., Appanah, R.,Yamanaka, S., Otte, A. P. and Brockdorff, N. (
2004
). Reactivation of the paternal X chromosome in early mouse embryos.
Science
303
,
666
-669.
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.
Marahrens, Y., Loring, J. and Jaenisch, R.(
1998
). Role of the Xist gene in X chromosome choosing.
Cell
92
,
657
-664.
McDonald, L. E., Paterson, C. A. and Kay, G. F.(
1998
). Bisulfite genomic sequencing-derived methylation profile of the xist gene throughout early mouse development.
Genomics
54
,
379
-386.
Nakanishi, T., Kuroiwa, A., Yamada, S., Isotani, A., Yamashita,A., Tairaka, A., Hayashi, T., Takagi, T., Ikawa, M., Matsuda, Y. et al.(
2002
). FISH analysis of 142 EGFP transgene integration sites into the mouse genome.
Genomics
80
,
564
-574.
Navarro, P., Pichard, S., Ciaudo, C., Avner, P. and Rougeulle,C. (
2005
). Tsix transcription across the Xist gene alters chromatin conformation without affecting Xist transcription: implications for X-chromosome inactivation.
Genes Dev.
19
,
1474
-1484.
Nesterova, T. B., Slobodyanyuk, S. Y., Elisaphenko, E. A.,Shevchenko, A. I., Johnston, C., Pavlova, M. E., Rogozin, I. B., Kolesnikov,N. N., Brockdorff, N. and Zakian, S. M. (
2001
). Characterization of the genomic Xist locus in rodents reveals conservation of overall gene structure and tandem repeats but rapid evolution of unique sequence.
Genome Res.
11
,
833
-849.
Ohhata, T., Tachibana, M., Tada, M., Tada, T., Sasaki, H.,Shinkai, Y. and Sado, T. (
2004
). X-inactivation is stably maintained in mouse embryos deficient for histone methyl transferase G9a.
Genesis
40
,
151
-156.
Ohhata, T., Hoki, Y., Sasaki, H. and Sado, T.(
2008
). Crucial role of antisense transcription across the Xist promoter in Tsix-mediated Xist chromatin modification.
Development
135
,
227
-235.
Okamoto, I., Tan, S. and Takagi, N. (
2000
). X-chromosome inactivation in XX androgenetic mouse embryos surviving implantation.
Development
127
,
4137
-4145.
Okamoto, I., Otte, A. P., Allis, C. D., Reinberg, D. and Heard,E. (
2004
). Epigenetic dynamics of imprinted X inactivation during early mouse development.
Science
303
,
644
-649.
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.
Sado, T., Wang, Z., Sasaki, H. and Li, E.(
2001
). Regulation of imprinted X-chromosome inactivation in mice by Tsix.
Development
128
,
1275
-1286.
Sado, T., Hoki, Y. and Sasaki, H. (
2005
). Tsix silences Xist through modification of chromatin structure.
Dev. Cell
9
,
159
-165.
Sado, T., Hoki, Y. and Sasaki, H. (
2006
). Tsix defective in splicing is competent to establish Xist silencing.
Development
133
,
4925
-4931.
Sheardown, S. A., Duthie, S. M., Johnston, C. M., Newall, A. E.,Formstone, E. J., Arkell, R. M., Nesterova, T. B., Alghisi, G. C., Rastan, S. and Brockdorff, N. (
1997
). Stabilization of Xist RNA mediates initiation of X chromosome inactivation.
Cell
91
,
99
-107.
Sugimoto, M. and Abe, K. (
2007
). X chromosome reactivation initiates in nascent primordial germ cells in mice.
PLoS Genet.
3
,
e116
.
Sun, B. K., Deaton, A. M. and Lee, J. T.(
2006
). A transient heterochromatic state in Xist preempts X inactivation choice without RNA stabilization.
Mol. Cell
21
,
617
-628.
Tada, T., Takagi, N. and Adler, I. D. (
1993
). Parental imprinting on the mouse X chromosome: effects on the early development of X0, XXY and XXX embryos.
Genet. Res.
62
,
139
-148.
Takagi, N. and Sasaki, M. (
1975
). Preferential inactivation of the paternally derived X chromosome in the extraembryonic membranes of the mouse.
Nature
256
,
640
-642.
Wutz, A., Rasmussen, T. P. and Jaenisch, R.(
2002
). Chromosomal silencing and localization are mediated by different domains of Xist RNA.
Nat. Genet.
30
,
167
-174.