Post-meiotic transcription was accepted to be essentially absent from Drosophila spermatogenesis. We identify 24 Drosophila genes whose mRNAs are most abundant in elongating spermatids. By single-cyst quantitative RT-PCR, we demonstrate post-meiotic transcription of these genes. We conclude that transcription stops in Drosophila late primary spermatocytes, then is reactivated by two pathways for a few loci just before histone-to-transition protein-to-protamine chromatin remodelling in spermiogenesis. These mRNAs localise to a small region at the distal elongating end of the spermatid bundles, thus they represent a new class of sub-cellularly localised mRNAs. Mutants for a post-meiotically transcribed gene (scotti), are male sterile, and show spermatid individualisation defects, indicating a function in late spermiogenesis.

In D. melanogaster spermatogenesis (reviewed by Fuller, 1993), the primary spermatogonium undergoes four mitotic divisions, each with incomplete cytokinesis. Two somatic cyst cells, analogous to mammalian Sertoli cells,encapsulate each spermatogonium, and thus, later, 16 spermatocytes, and, after meiosis, 64 inter-connected spermatids. After meiosis, morphological changes,including axoneme assembly and mitochondrial and membrane remodelling,transform round spermatids into mature 2-mm-long motile sperm. These post-meiotic events are believed to be driven by translation, as spermiogenesis genes are typically transcribed in primary spermatocytes, the mRNAs are then stored in spermatids and translated during elongation (reviewed by Schäfer et al.,1995).

In mammalian early spermatids, transcription is readily detected (by 3H-Uridine incorporation) and continues until chromatin compaction(Kierszenbaum and Tres, 1975; Monesi, 1965). Mammalian spermatocytes transcribe genes required during meiosis or in early spermatids,whereas mRNAs for spermiogenesis proteins are transcribed post-meiotically(Schultz et al., 2003), with some under additional translational control (reviewed by Braun, 1998). Radiographic studies showed no detectable transcription in Drosophila spermatids(Gould-Somero and Holland,1974; Olivieri and Olivieri,1965); however, RNA polymerase II activity in Drosophilaspermatid nuclei suggests that these cells actively transcribe genes(Rathke et al., 2007).

Spermatid elongation generates dramatic cellular asymmetry, with nuclei at one end, and axonemes extending towards the distal end, where sister cells are connected via ring canals (Hime et al.,1996). During individualisation, each spermatid is stripped of excess cytoplasm and organelles, and is separated from sister spermatids by an individualisation complex that progresses from the heads along to the tails, a process dependent on apoptotic pathway activation(Arama et al., 2003). Spermatids have a unique, highly compact, non-nucleosomal chromatin organisation. Histones are removed from spermatid chromosomes and replaced by other small basic proteins, initially transition proteins, then protamines or protamine-like proteins (reviewed by Braun,2001; Oliva,2006). Drosophila have at least one transition protein and several protamine-like proteins(Jayaramaiah Raja and Renkawitz-Pohl,2005; Rathke et al.,2007).

Here, we provide compelling evidence for the transcription of 24 genes in Drosophila spermatids, activated by at least two regulatory pathways. In Drosophila, transcription shuts off in late primary spermatocytes,then (for a few loci) is reactivated mid-elongation, just before histone-to-protamine chromatin remodelling. At least one post-meiotically expressed gene, scotti, is required for normal actin cone progression during spermatid individualisation, and thus for male fertility. mRNAs encoded by all of the post-meiotically transcribed genes are localised to the extreme distal ends of the elongating cells; thus, they represent a novel class of sub-cellularly localised mRNAs.

Drosophila strains and culture

Flies were raised on standard cornmeal sucrose agar at 25°C. Visible markers and balancer chromosomes are described in FlyBase(Crosby et al., 2007). Wild type was w1118. Protamine-EGFP flies were from Renate Renkawitz-Pohl (Jayaramaiah Raja and Renkawitz-Pohl, 2005) and H2A-mRFP1 flies from Jürgen Knoblich (IMBA, Vienna, Austria).

RNA in situ hybridisation

RT-PCR products (400-600 bp) were generated from total testis RNA. 3′PCR primers included a T3 RNA polymerase promoter for the in vitro transcription of digoxigenin (DIG)-labelled RNA probes. In situ hybridisation was as described previously (White-Cooper et al., 1998). Primer sequences are available on request.

Isolation of RNA from individual cysts or bundles

Testes were dissected from w1118 or the young male progeny of w; Mst35Ba-egfp × w; H2A-mRFP1, in testis buffer (183 mM KCl, 47 mM NaCl, 10 mM Tris pH 6.9), transferred onto a siliconised slide, opened with tungsten needles and examined using an Olympus BX50 upright microscope with a long working distance 10× objective. Individual photographed cysts were transferred into 100 μl lysis buffer using a pulled-out Pasteur pipette; total RNA was extracted according to manufacturer's instructions (RNAqueous-Micro kit, Ambion, Warrington, UK).

Q-RT-PCR

First-strand cDNA was synthesised using oligo(dT)20 and SuperScript III Reverse Transcriptase (Invitrogen, Paisley, UK) in a 20 μl reaction. Each PCR contained 0.33 μl cDNA, 10 μl 2× Power SYBR Green PCR Master Mix (Applied Biosystems, Warrington, UK), and 100 nM of each gene-specific primer, in a 20 μl final volume. Primers amplified cDNA only,as one of each pair spanned an exon-exon junction. Negative controls (testis RNA without reverse transcriptase) were performed for every primer pair in every PCR plate. Q-PCR reactions in 96-well thin-wall white plates(BIOplastics, Braintree, UK) were run in a Chromo4 with Opticon Monitor Software (GRI, Braintree, UK). Q-PCRs were run in triplicate and the internal reference control (CG10252) was quantified for each sample.

Fig. 1.

RNA in situ hybridisation reveals comet and cup transcript patterns.(A) Drosophila testis viewed by phase contrast. Asterisk,apical tip; bracket, spermatocytes; arrowheads, early spermatids; arrows,elongating spermatid bundles. (B-D) Control genes. (E,F)Typical cup and comet localisation. Scale bar: 100 μm for to A-F.(G-J) Typical cup localisation at spermatid distal ends. Scale bar: 25μm. (K) Diagram of a spermatid bundle showing the location of comet and cup transcripts relative to spermatid nuclei, and cyst cells (grey).

Fig. 1.

RNA in situ hybridisation reveals comet and cup transcript patterns.(A) Drosophila testis viewed by phase contrast. Asterisk,apical tip; bracket, spermatocytes; arrowheads, early spermatids; arrows,elongating spermatid bundles. (B-D) Control genes. (E,F)Typical cup and comet localisation. Scale bar: 100 μm for to A-F.(G-J) Typical cup localisation at spermatid distal ends. Scale bar: 25μm. (K) Diagram of a spermatid bundle showing the location of comet and cup transcripts relative to spermatid nuclei, and cyst cells (grey).

soti mutant generation

soti deletions were generated by FLP-mediated recombination between PBac[WH]f02605 inserted 2162 bp 3′ of soti and P[XP]d01837 inserted 468 bp 5′ of soti(Parks et al., 2004; Thibault et al., 2004). All five independent deletion lines were confirmed by soti ORF PCR, and were viable but male sterile.

FITC-Phalloidin staining of testes

Visualisation of actin cones in mutant and wild-type testes was carried out as described previously (White-Cooper,2004).

Drosophila `comet' and `cup' transcripts are specifically detected in spermatid bundles

Many genes with unknown functions have testes-specific expression. To determine when during spermatogenesis these proteins are made, we examined the transcript patterns of >1200 genes by in situ hybridisation(www.fly-ted.org). The expression of spermiogenesis genes in primary spermatocytes, and the storage of transcripts for later use during spermiogenesis, means that the translation of specific mRNAs in Drosophila spermatids correlates well with their disappearance, as translation exposes stored mRNAs to the RNA degradation machinery (Schäfer et al., 1995). The in situ hybridisation results will be described in detail elsewhere; in summary, 529 of the 553 mRNAs detected in spermatids were transcribed in primary spermatocytes, persisted in the spermatid cytoplasm,and were degraded at various stages in elongation(Fig. 1C,D). Unexpectedly, we found 24 germ-line expressed genes that did not conform to this pattern(Table 1; see also Table S1 in the supplementary material). We subdivided these on the basis of subtle differences in transcript localisation patterns, and refer to the genes collectively as `comets and cups'.

Table 1.

Comet and cup gene names, symbols and CG identifiers

GeneSymbolCG numberPCR
Comet genes
 

 

 

 
schumacher-levy schuy CG17736 Yes* 
hale-bopp hale CG7570 Yes* 
sungrazer sunz CG15179 Yes 
solwind sowi CG15178 Yes 
borrelly boly CG30362 Yes* 
comas sola cola CG30363 Yes* 
hug-bell hubl CG30364  
spacewatch spaw CG30365  
whipple whip CG34218  
swift-tuttle swif CG30366 Yes* 
scotti soti CG8489 Yes 
Pglym87
 

 
CG17645
 

 
Cup genes
 

 

 

 
calcutta-cup c-cup CG15623 Yes 
walker-cup wa-cup CG10113 Yes 
ryder-cup r-cup CG10998 Yes 
davis-cup d-cup CG14387 Yes 
presidents-cup p-cup CG12993 Yes 
world-cup w-cup CG7363  
stanley-cup s-cup CG30044  
tetleys-cup t-cup CG31858  
flyers-cup f-cup CG9611  
heineken-cup h-cup CG6130  
mann-cup m-cup CG11896  
oo18 RNA-binding protein orb CG10868  
GeneSymbolCG numberPCR
Comet genes
 

 

 

 
schumacher-levy schuy CG17736 Yes* 
hale-bopp hale CG7570 Yes* 
sungrazer sunz CG15179 Yes 
solwind sowi CG15178 Yes 
borrelly boly CG30362 Yes* 
comas sola cola CG30363 Yes* 
hug-bell hubl CG30364  
spacewatch spaw CG30365  
whipple whip CG34218  
swift-tuttle swif CG30366 Yes* 
scotti soti CG8489 Yes 
Pglym87
 

 
CG17645
 

 
Cup genes
 

 

 

 
calcutta-cup c-cup CG15623 Yes 
walker-cup wa-cup CG10113 Yes 
ryder-cup r-cup CG10998 Yes 
davis-cup d-cup CG14387 Yes 
presidents-cup p-cup CG12993 Yes 
world-cup w-cup CG7363  
stanley-cup s-cup CG30044  
tetleys-cup t-cup CG31858  
flyers-cup f-cup CG9611  
heineken-cup h-cup CG6130  
mann-cup m-cup CG11896  
oo18 RNA-binding protein orb CG10868  
*

Genes resembling hale in Q-RT-PCR transcript profiling.

Genes resembling sunz in Q-RT-PCR transcript profiling.

Comet and cup transcripts were detected at very low levels in primary spermatocytes by RNA in situ hybridisation(Fig. 1E,F; see Fig. S1 in the supplementary material), and were barely detected in early elongation spermatids. However, robust signals, with striking subcellular localisations,were evident in more elongated spermatids. Spermatid nuclei are located at one end of these elongated cells, in the basal-most region of the testis; comet and cup mRNAs were localised to the distal ends of the spermatids, in subtly different patterns (Fig. 1G-K).`Comet' mRNAs localised into a ball shape at the ends of spermatid bundles,trailing away proximally to a less abundant, speckled distribution(Fig. 1H,J). `Cup' transcripts localised in shallow cup-like shapes at the ends of spermatid bundles(Fig. 1G,I).

Fig. 2.

Q-RT-PCR confirms post-meiotic transcription. (A) Individual cyst Q-RT-PCR assay schematic. (B) Cysts used for RNA extraction, from primary spermatocytes (I1, I2) and round spermatids (R1, R2; scale bars: 25μm), and from elongating spermatids (0.19-1.89 mm; scale bars: 100 μm).(C,D) Control Q-RT-PCRs. (E-H) Peak of expression in spermatids confirms post-meiotic transcription of comet and cup genes. hale (E) and schuy (F) were expressed in more bundles than were d-cup (G) and sunz (H).

Fig. 2.

Q-RT-PCR confirms post-meiotic transcription. (A) Individual cyst Q-RT-PCR assay schematic. (B) Cysts used for RNA extraction, from primary spermatocytes (I1, I2) and round spermatids (R1, R2; scale bars: 25μm), and from elongating spermatids (0.19-1.89 mm; scale bars: 100 μm).(C,D) Control Q-RT-PCRs. (E-H) Peak of expression in spermatids confirms post-meiotic transcription of comet and cup genes. hale (E) and schuy (F) were expressed in more bundles than were d-cup (G) and sunz (H).

Quantitative RT-PCR confirms post meiotic transcription

Comet and cup expression patterns are extremely unusual. The obvious explanation for the abrupt mRNA appearance during spermatid differentiation is post-meiotic transcription. Alternatively, the transcripts could be present earlier, but either (1) diffuse or (2) masked, so undetectable by in situ hybridisation.

To verify the post-meiotic transcription, and to determine its timing with respect to cellular differentiation events, we developed a single-cyst quantitative reverse transcription PCR (Q-RT-PCR) protocol. Testes were dissected, and individual cysts isolated, photographed, and staged according to morphology; total RNA was then isolated and first strand cDNA synthesised(Fig. 2A). Each cyst yielded cDNA for 60 Q-RT-PCR reactions. Testis-specific control genes were chosen. The CG10252 transcript conforms to the conventional pattern for a late elongation protein (Fig. 1D)and CG10252 protein is detected in mature sperm(Dorus et al., 2006). CG3927 was detected exclusively in primary spermatocytes; CG11591 was expressed in primary spermatocytes and the signal disappeared from mid-elongation spermatids(Fig. 1B,C). CG3927and CG11591 controlled for cyst visual staging. For each cyst,expression levels of both staging controls and up to eight test genes were compared with the internal control CG10252. Fig. 2 shows a typical experiment, expression levels were normalised to primary spermatocyte cyst I1. Consistent with the in situ hybridisation patterns, CG3927 transcript was only detected in primary spermatocytes. CG11591 transcript was highest in primary spermatocytes, and persisted into early elongation stage spermatids. The 13 comet and cup gene transcripts assayed by isolated-cyst Q-RT-PCR showed broadly similar profiles in Q-RT-PCR assays (Figs 2, 3; see also Figs S2, S5, S6 in the supplementary material). All transcripts were detected in primary spermatocytes and round spermatids. sunz, sowi, soti, c-cup, d-cup,wa-cup, p-cup and r-cup were low or not detected in very short elongating cysts, but were detected at high levels in a few longer spermatid cysts. hale, schuy, boly, cola and swif were detected at a basal level in almost all cysts, but were much more abundant in a few mid-elongation bundles. From these differences, we infer two separate regulatory modules activating post-meiotic gene expression, with the hale group being transcribed in more cysts than the sunzgroup. Spermatid length measurements give good staging of the relative differentiation states of cysts from a single testis, but the exact length of spermatids expressing comets and cups varied between testes. The initial low-level signal in primary spermatocytes, the dip in signal intensity in early spermatids, then the dramatic appearance in later spermatids conclusively demonstrate that there is post-meiotic transcription in Drosophila testes.

Fig. 3.

Comet and cup transcription precedes the histone-protamine switch.Bars above graphs and colours within graphs indicate H2A-mRFP (histone, red)and Mst35Ba-EGFP (protamine, green; overlap orange/yellow) detection.(A-C) Control Q-RT-PCRs. (D-J) Comet and cup Q-RT-PCR expression profiles show that expression initiates in cysts negative for protamine.(K) Nuclear fluorescence in elongating cysts.

Fig. 3.

Comet and cup transcription precedes the histone-protamine switch.Bars above graphs and colours within graphs indicate H2A-mRFP (histone, red)and Mst35Ba-EGFP (protamine, green; overlap orange/yellow) detection.(A-C) Control Q-RT-PCRs. (D-J) Comet and cup Q-RT-PCR expression profiles show that expression initiates in cysts negative for protamine.(K) Nuclear fluorescence in elongating cysts.

Fig. 4.

soti is required for normal individualisation complex progression. Confocal maximum brightness projections of the entire cyst depth. DNA is labelled with propidium iodide (A-D, magenta); actin is labelled with FITC-phalloidin (A′-D′, green). Scale bar: 10 μm.(A-B″) In wild-type testes, investment cones assembled above spermatid nuclei (A-A″) and moved along spermatid bundles as a complex(B-B″). (C-D″) In soti mutant testes, investment cones assembled normally, but complexes dissociated during individualisation.

Fig. 4.

soti is required for normal individualisation complex progression. Confocal maximum brightness projections of the entire cyst depth. DNA is labelled with propidium iodide (A-D, magenta); actin is labelled with FITC-phalloidin (A′-D′, green). Scale bar: 10 μm.(A-B″) In wild-type testes, investment cones assembled above spermatid nuclei (A-A″) and moved along spermatid bundles as a complex(B-B″). (C-D″) In soti mutant testes, investment cones assembled normally, but complexes dissociated during individualisation.

Comet and cup genes are expressed before most of the DNA has been loaded with transition proteins or protamines

In Drosophila, bulk histone removal initiates in the early `canoe'stage of nuclear remodelling, and protamine deposition is complete by late canoe stage (Jayaramaiah Raja and Renkawitz-Pohl, 2005). To determine comet and cup transcription timing with respect to chromatin reorganisation, we staged cysts via combined fluorescent fusion-protein localisation and spermatid-length measurements. We isolated cysts from flies co-expressing Mst35Ba-GFP (protamine-GFP) and H2A-mRFP1 (histone-GFP). Protamine accumulation initiates before all histones have been removed, as some nuclei fluoresced both red and green(Fig. 3K). Transcription of comet and cup genes was detected in mid-elongation cysts that were positive for histone and negative for protamine(Fig. 3D-J). Thus, comet and cup transcription occurs just before the deposition of protamines. Some comet and cup mRNAs were also detected in cysts positive for protamine-GFP. This could be due to ongoing transcription, or to message stability. The recently described active transcription in spermatids(Rathke et al., 2007)coincides with the comet and cup gene transcription peak.

We repeated these experiments using cysts isolated from flies co-expressing Tpl94D-GFP (transition protein) and H2A-mRFP1. Initiation of post-meiotic comet and cup gene expression was found in cysts lacking nuclear Tpl94D (see Fig. S2B in the supplementary material), indicating that comet and cup gene transcription initiates before the deposition of transition protein, while chromatin is presumably still nucleosomal.

Rfx is not required for expression of the comet and cup genes

Rfx is a winged-helix transcription factor important for ciliogenesis gene expression. In testes, Rfx protein was detected transiently in canoe-stage spermatid nuclei only (Vandaele et al.,2001), and thus was an excellent comet and cup candidate transcription factor. All comet and cup genes tested (hale, schuy, cola,soti, c-cup, w-cup, p-cup and wa-cup) by in situ hybridisation to Rfx253/Rfx49(Dubruille et al., 2002) mutant testes were indistinguishable from wild type (see Fig. S3 in the supplementary material). Rfx49 is a null, whereas Rfx253 lacks DNA-binding activity. Thus, despite the intriguing localisation of Rfx to spermatid nuclei, its X-box binding function is not required for comet or cup expression.

scotti is required for spermatid individualisation

We deleted approximately 4 kb of genomic DNA, including the entire scotti (soti, a comet) ORF, by FLP-mediated recombination between flanking FRT-containing transposons. soti homozygous mutants were viable and female fertile, but male sterile. Phase contrast microscopy indicated no gross defects in soti testes organisation or spermatid elongation; however, empty seminal vesicles indicated spermiogenesis defects. Within each individualising spermatid cyst, 64 actin-rich investment cones move together as an individualisation complex, pushing ahead a cystic bulge of excess cytoplasm and organelles. This cytoplasm is discarded from spermatid distal ends as a waste bag. Waste bags were completely absent from mutant testes, and cystic bulges were rarely seen. FITC-phalloidin labelling revealed that investment cones formed normally in soti mutant males; however,nuclei failed to remain tightly clustered and were displaced distally along the cyst (Fig. 4). Although investment cones progressed away from the nuclei in mutants, investment cone coupling within individualisation complexes was lost, and cones never progressed the full length of mutant spermatids. Thus, soti function is required for spermatid individualisation.

Why are the comets and cups transcribed in spermatids?

Post-meiotic transcription, in early spermatids, has been reported for two loci in Drosophila, hsr-omega and Hsp70(Bendena et al., 1991);however, we have been unable to reproduce these findings (see Fig. S5 in the supplementary material). Ninety-six percent of genes whose mRNAs were detected in spermatids are not actively transcribed in these cells (being made in spermatocytes), so what is special about the exceptional 4% - the comets and cups? These genes are found throughout the euchromatin, including the X chromosome, and their local genomic environments showed no unusual features. Their flanking genes showed no bias towards or away from testis-specific expression in adults (Chintapalli et al.,2007). There are three comet and cup gene clusters, two of which clearly represent gene duplication events (see Table S2 in the supplementary material). The final cluster comprises hale and schuy; both encode glutamine-rich proteins, but their evolutionary history is unclear. We investigated the expression of all 10 related genes in the CG11635-CG8701 cluster (see Fig. S5A in the supplementary material). CG11635, CG18449,CG2127 and CG8701 were expressed in the conventional spermiogenesis gene pattern - transcribed in primary spermatocytes and stored until late elongation - while spaw, hubl, swif, cola, boly and whip were typical `comets' (see Tables S1, S2 and Fig. S5B,C in the supplementary material). Q-RT-PCR confirmed that the post-meiotic transcription and RNA localisation to distal ends of spermatids were correlated.

In mammals, the transcription of many genes in spermatids has been described, and new reports are frequent(Reynard et al., 2007; Schultz et al., 2003). These mammalian genes typically, although not exclusively, encode components of the mature sperm. By contrast, Drosophila comet and cup proteins, with the exception of Boly and Pglym87, are not sperm components (see Table S1 in the supplementary material) (Dorus et al.,2006). Perhaps comet and cup proteins function, like Soti, during spermiogenesis, rather than in sperm. Alternatively, perhaps they are present in sperm but at a very low copy number. mRNAs of several comet and cup gene homologues were transcribed in conventional patterns, and the encoded proteins detected in sperm (see Table S2 in the supplementary material). orb(a comet) encodes an RNA-binding protein, potentially anchoring other comet and cup mRNAs. The other comet or cup proteins have no predicted function. PKD2 encodes a Ca2+-activated non-selective cation channel, and it is intriguing that a Drosophila PKD2 homologue (Pkd2/Amo)concentrates at the distal ends of sperm, and is important for sperm function(Watnick et al., 2003). Sunz,Sowi and D-cup are EF-hand-containing proteins, and so could function with Pkd2 in mediating a Ca2+ signal at the spermatid tail tip. This signal could then be transduced along spermatid tails, perhaps via the mitochondrial derivative or the endoplasmic reticulum-derived axonemal sheath,which stretches the length of spermatid tails, activating the apoptotic pathway to synchronise individualisation and ensure normal investment cone progression.

In conclusion, there is significant transcription from several genomic loci in Drosophila spermatids, and the post-meiotically expressed transcripts localise to the growing ends of spermatids. This transcription and RNA localisation occurs before spermatid chromatin remodelling, and at least one post-meiotically-expressed gene is required for spermiogenesis.

We thank Bénédicte Durand for Rfx mutant flies,Renate Renkawitz-Pohl for protamine and Tpl94D-EGFP flies, Spyros Artavanis-Tsakonas for Exelixis stocks and Jennifer Mummery-Widmer (Knoblich lab) for H2A-mRFP1 flies. This work was funded by the BBSRC and the Royal Society.

Arama, E., Agapite, J. and Steller, H. (
2003
). Caspase activity and a specific cytochrome C are required for sperm differentiation in Drosophila.
Dev. Cell
4
,
687
-697.
Bendena, W. G., Ayme, S. A., Garbe, J. C. and Pardue, M. L.(
1991
). Expression of heat-shock locus hsr-omega in nonstressed cells during development in Drosophila melanogaster.
Dev. Biol.
144
,
65
-77.
Braun, R. (
1998
). Post-transcriptional control of gene expression during spermatogenesis.
Semin. Cell Dev. Biol.
9
,
483
-489.
Braun, R. (
2001
). Packaging paternal chromosomes with protamine.
Nat. Genet.
28
,
10
-12.
Chintapalli, V., Wang, J. and Dow, J. (
2007
). Using FlyAtlas to identify better Drosophila models of human disease.
Nat. Genet.
39
,
715
-720.
Crosby, M., Goodman, J., Strelets, V., Zhang, P., Gelbart, W. and Consortium, F. (
2007
). FlyBase: genomes by the dozen.
Nucleic Acids Res.
35
,
D486
-D491.
Dorus, S., Busby, S. A., Gerike, U., Shabanowitz, J., Hunt, D. F. and Karr, T. L. (
2006
). Genomic and functional evolution of the Drosophila melanogaster sperm proteome.
Nat. Genet.
38
,
1440
-1445.
Dubruille, R., Laurecon, A., Vandaele, C., Shishido, E.,Coulon-Bublex, M., Swoboda, P., Couble, P., Kernan, M. and Durand, B.(
2002
). Drosophila regulatory factor X is necessary for ciliated sensory neuron differentiation.
Development
129
,
5487
-5498.
Fuller, M. T. (
1993
). Spermatogenesis. In
The Development of Drosophila
. Vol.
1
(ed. M. Bate and A. Martinez-Arias), pp.
71
-147. Cold Spring Harbor, NY: Cold Spring Harbor Press.
Gould-Somero, M. and Holland, L. (
1974
). The timing of RNA synthesis for spermiogenesis in organ cultures of Drosophila melanogaster testes.
Wilhelm Rouxs Arch.
174
,
133
-148.
Hime, G. R., Brill, J. A. and Fuller, M. T.(
1996
). Assembly of ring canals in the male germ line from structural components of the contractile ring.
J. Cell Sci.
109
,
2779
-2788.
Jayaramaiah Raja, S. and Renkawitz-Pohl, R.(
2005
). Replacement by Drosophila melanogaster protamines and Mst77F of histones during chromatin condensation in late spermatids and role of sesame in the removal of these proteins from the male pronucleus.
Mol. Cell. Biol.
25
,
6165
-6177.
Kierszenbaum, A. and Tres, L. (
1975
). Structural and transcriptional features of the mouse spermatid genome.
J. Cell Biol.
65
,
258
-270.
Monesi, V. (
1965
). Synthetic activities during spermatogenesis in the mouse.
Exp. Cell Res.
39
,
197
-224.
Oliva, R. (
2006
). Protamines and male infertility.
Hum. Reprod. Update
12
,
417
-435.
Olivieri, G. and Olivieri, A. (
1965
). Autoradiographic study of nucleic acid synthesis during spermatogenesis in
Drosophila melanogaster. Mutat. Res.
2
,
366
-380.
Parks, A. L., Cook, K. R., Belvin, M., Dompe, N. A., Fawcett,R., Huppert, K., Tan, L. R., Winter, C. G., Bogart, K. P., Deal, J. E. et al. (
2004
). Systematic generation of high-resolution deletion coverage of the Drosophila melanogaster genome.
Nat. Genet.
36
,
288
-292.
Rathke, C., Barrends, W. M., Jayaramaiah Raja, S., Bartkuhn, M.,Renkawitz, R. and Renkawitz-Pohl, R. (
2007
). Transition from a nucleosome-based to a protamin-based chromatin configuration during spermiogenesis in
Drosophila. J. Cell Sci.
120
,
1689
-1700.
Reynard, L., Turner, J., Cocquet, J., Mahadevaiah, S., Toure,A., Hoog, C. and Burgoyne, P. (
2007
). Expression analysis of the mouse multi-copy X-linked gene Xlr-Related, meiosis-regulated (Xmr),reveals that Xmr encodes a spermatid-expressed cytoplasmic protein SLX/XMR.
Biol. Reprod.
77
,
329
-335.
Schäfer, M., Nayernia, K., Engel, W. and Schäfer,U. (
1995
). Translational control in spermatogenesis.
Dev. Biol.
172
,
344
-352.
Schultz, N., Hamra, F. and Garbers, D. (
2003
). A multitude of genes expressed solely in meiotic or postmeiotic spermatogenic cells offers a myriad of contraceptive targets.
Proc. Natl. Acad. Sci. USA
100
,
12201
-12206.
Thibault, S. T., Singer, M. A., Miyazaki, W. Y., Milash, B.,Dompe, N. A., Singh, C. M., Buchholz, R., Demsky, M., Fawcett, R.,Francis-Lang, H. L. et al. (
2004
). A complementary transposon tool kit for Drosophila melanogaster using P and piggyBac.
Nat. Genet.
36
,
283
-287.
Vandaele, C., Coulon-Bublex, M., Couble, P. and Durand, B.(
2001
). Drosophila regulatory factor X is an embryonic type I sensory neuron marker also expressed in spermatids and in the brain of
Drosophila. Mech. Dev.
103
,
159
-162.
Watnick, T. J., Jin, Y., Matunis, E., Kernan, M. and Montell,C. (
2003
). A flagellar Polycystin-2 homolog required for male fertility in
Drosophila. Curr. Biol.
13
,
2175
-2178.
White-Cooper, H. (
2004
). Spermatogenesis:analysis of meiosis and morphogenesis. In
Drosophila Cytogenetics Protocols.
Vol.
247
(ed. D. Henderson),pp.
45
-75.Totowa, NJ: Humana Press.
White-Cooper, H., Schafer, M. A., Alphey, L. S. and Fuller, M. T. (
1998
). Transcriptional and post-transcriptional control mechanisms coordinate the onset of spermatid differentiation with meiosis I in
Drosophila. Development
125
,
125
-134.

Supplementary information