male-specific lethal 2


DEVELOPMENTAL BIOLOGY

Embryonic

In embryonic nuclei, MSLs and monoacetylated histone H4 associate with the X chromosome. This was established using double fluorescent in situ hybridization with an X chromosome DNA probe and immunostaining for MSL protein. The late blastoderm/early gastrula are the earliest stages at which the MSLs are detected in the subnuclear localization pattern characteristic of male embryos. Earlier than this, MSLs are rather homogeneously distributed in the nucleus. It seems that MSLs become localized simultaneously, or over a very short period of time. MSL localization to the X chromosome without colocalization of monoacetylated H4 has been thought to suggest that MSLs localize prior to histone H4 (Franke, 1996).

Effects of Mutation or Deletion

There is no significant amount of embryonic death caused by msl-2 mutation, and no significant effect of msl-2 on viability of females or heterozygous males. Although most msl-2 males are killed in the larval stages, some do reach the prepupal stage. Mutants develop very slowly, most remaining in the third-instar stage for several days until dying as larvae or in the process of forming a puparium (Belote, 1980).

Msl-2 is required for the male-specific assembly of a dosage compensation regulatory complex on the X chromosome of Drosophila. Msl-2 binds in a reproducible, partial pattern to the male X chromosome in the absence of Mle or Msl-3, or when ectopically expressed at a low level in females. Moreover, the pattern of Msl-2 binding corresponds precisely in each case to that of MSL-1, suggesting that the two proteins function together to associate with the X chromosome. When MSL-1 function is compromised, Msl-2 is only localized to sites where the mutant form of MSL-1 is bound. Thus, a partial Msl-2 pattern is observed in msl-1 mutants. Likewise, when Msl-2 is present in limiting quantities, MSL-1 is precisely restricted to the sites of Msl-2 localization. Ectopic Msl-2 expression in females results in a dominant female phenotype. EMS-induced loss of function msl-1 and msl-2 alleles were isolated in a screen for suppressors of the toxic effects of Msl-2 expression in females. One such mutant lacks the RING finger motif. Site-directed mutagenesis was also used to determine the importance of the Msl-2 RING finger domain and second cysteine-rich motif. The mutations, including those in conserved zinc coordinating cysteines, confirm that the RING finger is essential for Msl-2 function, while suggesting a less stringent requirement for an intact second motif, the metalothionein-like cysteine cluster known as a PHD finger (Lyman, 1997).

The binding of each of the MSLs, and of histone H4 acetylated on lysine 16, to the X chromosome requires the presence of all four MSL proteins. For example, no association of the MSL-3 protein with the X chromosome is detected in msl-1 mutants. Throughout embryogenesis, in none of the homozygous male embryos mutant for msl-2 or msl-3 could association of the other MSL proteins or monoacetylated histone with the X chromosome be detected. msl-1 mutant embryos show association of MSL proteins with X chromosomes up to stage 9, while in mle mutants such association is seen up to stage 10, but this binding can be attributed to the presence of maternal MLE and MSL-1 (Franke, 1996).

In Drosophila, dosage compensation requires assembly of the Male Specific Lethal (MSL) protein complex for doubling transcription of most X-linked genes in males. The recognition of the X chromosome by the MSL complex has been suggested to include initial assembly at ~35 chromatin entry sites and subsequent spreading of mature complexes in cis to numerous additional sites along the chromosome. To understand this process further, MSL patterns were examined in a range of wild-type and mutant backgrounds producing different amounts of MSL components. The data support a model in which MSL complex binding to the X is directed by a hierarchy of target sites that display different affinities for the MSL proteins. Chromatin entry sites differ in their ability to provide local intensive binding of complexes to adjacent regions, and need high MSL complex titers to achieve this. A set of definite autosomal regions (~70), competent to associate with the functional MSL complex in wild-type males, was mapped. Overexpression of both MSL1 and MSL2 stabilizes this binding and results in inappropriate MSL binding to the chromocenter and the 4th chromosome. Thus, wild-type MSL complex titers are critical for correct targeting to the X chromosome (Demakova, 2003).

A functional dosage compensation complex required for male killing in Drosophila

Bacteria that selectively kill males ('male-killers') were first characterized more than 50 years ago in Drosophila and have proved to be common in insects. However, the mechanism by which sex specificity of virulence is achieved has remained unknown. This study tested the ability of Spiroplasma poulsonii to kill Drosophila melanogaster males carrying mutations in genes that encode the dosage compensation complex. The bacterium failed to kill males lacking any of the five protein components of the complex (Veneti, 2005).

Certain isofemale lines of Drosophila only give rise to daughters following the death of male embryos. Male death is due to the presence of intracellular bacteria that pass from a female to her progeny and that selectively kill males during embryogenesis. These male-killing bacteria are found in a wide range of other insect species, and many different bacteria have evolved male-killing phenotypes. In some host species, male-killers drive the host population sex ratio to levels as high as 100 females per male and alter the pattern of mate competition. However, the underlying processes that produce male-limited mortality are unclear. This study examined the interaction between the male-killing bacterium Spiroplasma poulsonii and the sex determination pathway of D. melanogaster (Veneti, 2005).

The primary signal of sex in Drosophila is the X-to-autosome ratio. This signal is permanently established in expression of Sex-lethal (Sxl) in females and its absence in males. This, in turn, effects three processes: germline sexual identity, somatic sexual differentiation, and dosage compensation, the process by which the gene expression titer on the X chromosome is equalized between two sexes despite their difference in X chromosome number. Mutations in the gene tra that convert XX individuals to male somatic sex do not induce female death. These observations indicate germline formation and migration happen correctly in male embryos and that dying male embryos do not express Sxl. Therefore, the requirement of the Spiroplasma for genes within the system of dosage compensation was examined (Veneti, 2005).

In Drosophila, the single X of males is hypertranscribed. This process of hypertranscription requires the formation of the dosage compensation complex (DCC) and its binding to (and modification of) the X chromosome. SXL in female Drosophila inhibits the production of MSL-2 protein, which is thus only present in male Drosophila. MSL-2 forms a complex with four other proteins, MSL-1, MSL-3, MLE, and MOF, which collectively form the DCC. MSL-1, MSL-3, MLE, and MOF are constitutively present in both males and females and are also supplied maternally. The complete DCC binds, with JIL-1, to the male X chromosome at various entry points, and, with the products of two noncoding RNAs, RoX1 and RoX2, it affects the modification of the single X chromosome and its hypertranscription (Veneti, 2005).

The effect was examined of mutations within the host DCC on the ability of the male-killer to function. The survival of male progeny beyond embryogenesis to L2/L3 (and in one case adult) was scored in the presence of different loss-of-function mutations within the dosage compensation system (normal male-killing occurs during embryogenesis), in the presence and absence of infection. Because many genes within this group additionally show strong maternal effects, the effect of mutations was in each case tested by using both mothers that were heterozygous for the loss-of-function mutation and mothers homozygous for it (Veneti, 2005).

Uninfected males homozygous for loss-of-function mutations within the dosage compensation system generally survive to the third larval instar. Survival to the third larval instar was studied for loss-of-function alleles of msl-1 (alleles msl-11 and msl-1b), msl-2 (msl-2g227 and msl-2g134), msl-3 (msl-3132), mle (mle9, mle1), and mof (mof1), and survival to adult for mle1/mle6 transheterozygotes. In the case of all alleles of msl-1, msl-3, mle, and mof, a similar pattern is observed: Males homozygous or hemizygous for the loss-of-function mutation have appreciable survival in the presence of infection when their mother is also homozygous for the loss-of-function mutation. In contrast, heterozygous male embryos that were siblings of the above (that will have a wild-type DCC) were always killed, as were all male embryos in the case where the mother was heterozygous (where maternal supply of these proteins enables dosage compensation to be initiated, although not maintained). In the case of mle1/mle6 transheterozygotes, male survival to adult was observed. For the case of mof, male-killing was restored to full efficiency when 18H1, a transgenic copy of mof, was added to the mof1 loss-of-function background. Within the above crosses, three observations make sure the observation that Spiroplasma was fully operational. First, crosses involving heterozygous mothers, where male-killing was complete, were conducted concurrently with those using homozygous mothers, and the females in these crosses were siblings from the same vials. Second, in each cross and vial where homozygous males survived, heterozygous males (with wild-type function) still died. Finally, F1 females derived from these crosses, when mated to wild-type males, produced a full, female-biased sex ratio (Veneti, 2005).

In the case of msl-2, where there is no maternal supply of MSL-2, survival of homozygous sons was observed for both homozygous and heterozygous mothers for the case of the mutation msl-2g134. For the case of msl-2 g227, no male progeny were observed from infected females (table S5). This mutation does not rescue males in the assay, probably because the two mutations have different effects on msl-2 expression. The msl-2g134 allele prevents formation of MSL-2 protein, whereas msl-2 g227 potentially encodes the RING-finger element of a truncated MSL-2 protein (Veneti, 2005).

Thus, absence or reduced function of any of the proteins of the DCC can reduce the efficiency of male killing, and a functional DCC is required for male killing by S. poulsonii. The fact that the genes mediating this process in Drosophila have been well studied can be exploited to yield further insights into the mechanism of male killing (Veneti, 2005).

X chromosome dosage compensation via enhanced transcriptional elongation in Drosophila

The evolution of sex chromosomes has resulted in numerous species in which females inherit two X chromosomes but males have a single X, thus requiring dosage compensation. MSL (Male-specific lethal) complex increases transcription on the single X chromosome of Drosophila males to equalize expression of X-linked genes between the sexes. The biochemical mechanisms used for dosage compensation must function over a wide dynamic range of transcription levels and differential expression patterns. It has been proposed that the MSL complex regulates transcriptional elongation to control dosage compensation, a model subsequently supported by mapping of the MSL complex and MSL-dependent histone 4 lysine 16 acetylation to the bodies of X-linked genes in males, with a bias towards 3' ends. However, experimental analysis of MSL function at the mechanistic level has been challenging owing to the small magnitude of the chromosome-wide effect and the lack of an in vitro system for biochemical analysis. This study used global run-on sequencing (GRO-seq) to examine the specific effect of the MSL complex on RNA Polymerase II (RNAP II) on a genome-wide level. Results indicate that the MSL complex enhances transcription by facilitating the progression of RNAP II across the bodies of active X-linked genes. Improving transcriptional output downstream of typical gene-specific controls may explain how dosage compensation can be imposed on the diverse set of genes along an entire chromosome (Larschan, 2011).

To investigate how the MSL complex specifically increases transcription of X-linked genes, GRO-seq was performed in SL2 cells, a male Drosophila cell line that has been extensively characterized for MSL function. To show the average enrichment across genes, a 3-kb 'metagene' profile was plotted in which the internal regions were rescaled so that all genes appear to have the same length. Analysis was restricted to expressed genes that were sufficiently large (>2.5 kb) so that gene-body effects could be clearly assessed (822 X-linked genes, 3,420 autosomal genes), and all gene profiles were normalized by their copy number as determined by analysis of SL2 DNA content. High correlation coefficients were observed between replicate libraries. The metagene profiles revealed a prominent 5' peak of paused RNAP II consistent with previous chromatin immunoprecipitation (ChIP) and analysis of short 5' RNAs. In addition, a peak of RNAP II density downstream of the metagene 3' processing site is evident, possibly due to slow release in regions of transcription termination. The 3' peak is present even when the influence of neighbouring gene transcription is eliminated (Larschan, 2011).

The central question with regard to dosage compensation is how genes on the X chromosome differ on average from genes on autosomes. Overall, it was found that RNAP II density on active X-linked genes was higher than on autosomal genes, specifically over gene bodies. The increase in tag density over the bodies of X-linked genes compared to autosomal genes was approximately 1.4-fold, consistent with previous estimates of MSL-dependent dosage compensation. RNAP II ChIP was performed in SL2 cells, confirming higher occupancy on X-linked genes compared to autosomes but with lower resolution and reduced sensitivity. Therefore, GRO-seq was performed to analyse X and autosomal differences (Larschan, 2011).

To measure how X and autosomes differed on average in the distribution of elongating RNAP II, genes were segmented into their 5' 500 bp and the remainder of the coding region. The remainder of the coding region was subdivided further into 5' and 3' segments (25% and 75%, respectively). Using this segmentation, RNAP II pausing and elongation were quantitated separately on the basis of the unscaled GRO-seq signal. The pausing index (PI) was previously defined as the ratio of the GRO-seq signal at the 5' peak to the average signal over gene bodies. Herewe calculated the PI was calculated for X and autosomal genes as the ratio of the 5' peak to the first 25% of the remaining gene body, and no statistically significant difference was found when the two groups were compared (Larschan, 2011).

To examine separately transcription elongation across gene bodies, the elongation density index (EdI) was defined as the ratio of tag density in the 3' region of each gene compared to its 5' region after the first 500 bp. In contrast to the analysis of 5' pausing, statistically significant differences was found in EdI between X and autosomes. This conclusion was robust to how the 5' and 3' regions of genes were divided. As defined, the average PI (log scale) is a positive number because RNAP II is generally enriched at 5' ends compared to gene bodies; the average EdI (log scale) is a negative number, as the relative density of RNAP II typically decreases from the beginning to the end of gene bodies. It is concluded that X-linked genes, on average, show a significantly smaller decrease in RNAP II density along their gene bodies when compared to autosomal genes (Larschan, 2011).

To measure the specific contribution of the MSL complex to the increase in RNAP II within X-linked gene bodies, MSL2 RNA interference (RNAi) was used to reduce complex levels in male SL2 cells. Excellent correlations between replicate data sets were observed. To confirm the X-specific effect of MSL2 RNAi, the distributions of the GRO-seq signal (averaged over the bodies of genes excluding the 5' peak) were computed for all genes before and after RNAi. When comparing X versus autosomes, a preferential decrease was found on the X chromosome, with an average control:MSL RNAi ratio of 1.4. MSL-dependent changes in average GRO-seq density showed a weak but statistically significant correlation with changes in steady-state messenger RNA levels assayed by expression array or mRNA-Seq10. These results confirm that MSL-dependent changes in steady-state RNA levels reflect differences in active transcription on the X chromosome (Larschan, 2011).

In addition to assessing the average decrease of X-linked RNAP II density after MSL2 RNAi, it was asked whether any genes showed strong MSL-dependence, a hallmark of the roX genes that encode RNA components of the complex. It was found that roX2 showed a strong loss in GRO-seq density (ninefold) after MSL2 RNAi, as predicted. Interestingly, in the untreated or control RNAi samples, there is a prominent GRO-seq peak downstream of the major roX2 3' end, coincident with an MSL recruitment site. roX1 expression is low in this isolate of SL2 cells, and no other expressed genes on X or autosomes showed strong MSL dependence in these assays (Larschan, 2011).

Next the average RNAP II density was compared along X and autosomal metagene profiles after control and MSL2 RNAi. Unlike the initial analysis of X and autosomes, where different gene populations were compared, here it was possible to examine the same genes in the presence and absence of the MSL complex. It was found that after MSL2 RNAi, the density of elongating RNAP II over the bodies of X-linked genes decreased, approaching the level on autosomes. The presence of the MSL complex affected RNAP II density starting just downstream of the 5' peak and continuing through the bodies of X-linked genes. Thus, GRO-seq functional data correlate with physical association of the MSL complex, which is biased towards the 3' ends of active genes on the male X chromosome (Larschan, 2011).

To quantify the differences in density of engaged RNAP II in the presence and absence of the MSL complex, the PI and EdI were calculated for each gene, followed by the PI and EdI ratios comparing MSL2 and control RNAi treatment. It was found that both X and autosomes increased PI and decreased EdI after MSL2 RNAi treatment. However, in each case the change was larger on X than on autosomes, and the most profound difference was an MSL-dependent change in EdI on X compared with autosomes. EdI was computed, as before, by defining the 5' and 3' regions as 25% and 75%, respectively, of the gene body after removing the 5' peak, but the difference was statistically significant for all other values until the 3' end was reached. When these analyses were performed separately for two independently prepared sets of GRO-seq libraries, the results were also statistically significant. It is concluded that the MSL complex causes the transcriptional elongation profiles of X-linked genes to differ from those of autosomal genes (Larschan, 2011).

To visualize the location along gene bodies at which the MSL complex functions, control:MSL2 RNAi GRO-seq ratios were calculated and a metagene profile was generated. Here, values above zero represent higher relative amounts of engaged RNAP II in the presence of the MSL complex compared to after RNAi treatment. In contrast, values below zero represent a relative increase in engaged RNAP II after MSL2 RNAi. In the absence of the MSL complex, there is a relative increase in the amount of RNAP II localized to the 5' ends of both autosomal and X-linked genes, perhaps due to relocalization of RNAP II from the bodies of X-linked genes. A limitation of the GRO-seq assay is that it is not possible to distinguish between initiating and 5' paused polymerase, so a definitive role cannot be assigned for this 5' increase in RNAP II after MSL2 RNAi treatment. However, relative RNAP II levels over autosomal gene bodies do not increase, indicating that any relocalized enzyme in this experiment is likely to remain paused rather than progressing across transcription units. This is consistent with a model in which the functional outcome of MSL2 RNAi is to shift RNAP II density away from productive transcription through X-linked gene bodies (Larschan, 2011).

The local effect of the MSL complex was plotted to compare it to the status of histone 4 lysine 16 (H4K16) acetylation catalysed by the MOF component of the MSL complex. H4K16 acetylation typically is enriched at the 5' ends of most active genes in mammals and flies; in contrast, a 3' bias of this mark is a distinctive characteristic of the dosage compensated male X chromosome in Drosophila. Interestingly, there is an overall coincidence across gene bodies between the MSL-complex-dependent GRO-seq signal and the presence of H4K16 acetylation. How might H4K16 acetylation biased towards the 3' end of genes generate the improved transcriptional elongation indicated by the GRO-seq results? During transcription elongation, nucleosomes are thought to comprise a barrier to the progress of RNAP II and several well-studied elongation factors, including Spt6 and the FACT complex, are proposed to function by removing nucleosomes that block RNAP II progression and replacing them in the wake of transcription. Interestingly, H4K16 acetylation of nucleosomes has been observed to act in opposition to the formation of higher-order chromatin structure in vitro. Thus, H4K16 acetylation is likely to reduce further the steric hindrance to RNAP II progression through chromatin. Improving the entry of RNAP II into the bodies of genes may allow 5', gene-specific events to proceed at an increased but still regulated rate. Furthermore, reduction in the repressive effect of nucleosomes could increase mRNA output by improving the processivity of RNAP II on each template. Available methodologies cannot distinguish between these mechanisms in vivo, and therefore future approaches will be required to assess their relative contributions to dosage compensation (Larschan, 2011).

In addition to increasing the transcription of X-linked genes for dosage compensation, the MSL complex also positively regulates the roX noncoding RNA components of the complex, to promote their male specificity. roX1 expression is low in the SL2 cell line, but GRO-seq data indicate that active transcription of roX2 is highly dependent on MSL2 as predicted. Interestingly, there is a strong GRO-seq peak at the 3' roX2 DHS (DNaseI hypersensitive site), which contains sequences important for targeting the MSL complex to the X chromosome. Sites of roX gene transcription are thought to be critical for MSL complex assembly. Therefore, it is possible that paused RNAP II at the roX2 DHS could promote an open chromatin structure that facilitates MSL complex targeting or incorporation of noncoding roX2 RNA into the complex (Larschan, 2011).

In summary, it is proposed that the MSL complex functions on the male X chromosome to promote progression and processivity of RNAP II through the nucleosomal template. Improving transcriptional output downstream of typical gene-specific regulation makes biological sense when compensating the diverse set of genes found along an entire chromosome (Larschan, 2011).


REFERENCES

Abaza, I., Coll, O., Patalano, S. and Gebauer, F. (2006). Drosophila UNR is required for translational repression of male-specific-lethal 2 mRNA during regulation of X-chromosome dosage compensation. Genes Dev. 20: 380-389. 16452509

Abaza, I. and Gebauer, F. (2008). Functional domains of Drosophila Unr in translational control. RNA 14: 482-490. PubMed Citation: 18203923

Alekseyenko, A. A., et al. (2012). Sequence-specific targeting of dosage compensation in Drosophila favors an active chromatin context. PLoS Genet. 8(4): e1002646. PubMed Citation: 22570616

Amrein, H. and Axel, R. (1997). Genes expressed in neurons of adult male Drosophila. Cell 88: 459-469. PubMed Citation: 9038337

Bashaw,G.J.and Baker,B.S. (1995). The msl-2 dosage compensation gene of Drosophila encodes a putative DNA-binding protein whose expression is sex specifically regulated by Sex-lethal. Development 121: 3245-3258. PubMed Citation: 7588059

Bashaw, G. J. and Baker, B. S. (1997). The regulation of Drosophila msl-2 gene reveals a function for Sex-lethal in translational control. Cell 89: 789-798. PubMed Citation: 9182767

Beckmann, K., et al. (2005). A dual inhibitory mechanism restricts msl-2 mRNA translation for dosage compensation in Drosophila. Cell 122: 529-540. 16122421

Belote, J., and Lucchesi, J. C. (1980). Male-specific lethal mutations in Drosophila melanogaster. Genetics 96: 165-185

Bernstein, M. and Cline, T. W. (1994). Differential effects of Sex-lethal mutations on dosage compensation early in Drosophila development. Genetics 136: 1051-61. PubMed Citation: 8005414

Bhadra, U., Pal-Bhadra, M. and Birchler, J. A. (1999). Role of the male specific lethal (msl) genes in modifying the effects of sex chromosomal dosage in Drosophila. Genetics 152(1): 249-268. PubMed Citation: 10224258

Bhadra, U., Pal-Bhadra, M. and Birchler, J. A. (2000). Histone acetylation and gene expression analysis of Sex lethal mutants in Drosophila. Genetics 155: 753-763. PubMed Citation: 10835396

Braunstein, M., et al. (1993). Transcriptional silencing in yeast is associated with reduced nucleosome acetylation. Genes Dev. 7: 592-604. PubMed Citation: 8458576

Bone, J. R., et al. (1994). Acetylated histone H4 on the male X chromosome is associated with dosage compensation in Drosophila. Genes Dev. 8: 96-104. PubMed Citation: 8288132

Buscaino, A., et al. (2003). MOF-regulated acetylation of MSL-3 in the Drosophila dosage compensation complex. Molec. Cell 11: 1265-1277. 12769850

Chang, K. A. and Kuroda, M. I. (1998). Modulation of MSL1 abundance in female Drosophila contributes to the sex specificity of dosage compensation. Genetics 150(2): 699-709. PubMed Citation: 9755201

Copps, K., et al. (1998). Complex formation by the Drosophila MSL proteins: role of the MSL2 RING finger in protein complex assembly. EMBO J. 17: 5409-5417. PubMed Citation: 9755201

Dahlsveen, I. K., Gilfillan, G. D., Shelest, V. I., Lamm, R. and Becker, P. B. (2006). Targeting determinants of dosage compensation in Drosophila. PLoS Genet. 2(2): e5. 16462942

Demakova, O. V., et al. (2003). The MSL complex levels are critical for its correct targeting to the chromosomes in Drosophila melanogaster. Chromosoma 112(3): 103-15. 14579126

Duncan, K., et al. (2006). Sex-lethal imparts a sex-specific function to UNR by recruiting it to the msl-2 mRNA 3' UTR: translational repression for dosage compensation. Genes Dev. 20: 368-379. 16452508

Forch, P., Merendino, L., Martinez, C. and Valcarcel, J. (2001). Modulation of msl-2 5' splice site recognition by Sex-lethal. RNA 7(9): 1185-91. 11565743

Franke, A., et al. (1996). Evidence that MSL-mediated dosage compensation in Drosophila begins at blastoderm. Development 122 (9): 2751-2760. PubMed Citation: 8787749

Franke, A. and Baker, B. S. (1999). The rox1 and rox2 RNAs are essential components of the compensasome, which mediates dosage compensation in Drosophila. Mol. Cell 4: 117-122. PubMed Citation: 10445033

Friedman, J. R., et al. (1996). KAP-1, a novel corepressor for the highly conserved KRAB repression domain. Genes Dev. 10: 2067-78. PubMed Citation: 8769649

Gebauer, F., et al. (1998). The Drosophila splicing regulator Sex-lethal directly inhibits translation of Male-specific-lethal 2 mRNA. RNA 4(2): 142-50. PubMed Citation: 9570314

Gebauer, F., et al. (1999). Translational control of dosage compensation in Drosophila by Sex-lethal: cooperative silencing via the 5' and 3' UTRs of msl-2 mRNA is independent of the poly(A) tail. EMBO J. 18: 6146-6154. PubMed Citation: 10545124

Gebauer, F., Grskovic, M. and Hentze, M. W. (2003). Drosophila Sex-lethal inhibits the stable association of the 40S ribosomal subunit with msl-2 mRNA. Molec. Cell 11: 1397-1404. 12769862

Gladstein, N., McKeon, M. N. and Horabin, J. I. (2010). Requirement of male-specific dosage compensation in Drosophila females--implications of early X chromosome gene expression. PLoS Genet 6: e1001041. PubMed ID: 20686653

Gorchakov, A. A., et al. (2009). Long-range spreading of dosage compensation in Drosophila captures transcribed autosomal genes inserted on X. Genes Dev. 23(19): 2266-71. PubMed Citation: 19797766

Graham, P. L., Yanowitz, J. L., Penn, J. K., Deshpande, G. and Schedl, P. (2011). The translation initiation factor eIF4E regulates the sex-specific expression of the master switch gene Sxl in Drosophila melanogaster. PLoS Genet. 7(7): e1002185. PubMed Citation: 21829374

Greenberg, A. J., Yanowitz, J. L. and Schedl, P. (2004). The Drosophila GAGA factor is required for dosage compensation in males and for the formation of the male-specific-lethal complex chromatin entry site at 12DE. Genetics 166: 279-289. 15020425

Grimaud, C. and Becker, P. B. (2009). The dosage compensation complex shapes the conformation of the X chromosome in Drosophila. Genes Dev. 23(21): 2490-5. PubMed Citation: 19884256

Grskovic, M., Hentze, M. W. and Gebauer, F. (2003). A co-repressor assembly nucleated by Sex-lethal in the 3'UTR mediates translational control of Drosophila msl-2 mRNA. EMBO J. 22: 5571-5581. 14532129

Gu, W., Szauter, P. and Lucchesi, J. C. (1998). Targeting of MOF, a putative histone acetyl transferase, to the X chromosome of Drosophila melanogaster. Dev. Genet. 22(1): 56-64. 9816055

Hamada, F. N., et al. (2005). Global regulation of X chromosomal genes by the MSL complex in Drosophila melanogaster. Genes Dev. 19: 2289-2294. 16204180

Hennig, J., Militti, C., Popowicz, G. M., Wang, I., Sonntag, M., Geerlof, A., Gabel, F., Gebauer, F., Sattler, M. (2014) Structural basis for the assembly of the Sxl-Unr translation regulatory complex. Nature. PubMed ID: 25209665

Hilfiker, A., et al. (1995). The gene virilizer is required for female-specific splicing controlled by Sxl, the master gene for sexual development in Drosophila. Development 121: 4017-4026. 8575302

Hilfiker, A., et al. (1997). mof, a putative acetyl transferase gene related to the Tip60 and MOZ human genes and to the SAS genes of yeast, is required for dosage compensation in Drosophila. EMBO J. 16(8): 2054-2060. PubMed Citation: 9155031

Ilik, I. A., Quinn, J. J., Georgiev, P., Tavares-Cadete, F., Maticzka, D., Toscano, S., Wan, Y., Spitale, R. C., Luscombe, N., Backofen, R., Chang, H. Y. and Akhtar, A. (2013). Tandem stem-loops in roX RNAs act together to mediate X chromosome dosage compensation in Drosophila. Mol Cell 51: 156-173. PubMed ID: 23870142

Jin, Y., et al. (1999). JIL-1: a novel chromosomal tandem kinase implicated in transcriptional regulation in Drosophila. Mol. Cell 4(1): 129-35. PubMed Citation: 10445035 <>Jin, Y., et al. (2000). JIL-1, a chromosomal kinase implicated in regulation of chromatin structure, associates with the male specific lethal (MSL) dosage compensation complex. J. Cell Biol. 149(5): 1005-10. PubMed Citation: 10831604

Johansson, A. M., Allgardsson, A., Stenberg, P. and Larsson, J. (2011). msl2 mRNA is bound by free nuclear MSL complex in Drosophila melanogaster. Nucleic Acids Res. 39(15): 6428-39. PubMed Citation: 21551218

Kelly, R. L., et al. (1995). Expression of msl-2 causes assembly of dosage compensation regulators on the X chromosomes and female lethality in Drosophila. Cell 81: 867-877. PubMed Citation: 7781064

Kelley, R. L., et al. (1997). Sex lethal controls dosage compensation in Drosophila by a non-splicing mechanism. Nature 387 (6629): 195-199. PubMed Citation: 9144292

Kelley, R. I. and Kuroda, M. I. (2003). The Drosophila roX1 RNA gene can overcome silent chromatin by recruiting the male-specific lethal dosage compensation complex. Genetics 164: 565-574. 12807777

Larschan, E., et al. (2007). MSL complex is attracted to genes marked by H3K36 trimethylation using a sequence-independent mechanism. Mol. Cell 28: 121-133. PubMed Citation: 17936709

Larschan, E., et al. (2011). X chromosome dosage compensation via enhanced transcriptional elongation in Drosophila. Nature 471(7336): 115-8. PubMed Citation: 21368835

Li, F., Parry, D. A. and Scott, M. J. (2005). The amino-terminal region of Drosophila MSL1 contains basic, glycine-rich, and leucine zipper-like motifs that promote X chromosome binding, self-association, and MSL2 binding, respectively. Mol. Cell. Biol. 25(20): 8913-24. 16199870

Li, F., Schiemann, A. H. and Scott, M. J. (2008). Incorporation of the noncoding roX RNAs alters the chromatin-binding specificity of the Drosophila MSL1/MSL2 complex. Mol Cell Biol 28: 1252-1264. PubMed ID: 18086881

Lim, C. K. and Kelley, R. L. (2012). Autoregulation of the Drosophila Noncoding roX1 RNA Gene. PLoS Genet 8: e1002564. PubMed ID: 22438819

Lyman, L. M., et al. (1997). Drosophila male-specific lethal-2 protein: structure/function analysis and dependence on MSL-1 for chromosome association. Genetics 147(4): 1743-1753. PubMed Citation: 9409833

Maenner, S., Muller, M., Frohlich, J., Langer, D. and Becker, P. B. (2013). ATP-dependent roX RNA remodeling by the helicase maleless enables specific association of MSL proteins. Mol Cell 51: 174-184. PubMed ID: 23870143

Marcand, S., et al. (1996). Silencing of genes at nontelomeric sites in yeast is controlled by sequestration of silencing factors at telomeres by Rap 1 protein. Genes Dev. 10:1297-1309. PubMed Citation: 8647429

Marin, I. (2003). Evolution of chromatin-remodeling complexes: comparative genomics reveals the ancient origin of 'Novel' compensasome genes. J. Mol. Evol 56: 527-539. 12698291

McDowell, K. A., Hilfiker, A. and Lucchesi, J. C. (1996). Dosage compensation in Drosophila: the X chromosome binding of MSL-1 and MSL-2 in female embryos is prevented by the early expression of the Sxl gene. Mech. Dev. 57: 113-119. PubMed Citation: 8817458

Meller, V. H., et al. (1997). roX1 RNA paints the X chromosome of male Drosophila and is regulated by the dosage compensation system. Cell 88: 445-457. PubMed Citation: 9038336

Meller, V. H., Gordadze, P. R., Park, Y., Chu, X., Stuckenholz, C., Kelley, R. L. and Kuroda, M. I. (2000). Ordered assembly of roX RNAs into MSL complexes on the dosage-compensated X chromosome in Drosophila. Curr Biol 10: 136-143. PubMed ID: 10679323

Meller, V. H. (2003). Initiation of dosage compensation in Drosophila embryos depends on expression of the roX RNAs. Mech Dev 120: 759-767. PubMed ID: 12915227

Mendjan. S., et al. (2006). Nuclear pore components are involved in the transcriptional regulation of dosage compensation in Drosophila. Mol. Cell 21: 811-823. 16543150

Merendino, L., Guth, S., Bilbao, D., Martinez, C. and Valcarcel, J. (1999). Inhibition of msl-2 splicing by Sex-lethal reveals interaction between U2AF35 and the 3' splice site AG. Nature 402(6763): 838-41. 10617208

Oh, H., Park, Y. and Kuroda, M. I. (2003). Local spreading of MSL complexes from roX genes on the Drosophila X chromosome. Genes Dev. 17: 1334-1339. 12782651

Oh, H., Bone, J. R. and Kuroda, M. I. (2004). Multiple classes of MSL binding sites target dosage compensation to the X chromosome of Drosophila. Curr. Biol. 14: 481-487. 15043812

Park, Y., Mengus, G., Bai, X., Kageyama, Y., Meller, V. H., Becker, P. B. and Kuroda, M. I. (2003). Sequence-specific targeting of Drosophila roX genes by the MSL dosage compensation complex. Mol Cell 11: 977-986. PubMed ID: 12718883

Patalano, S., Mihailovich, M., Belacortu, Y., Paricio, N. and Gebauer, F. (2009). Dual sex-specific functions of Drosophila Upstream of N-ras in the control of X chromosome dosage compensation. Development 136(4):689-98. PubMed Citation: 19168682

Park, Y., et al. (2003). Sequence-specific targeting of Drosophila roX genes by the MSL dosage compensation complex. Molec. Cell 11: 977-986. 12718883

Ramírez, F.,Lingg, T., Toscano, S., Lam, K.C., Georgiev, P., Chung, H.R., Lajoie, B.R., de Wit, E., Zhan, Y., de Laat, W., Dekker, J., Manke, T. and Akhtar, A. (2015). High-affinity sites form an interaction network to facilitate spreading of the MSL complex across the X chromosome in Drosophila. Mol Cell 60: 146-162. PubMed ID: 26431028

Rattner, B. P. and Meller, V. H. (2004). Drosophila Male-specific lethal 2 protein controls sex-specific expression of the roX genes. Genetics 166: 1825-1832. 15126401

Scott, M. J., et al. (2000). MSL1 plays a central role in assembly of the MSL complex, essential for dosage compensation in Drosophila. EMBO J. 19: 144-155. PubMed Citation: 7862667

Sobel, R. E., et al. (1995). Conservation of deposition-related acetylation sites in newly synthesized histones H3 and H4. Proc. Natl. Acad. Sci. 92: 1237-41. PubMed Citation: 7862667

Spierer, A., Seum, C., Delattre, M. and Spierer, P. (2005). Loss of the modifiers of variegation Su(var)3-7 or HP1 impacts male X polytene chromosome morphology and dosage compensation. J. Cell Sci. 118: 5047-5057. PubMed Citation: 16234327

Spierer, A., Begeot, F., Spierer, P. and Delattre, M. (2008). SU(VAR)3-7 links heterochromatin and dosage compensation in Drosophila. PLoS Genet. 4(5): e1000066. PubMed Citation: 18451980

Straub, T., et al. (2005a). Stable chromosomal association of MSL2 defines a dosage-compensated nuclear compartment. Chromosoma 114(5): 352-64. 16179989

Straub. T., Gilfillan, G. C., Maier, V. K. and Becker, P. B. (2005b). The Drosophila MSL complex activates the transcription of target genes. Genes Dev. 19: 2284-2288. 16204179

Turner, B. M., Birley, S. J., and Lavender, J. (1992). Histone H4 isoforms acetylated at specific lysine residues define individual chromosomes and chromatin domains in Drosophila polytene nuclei. Cell 69: 375-384. PubMed Citation: 1568251

Veneti, Z., Bentley, J. K., Koana, T., Braig, H. R. and Hurst, G. D. (2005). A functional dosage compensation complex required for male killing in Drosophila. Science 307(5714): 1461-3. 15746426

Villa, R., Schauer, T., Smialowski, P., Straub, T. and Becker, P. B. (2016). PionX sites mark the X chromosome for dosage compensation. Nature 537(7619): 244-248. PubMed ID: 27580037

Wu, L., et al. (2011). The RING finger protein MSL2 in the MOF complex is an E3 ubiquitin ligase for H2B K34 and is involved in crosstalk with H3 K4 and K79 methylation. Mol. Cell 43(1): 132-44. PubMed Citation: 21726816

Zheng, S., Villa, R., Wang, J., Feng, Y., Wang, J., Becker, P. B. and Ye, K. (2014). Structural basis of X chromosome DNA recognition by the MSL2 CXC domain during Drosophila dosage compensation. Genes Dev 28: 2652-2662. PubMed ID: 25452275

Zhou, S. Y., et al. (1995). Male-specific lethal 2, a dosage compensation gene of Drosophila, undergoes sex-specific regulation and encodes a protein with a RING finger and a metallothionein-like cysteine cluster. EMBO J. 14: 2884-95. PubMed Citation: 7796814


male-specific lethal 2: Biological Overview | Evolutionary Homologs | Regulation | Protein Interactions | Developmental Biology | Effects of Mutation

date revised: 22 December 2017
 

Home page: The Interactive Fly © 1997 Thomas B. Brody, Ph.D.

The Interactive Fly resides on the
Society for Developmental Biology's Web server.