Main

During meiotic recombination in mice, DSBs must occur within the tiny (around 700 kb3,4) PAR2,3,4,5,6. Because on average one DSB forms per ten megabases, the PAR would be at risk of frequent recombination failure if it behaved similarly to a typical autosomal segment2. Consequently, DSBs and recombination are disproportionately frequent in the PAR2,6,7,8 (Supplementary Discussion). The mechanisms that promote these frequent DSBs are—in any species—not known.

DSBs arise concomitantly with linear axial structures that anchor the chromatin loops in which DSBs occur9,10. Axes begin to form during replication and become assembly sites for proteins that promote SPO11-mediated DSBs11,12,13. PAR chromatin in spermatocytes forms relatively short loops on a long axis2. However, only a low-resolution view of the structure of the PAR is presently available, and the controlling cis- and trans-acting factors are unknown. Moreover, it was unclear how spermatocytes—but not oocytes—make the PAR prone to such a high rate of recombination.

A distinctive PAR ultrastructure

The X and Y chromosomes usually pair late, with PARs paired in less than 20% of spermatocytes at late zygonema, when most autosomes are paired2,14. At this stage, unsynapsed PAR axes (defined by SYCP2 and SYCP3 immunostaining) appeared thickened relative to other unsynapsed axes and showed bright staining for HORMAD1 and HORMAD215 (Fig. 1a, Extended Data Fig. 1a, b). Moreover, the PAR was highly enriched for REC114, MEI4, MEI1 and IHO1—which are required for genome-wide DSB formation16,17,18,19—as well as ANKRD31, a partner of REC114 that is essential for DSBs in the PAR20,21.

Fig. 1: Ultrastructure of the PAR during male meiosis.
figure 1

a, Axis thickening (SYCP2 and SYCP3) and ANKRD31 accumulation on X and Y chromosome PARs (arrowheads) in late zygonema. The asterisk shows an autosomal ANKRD31 blob. Scale bars, 2 μm. b, Ultrastructure of the PAR before and after synapsis (montage of representative SIM images). Dashed lines indicate where chromosomes are cropped. c, Enrichment of ANKRD31 protein along split PAR axes in late zygonema (SIM image). d, Schematic showing the dynamic remodelling of the PAR loop–axis ensemble during prophase I. See measurements in Extended Data Fig. 3b. Scale bars, 1 μm (bd).

All five proteins (hereafter referred to as RMMAI) colocalized in several bright ‘blobs’ for most of prophase I (Fig. 1a, Extended Data Fig. 1c). Two blobs were on the PARs of X and Y chromosomes and others highlighted the ends of specific autosomes (Fig. 1a, Extended Data Fig. 1d). Similar blobs in previously published micrographs were not characterized16,17,19,22. The proteins also colocalized in smaller foci along unsynapsed axes16,17,19,20,21,22 (Extended Data Fig. 1c). Enrichment of RMMAI proteins at the PAR was already detectable at the pre-leptotene stage17,22 (Extended Data Fig. 1e), but not in spermatogonia (Extended Data Fig. 1f). Mass spectrometry of immunoprecipitates from testes also identified ZMYM3 and PTIP as ANKRD31-interacting proteins that were enriched at the PAR (Extended Data Fig. 1g–i, Supplementary Discussion).

Structured illumination microscopy (SIM) resolved the thickened PAR as two axial cores (Fig. 1b, Extended Data Fig. 2a, b) decorated with RMMAI proteins (Fig. 1c). The PAR axes were extended and separated in late zygonema before the synapsis of X and Y chromosomes, then collapsed during X–Y chromosome synapsis in early pachynema (Fig. 1b). Each axial core is a sister chromatid, with a ‘bubble’ from near the PAR boundary almost to the telomere (Extended Data Fig. 2c–h). This PAR structure is distinct from that seen at all other chromosome ends later in prophase I (Supplementary Discussion). Axis splitting and REC114 enrichment occurred independently of the formation of DSBs (Extended Data Fig. 2i).

Dynamic remodelling of the PAR structure

We investigated temporal patterns of axis differentiation, RMMAI composition and chromatin-loop configuration at the PAR using SIM and conventional microscopy (Fig. 1d, Extended Data Fig. 3a, b). The SYCP3-defined axis was long as soon as it became detectable in leptonema, and the fluorescence in situ hybridization (FISH) signal for the PAR boundary (PARb) was compact and remained so while the axis lengthened further through late zygonema, when the sister axes separated. Throughout, abundant ANKRD31 and REC114 signals stretched along the PAR axes, decorating the compact chromatin (Extended Data Fig. 3a, b). After synapsis, the axes shortened and the chromatin loops decompacted, with a concomitant dissociation of RMMAI proteins. A focus of the meiotic cohesin subunit REC8 was juxtaposed with ANKRD31 blobs at the pre-leptotene stage; REC8 was mostly restricted to the borders of the PAR as its axes elongated and split, and remained highly enriched on the short axis after RMMAI proteins disappeared (Extended Data Fig. 3a, b). Collapse of the loop–axis structure and dissociation of REC114 also occurred when the PAR underwent non-homologous synapsis in a Spo11−/− mutant (Extended Data Fig. 3c), showing that synapsis without recombination is sufficient for the reconfiguration of the PAR. DSB formation without synapsis may also be sufficient (Supplementary Discussion). These findings reveal a large-scale reconfiguration of the loop–axis structure and establish spatial and temporal correlations between the accumulation of RMMAI proteins at the PAR and the formation of a long axis associated with compact chromatin.

Heterochromatic mo-2 minisatellites

We deduced that specific DNA sequences might recruit RMMAI proteins because autosomal blobs also hybridized to the PARb probe (Extended Data Fig. 1d). This repetitive probe includes a tandem array of around 20 kb of a minisatellite (mo-2), with a 31-bp repeat23,24 (Fig. 2a). Clusters of mo-2 are also present at the non-centromeric ends of chromosomes 4, 9 and 1323,24 (Fig. 2a, b, Extended Data Fig. 4a, b). FISH with an mo-2 oligonucleotide probe showed that RMMAI blobs colocalize completely with mo-2 arrays (Fig. 2b, Extended Data Fig. 4c, d). Mo-2 arrays become enriched at the onset of meiosis for heterochromatic histone modifications (histone H3 trimethylated at K9 (H3K9me3) and H4K20me3) and proteins (heterochromatin protein 1β (HP1β), HP1γ and others), independent of DSB formation (Extended Data Fig. 5).

Fig. 2: Arrays of the mo-2 minisatellite are sites of enrichment of RMMAI proteins in the PAR and on autosomes.
figure 2

a, Left, self-alignment of the PARb FISH probe. The circled block is a 20-kb mo-2 cluster. Right, schematic showing the non-centromeric chromosome ends identified using a Basic Local Alignment Search Tool (BLAST) search with the mo-2 consensus sequence. b, Colocalization of REC114 blobs with the mo-2 oligonucleotide FISH signal (zygotene spermatocyte). Scale bar, 2 μm. c, Enrichment of ANKRD31 and RPA2 at the PAR correlates with mo-2 copy number. Top, late-zygotene spermatocyte from an F1 hybrid from crosses of B6 × MSM. Scale bar, 1 μm. Bottom, PAR-associated signals (AU, arbitrary units) on B6-derived (XB6) and MSM-derived (YMSM) chromosomes from the indicated number of spermatocytes (n). Red lines indicate mean ± s.d. Differences between X and Y chromosome PAR intensities are significant both for proteins and for mo-2 FISH (P < 10−6, paired t-test; for exact P values, see the associated Source Data).

Source Data

To test whether mo-2 arrays are cis-acting determinants of RMMAI recruitment, we exploited the fact that Mus musculus molossinus has an mo-2 copy number substantially lower than that of other subspecies of Mus musculus24. Mice of the MSM/MsJ (MSM) strain showed a lower hybridization signal with the mo-2 FISH probe than did mice of the C57BL/6J (B6) strain, as well as a lower REC114 intensity in blobs (Extended Data Fig. 4e).

To avoid confounding strain effects, we examined spermatocytes of F1 hybrid mice (Fig. 2c, Extended Data Fig. 4f, g). Less ANKRD31 accumulated on the MSM PARs: in offspring from B6 mothers and MSM fathers, the YMSM PAR had eightfold less ANKRD31 than the XB6 PAR (Fig. 2c, Extended Data Fig. 4g), and in the reciprocal cross the XMSM PAR had 6.5-fold less ANKRD31 than the YB6 PAR (Extended Data Fig. 4f, g). The relative levels of ANKRD31 matched those that were observed in mo-2 FISH. Nevertheless, MSM PARs supported a similar efficiency and timing of sex-chromosome pairing to the B6 PARs (Extended Data Fig. 4h)—which is not surprising, as MSM mice are fertile. Notably, the single-strand DNA (ssDNA)-binding protein RPA2 was present at lower intensity on MSM PARs (Fig. 2c, Extended Data Fig. 4f).

Determinants that act in trans

To identify factors that are important for PAR behaviour, we eliminated different RMMAI proteins (MEI4 and ANKRD31) and axis proteins (HORMAD1 and REC8)16,20,25,26. Requirements for RMMAI blobs overlap with but are distinct from those for smaller RMMAI foci—for which HORMAD1 is key and MEI4 even more so, but ANKRD31 contributes only partially17,20,22 (Fig. 3a). HORMAD1 and REC8 were dispensable for RMMAI assembly in mo-2 regions, and for the elongation of the PAR axis, splitting of sister axes and formation of short loops (that is, compact mo-2 and REC114 signals) (Fig. 3a–c, Extended Data Fig. 6a, b). The distal PAR axes were separated in Rec8−/− mutants (Fig. 3c, Extended Data Fig. 6c), demonstrating that REC8 is essential for cohesion at the PAR end.

Fig. 3: Requirements for recruitment of RMMAI proteins and PAR axis remodelling.
figure 3

a, Top, quantification of REC114, ANKRD31, MEI4 and IHO1 foci along unsynapsed axes in wild-type and mutant leptotene or early-zygotene spermatocytes. Error bars are mean ± s.d. Comparisons to the wild type are indicated (*P < 0.02, **P ≤ 10−7, NS, not significant (P > 0.05), two-sided Student’s t-test; for exact P values, see the associated Source Data). Bottom, representative micrographs of REC114 staining; other proteins are shown in Extended Data Fig. 6a. Arrowheads indicate mo-2 associated blobs. Presence (+) or absence (–) of blobs is indicated below. Scale bars, 2 μm. b, Genetic requirements for PAR loop–axis organization (length of REC114 and mo-2 FISH signals along the PAR axis and axis-orthogonal extension of mo-2). Error bars are mean ± s.d. c, Representative SIM images of the loop–axis structure of the Y chromosome PAR in each mutant at late zygonema. Scale bar, 1 μm.

Source Data

The smaller MEI4 and REC114 foci still formed in Ankrd31−/− mutants, but were fewer and weaker20 (Fig. 3a, Extended Data Fig. 6a, d, e). In mo-2 regions, by contrast, RMMAI proteins did not accumulate detectably in Mei4−/− and Ankrd31−/− mutants (Fig. 3a, Extended Data Fig. 6a, b). ANKRD31 was dispensable for enrichment of heterochromatin factors (Extended Data Fig. 6f). REC114—though not IHO1—is also essential for RMMAI blobs21. The normal ultrastructure of the PAR was absent in Mei4−/− and Ankrd31−/− mutants: axes were short with no sign of splitting and mo-2 was decompacted (Fig. 3b, c, Extended Data Fig. 6b). We conclude that RMMAI blobs at the PAR share genetic requirements with autosomal mo-2 blobs, and that the presence of blobs correlates with normal structural differentiation of the PAR.

Axis remodelling and mo-2

If mo-2 arrays are cis-acting determinants of a high level of RMMAI recruitment, which in turn governs the structural dynamics of the PAR, then autosomal mo-2 regions should also form PAR-like structures. Consistent with this, the distal end of chromosome 9 underwent splitting in spermatocytes in which this region was late to synapse (Fig. 4a) and showed a PAR-like pattern of extended axes and compact chromatin dependent on Ankrd31 (Extended Data Fig. 7a). Thus, mo-2 (and/or linked elements) may be sufficient for both RMMAI recruitment and axis remodelling. We observed less axis remodelling for MSM PARs than B6 PARs (Extended Data Fig. 7b), which reinforces the correlation between mo-2 copy number, RMMAI levels and PAR ultrastructure.

Fig. 4: PAR-like structural reorganization and DSB formation on autosomal mo-2 arrays.
figure 4

a, The mo-2 region of chromosome 9 undergoes axis elongation and splitting similar to PARs (SIM image of a wild-type zygotene spermatocyte). Scale bar, 1 μm. b, ANKRD31 is required for high levels of DSB formation in mo-2 regions and for X–Y chromosome pairing. Immunolabelling with FISH (immuno-FISH) for RPA2 and mo-2 was used to detect DSBs. Immuno-FISH images for other mutants are shown in Extended Data Fig. 7c. Scale bar, 1 μm. c, PAR-like DSB formation near autosomal mo-2 regions. SSDS coverage7,20 is shown for the Y chromosome PAR (left) and the mo-2-adjacent region of chromosome 9 (right). The positions of mo-2 repeats are shown below. Examples for other chromosomes are shown in Extended Data Fig. 8a. d, Early pachytene XY oocyte showing bright RPA2 foci in the PAR (arrowheads). Scale bars, 2 μm (main image); 1 μm (inset).

DSB formation in spermatocytes

We hypothesized that RMMAI recruitment and axis remodelling create an environment that is conducive to a high level of DSB formation. This idea predicts that mutations should affect all of these processes (RMMAI recruitment, axis remodelling and DSB formation) coordinately and that PAR-like DSB formation should occur in autosomal mo-2 regions. We counted axial RPA2 foci as a proxy for the global number of DSBs and assessed the overlap of mo-2 with RPA2 (Fig. 4b, Extended Data Fig. 7c–f).

In wild-type zygotene spermatocytes, RPA2 foci overlapped with 35% of the mo-2 regions of each cell on average, increasing to 70% at pachynema (Extended Data Fig. 7e). Similar to the PAR2, autosomal mo-2 regions often acquired DSBs late (Extended Data Fig. 7g). By contrast, Ankrd31−/− mutants showed a substantial reduction in the overlap of RPA2 foci with mo-2, and X and Y chromosomes were paired in only 6% of mid-pachytene spermatocytes (Fig. 4b, Extended Data Fig. 7e, h). This is distinct from autosomes: global RPA2 foci were only modestly reduced in Ankrd31−/− cells (Extended Data Fig. 7d), and in most cells all autosomes were paired and synapsed20,21. (Ankrd31−/− mutants form fewer RPA2 foci at leptonema and early zygonema, but normal numbers thereafter20,21.)

Rec8 deficiency did not reduce the formation of RPA2 foci at mo-2 regions or more globally relative to a synapsis-deficient control (Syce1−/−) (Extended Data Fig. 7c–e). However, X–Y chromosome pairing was reduced (Extended Data Fig. 7h), presumably because REC8 promotes inter-homologue recombination27. Hormad1−/− spermatocytes had comparable or higher frequencies of mo-2-overlapping RPA2 foci and X–Y chromosome pairing to the Syce1−/− control (Extended Data Fig. 7e, h). The high frequency of RPA2 foci at mo-2 regions was notable, given the global reduction in the number of RPA2 foci (Extended Data Fig. 7d, f) and DSBs28, but is consistent with the dispensability of HORMAD1 both for the recruitment of RMMAI proteins to mo-2 and for the ultrastructure of the PAR (Fig. 3a–c).

These findings show that RMMAI recruitment and axis remodelling are tightly correlated with a high frequency of DSB formation. Further strengthening this correlation, we noted above that MSM PARs exhibit a lower RPA2 intensity than B6 PARs (Fig. 2c)—perhaps reflecting a lesser tendency to make multiple DSBs. Indeed, multiple PAR RPA2 foci were resolved more frequently in B6 than in MSM mice (Extended Data Fig. 7i, j).

We used maps of ssDNA bound by the strand-exchange protein DMC1 (ssDNA sequencing; SSDS)7,29,30 to test more directly whether PAR-like DSB formation occurs in autosomal mo-2 regions (that is, dependent on ANKRD31 but largely independent of the histone methyltransferase PRDM9)7,20,21 (Fig. 4c, Extended Data Fig. 8a). We found that the region that encompasses the chromosome 9 mo-2 cluster showed an accumulation of SSDS reads that was substantially reduced in Ankrd31−/− but not in Prdm9−/− mutants. A modest ANKRD31-dependent, PRDM9-independent peak was also observed near the mo-2 cluster on chromosome 13 (Extended Data Fig. 8a). Thus, autosomal mo-2 regions not only accumulate PAR-like levels of RMMAI proteins and undergo PAR-like axis remodelling in spermatocytes, but also frequently form DSBs in a PAR-like manner.

Mo-2 regions in oocytes

In females, recombination between the two X chromosomes is not restricted to the PAR, so oocytes do not require DSBs in the PAR as with spermatocytes31. We therefore asked whether the PAR undergoes spermatocyte-like structural changes in oocytes. RMMAI proteins robustly accumulated at the PAR and in autosomal mo-2 regions from leptonema to pachynema (Extended Data Fig. 9a), consistent with studies of MEI4 and ANKRD3116,21. Oocytes also had an extended PAR axis and a compact PARb FISH signal from leptonema to zygonema and transitioned to a shorter axis and a more extended PARb signal in pachynema, with a loss of REC114 signal upon synapsis (Extended Data Fig. 9b). Heterochromatin factors were also enriched (Extended Data Fig. 9c). However, in contrast to spermatocytes, we did not detect a thickening or splitting of the PAR axis, or the accumulation of REC8 (Extended Data Fig. 9d)—even in the absence of synapsis in Syce1−/− mutants (Extended Data Fig. 9e). Moreover, similar to the PAR31, autosomal mo-2 regions showed little enrichment for SSDS signal in wild-type ovaries (Extended Data Fig. 8b, c).

A low SSDS signal despite the enrichment of RMMAI proteins and the presence of long axes could indicate that oocytes lack a critical factor (or more than one factor) that promotes DSBs in the PAR of spermatocytes. Alternatively, oocyte PARs may not realize their full DSB potential because of negative feedback tied to homologue engagement32,33; perhaps synapsis that is initiated elsewhere on the X chromosome often spreads into the PAR and disrupts its ultrastructure before DSBs can form. To investigate this idea, we tested the effects of delaying or blocking PAR synapsis using sex-reversed XY females34 and Syce1−/− mutants.

XY oocytes pair and synapse their PARs relatively late—only 28% of late-zygotene cells had X and Y chromosomes that were paired and/or synapsed (25 of 90 cells from two mice), increasing to 66% at pachynema (115 of 174 cells). This late pairing and synapsis is reminiscent of spermatocytes, but appears to be less efficient. Most pachytene XY oocytes that synapsed their PARs had a PAR-associated RPA2 focus, which occurred at twice the frequency and with higher immunofluorescence intensity than in XX oocytes (Fig. 4d, Extended Data Fig. 9f). RPA2 foci were also seen on most PARs that failed to synapse (Extended Data Fig. 9g). By contrast, chromosomes 9 and 13 showed a lower frequency and intensity of RPA2 foci, comparable to those in XX PARs and not differing between XY and XX oocytes (Extended Data Fig. 9f).

These findings suggest that delayed PAR synapsis allows oocytes to form DSBs more efficiently. In support of this conclusion, the absence of synapsis in Syce1−/− oocytes was accompanied by an increase in both the frequency and the intensity of RPA2 foci at PARs and autosomal mo-2 regions alike (Extended Data Fig. 9h). Our results do not exclude the possibility of differences between the trans-acting factors of spermatocytes and oocytes, but we infer that the ability to manifest a high level of DSB formation depends substantially on the result of a race between the formation of DSBs and the completion of synapsis (Supplementary Discussion).

Discussion

We have shown here that the PAR in male mice undergoes a notable rearrangement of loop–axis structure before DSBs form, and that this rearrangement involves the recruitment of RMMAI proteins, dynamic axis elongation and splitting of sister chromatid axes (Extended Data Fig. 10). Most of these behaviours also occur in oocytes, which can also support a high rate of DSB formation in the PAR if synapsis is delayed. The mo-2 array may be a key cis-acting determinant, and RMMAI proteins are crucial trans-acting determinants. Although the function of sister axis splitting is unclear (Supplementary Discussion), the full suite of PAR behaviours appears essential for the pairing, recombination and segregation of heteromorphic sex chromosomes.

Budding yeast also uses the robust recruitment of Rec114 and Mer2 (the IHO1 orthologue) to ensure that its smallest chromosomes incur DSBs35. Thus, such preferential recruitment is an evolutionarily recurrent strategy for mitigating the risk of recombination failure when the length of chromosomal homology is limited.

The hyperaccumulation of RMMAI proteins may reflect the binding of one or more of these proteins to an mo-2-associated chromatin structure and/or direct binding to mo-2 repeats or another tightly linked DNA element. We note that the repetitive mo-2 array imposes a risk of unequal exchange23,36. Thus, paradoxically, the DNA structure of the PAR stabilizes the genome by supporting the segregation of sex chromosomes, but also promotes the rapid evolution of mammalian PARs4.

Methods

Data reporting

No statistical methods were used to predetermine sample size. The experiments were not randomized and the investigators were not blinded to allocation during experiments and outcome assessment.

Mice

Mice were maintained and killed under USA regulatory standards and experiments were approved by the Memorial Sloan Kettering Cancer Center (MSKCC) Institutional Animal Care and Use Committee (IACUC, protocol number 01-03-007). Mice were fed regular rodent chow with ad libitum access to food and water. The Ankrd31 knockout allele (Ankrd31em1Sky) is a single-base-insertion mutation (+A) in exon 3; its generation and phenotypic characterization have previously been described20. Mice with the Mei4 knockout allele16 were provided by B. de Massy. All other mouse strains were purchased from the Jackson Laboratory: C57BL/6J (stock no. 00664), MSM/MsJ (stock no. 003719), B6N(Cg)-Syce1tm1b(KOMP)Wtsi/2J (stock no. 026719), B6;129S7-Hormad1tm1Rajk/Mmjax (stock no. 41469-JAX), B6;129S4-Rec8mei8/JcsMmjax (stock no. 34762-JAX) and B6.Cg-Tg(Sry)2Ei Srydl1Rlb/ArnoJ (stock no. 010905). Mice were genotyped using Direct Tail lysis buffer (Viagen) following the manufacturer’s instructions.

B6.Cg-Tg(Sry)2Ei Srydl1Rlb/ArnoJ males have a Y chromosome with a deletion of the sex-determining Sry gene and also have an Sry transgene integrated on an autosome. When these males are crossed with C57BL/6J females, those XY and XX mice that do not inherit the Sry transgene develop as females.

Generation of REC8 and REC114 antibodies

To produce antibodies against REC8, a fragment of the mouse Rec8 gene encoding amino acids 36 to 253 (NCBI reference sequence: NP_001347318.1) was cloned into a pGEX-4T-2 vector. The resulting fusion of the REC8 fragment fused to glutathione S-transferase (GST) was expressed in Escherichia coli, affinity-purified on glutathione sepharose 4B and cleaved with Precision protease. Antibodies were raised in rabbits by Covance against the purified recombinant REC8 fragment, and antibodies were affinity-purified using GST-REC836-253 that had been immobilized on glutathione sepharose by cross-linking with dimethyl pimelimidate; bound antibodies were eluted with 0.1 M glycine, pH 2.5. Purified antibodies were tested in western blots of testis extracts and specificity was validated by immunostaining of spread meiotic chromosomes from wild-type and Rec8−/− mice.

To produce antibodies against REC114, a fragment of the mouse Rec114 gene encoding a truncated polypeptide lacking the N-terminal 110 amino acids (NCBI reference sequence: NP_082874.1) was cloned into a pET-19b expression vector. The resulting His6-tagged REC114(111–259) fragment was insoluble when expressed in E. coli, so the recombinant protein was solubilized and affinity-purified on Ni-NTA resin in the presence of 8 M urea. Eluted protein was dialysed against 100 mM NaH2PO4, 10 mM Tris-HCl, 6 M urea, pH 7.3 and used to immunize rabbits (Covance). Antibodies were affinity-purified against purified recombinant His6–REC114(111–259) protein immobilized on cyanogen-bromide-activated sepharose and eluted in 0.2 M glycine, pH 2.5. The affinity-purified antibodies were used in a previous study17 that reported detection of a band of appropriate molecular weight in western blots of testis extracts. Subsequent analysis showed that this band is also present in extracts of Rec114−/− testes, and thus is non-specific (C. Brun and B. de Massy, personal communication). However, the initial study17 also reported detection of immunostaining foci on spread meiotic chromosomes similar to findings reported here and in another previous study20. This immunostaining signal is absent from chromosome spreads prepared from Rec114−/− mutant mice (C. Brun and B. de Massy, personal communication). Moreover, this immunostaining signal is indistinguishable from that reported using independently generated and validated anti-REC114 antibodies19. We conclude that our anti-REC114 antibodies are highly specific for the cognate antigen when used for immunostaining of meiotic chromosome spreads.

Chromosome spreads

Testes were dissected and deposited after removal of the tunica albuginea in 1 × PBS, pH 7.4. Seminiferous tubules were minced using forceps to form a cell suspension. The cell suspension was filtered through a 70-μm cell strainer into a 15-ml Falcon tube pre-coated with 3% (w/v) BSA, and was centrifuged at 200g for 5 min. The cell pellet was resuspended in 12 ml of 1 × PBS for an additional centrifugation step at 200g for 5 min and the pellet was resuspended in 1 ml hypotonic buffer containing 17 mM sodium citrate, 50 mM sucrose, 30 mM Tris-HCl, pH 8, 5 mM EDTA, pH 8, 0.5 mM dithiothreitol (DTT) and 10 μl of 100 × Halt protease inhibitor cocktail (Thermo Fisher Scientific), and incubated for 8 min. Next, 9 ml of 1 × PBS was added and the cell suspension was centrifuged at 200g for 5 min. The cell pellet was resuspended in 100 mM sucrose pH 8 to obtain a slightly turbid cell suspension, and incubated for 10 min. Superfrost glass slides were divided into two squares using an ImmEdge hydrophobic pen (Vector Labs), then 110 μl of 1% paraformaldehyde (PFA) (freshly dissolved in presence of NaOH at 65 °C, 0.15% Triton, pH 9.3, cleared through a 0.22-μm filter) and 30 μl of cell suspension were added per square and swirled three times for homogenization, and the slides were placed horizontally in a closed humid chamber for 2 h. The humid chamber was opened for 1 h to allow almost complete drying of the cell suspension. Slides were washed in a Coplin jar 2 × 5 min in 1 × PBS on a shaker, and 2 min with 0.4% Photo-Flo 200 solution (Kodak), air-dried and stored in aluminium foil at −80 °C.

Ovaries were extracted from mice 14.5–18.5 days post-coitum (dpc), and collected in 1 × PBS, pH 7.4. After a 15-min incubation in hypotonic buffer, the ovaries were placed on a slide containing 30 μl of 100 mM sucrose, pH 8, and dissected with forceps to form a cell suspension. The remaining tissues were removed, 110 μl of 1% PFA–0.15% Triton was added and the slides were gently swirled for homogenization, before incubation in a humid chamber as described above for spermatocyte chromosome spreads.

Immunostaining

Slides of meiotic chromosome spreads were blocked for 30 min at room temperature horizontally in a humid chamber with an excess of blocking buffer containing 1 × PBS, pH 7.4 with 0.05% Tween-20, 7.5% (v/v) donkey serum, 0.5 mM EDTA, pH 8.0 and 0.05% (w/v) sodium azide, and cleared by centrifugation at 13,000 rpm for 15 min in an Eppendorf microcentrifuge (with rotor FA-45-24-11). Slides were incubated with primary antibody overnight in a humid chamber at 4 °C, or for at least 3 h at room temperature. Slides were washed 3 × 5 min in 1 × PBS, 0.05% Tween-20, then blocked for 10 min and incubated with secondary antibody for 1–2 h at 37 °C in a humid chamber. Slides were washed 3 × 5 min in the dark on a shaker with 1 × PBS and 0.05% Tween-20, then rinsed in H2O, and mounted before air-drying with Vectashield (Vector Labs). Antibody dilutions were centrifuged at 13,000 rpm for at least 5 min before use. The primary antibodies used were rabbit and guinea pig anti-ANKRD3120 (1:200 dilution), rabbit anti-HORMAD2 (Santa Cruz, sc-82192, 1:50), guinea pig anti-HORMAD2 (1:200) and guinea pig anti-IHO1 (1:200) (gifts from A. Tóth), goat anti-MEI1 (Santa Cruz, sc-86732, 1:50), rabbit anti-MEI4 (gift from B. de Massy, 1:200), rabbit anti-REC8 (this study, 1:100), rabbit anti-REC114 (this study, 1:200), rabbit anti-RPA2 (Santa Cruz, sc-28709, 1:50), goat anti-SYCP1 (Santa Cruz, sc-20837, 1:50), rabbit anti-SYCP2 (Atlas Antibodies, HPA062401, 1:100), mouse anti-SYCP3 (Santa Cruz, sc-74569, 1:100), goat anti-SYCP3 (Santa Cruz, sc-20845, 1:50), rabbit anti-TRF1 (Alpha Diagnostic, TRF12-S, 1:100), rabbit anti-H4K20me3 (Abcam, ab9053, 1:200), rabbit anti-H3K9me3 (Abcam, ab8898, 1:200), mouse anti-macroH2A1.2 (Active Motif, 61428, 1:100), mouse anti-HP1γ (Millipore, MAB3450, 1:100), mouse anti-HP1β (Millipore, MAB3448, 1:100), rabbit anti-HP1β (Genetex, GTX106418, 1:100), rabbit anti-Mi2 (recognizes CHD3 and CHD4; Santa Cruz, sc-11378, 1:50), rabbit anti-ATRX (Santa Cruz, sc-15408, 1:50), mouse anti-DMRT1 (Santa Cruz, sc-377167, 1:50), rabbit anti-ZMYM3 (Abcam, ab19165, 1:300), rabbit anti-PAXIP1 (EMD Millipore, ABE1877, 1:300). The secondary antibodies used were CF405S anti-guinea pig (Biotium, 20356), CF405S anti-rabbit (Biotium, 20420), CF405S anti-mouse (Biotium, 20080), Alexa Fluor 488 donkey anti-mouse (Life Technologies, A21202), Alexa Fluor 488 donkey anti-rabbit (Life Technologies, A21206), Alexa Fluor 488 donkey anti-goat (Life Technologies, A11055), Alexa Fluor 488 donkey anti-guinea pig (Life Technologies, A11073), Alexa Fluor 568 donkey anti-mouse (Life Technologies, A10037), Alexa Fluor 568 donkey anti-rabbit (Life Technologies, A10042), Alexa Fluor 568 goat anti-guinea pig (Life Technologies, A11075), Alexa Fluor 594 donkey anti-mouse (Life Technologies, A21203), Alexa Fluor 594 donkey anti-rabbit (Life Technologies, A21207), Alexa Fluor 594 donkey anti-goat (Life Technologies, A11058), Alexa Fluor 647 donkey anti-rabbit (Abcam, ab150067), Alexa Fluor 647 donkey anti-goat (Abcam, ab150131), all at 1:250 dilution.

Immuno-FISH and DNA probe preparation

All steps were performed in the dark to prevent loss of fluorescence from prior immunostaining. After the last washing step in the immunostaining protocol, slides were placed horizontally in a humid chamber and the chromosome spreads were re-fixed with an excess of 2% (w/v) PFA in 1 × PBS (pH 9.3) for 10 min at room temperature. Slides were rinsed once in H2O, washed for 4 min in 1 × PBS, sequentially dehydrated with 70% (v/v) ethanol for 4 min, 90% ethanol for 4 min and 100% ethanol for 5 min, and air-dried vertically for 5–10 min. Next, 15 μl of hybridization mix was applied containing the DNA probe(s) in 70% (v/v) deionized formamide (Amresco), 10% (w/v) dextran sulfate, 2 × saline sodium citrate (SSC) buffer, 1 × Denhardt’s buffer, 10 mM EDTA, pH 8 and 10 mM Tris-HCl, pH 7.4. Cover glasses (22 × 22 mm) were applied and sealed with rubber cement (Weldwood contact cement), then the slides were denatured on a heat block for 7 min at 80 °C, followed by overnight incubation (for at least 14 h) at 37 °C. Cover glasses were carefully removed using a razor blade, slides were rinsed in 0.1 × SSC buffer, washed in 0.4 × SSC, 0.3% NP-40 for 5 min, washed in PBS–0.05% Tween-20 for 3 min, rinsed in H2O and mounted with Vectashield before air-drying.

To generate FISH probes, we used the nick translation kit from Abbott Molecular following the manufacturer’s instructions and using CF-dye-conjugated dUTP (Biotium), on bacterial artificial chromosome (BAC) DNA from the clones RP24-500I4 (maps to the region of the PAR boundary; PARb probe) CH25-592M6 (maps to the distal PAR, PARd probe), RP23-139J18, RP24-136G21 and CH36-200G6 (centromere-distal ends of chromosomes 4, 9 and 13, respectively). BAC clones were obtained from the BACPAC Resource Center (CHORI). Labelled DNA (500 ng) was precipitated during a 30-min incubation at −20 °C after adding 5 μl mouse Cot-1 DNA (Invitrogen), 0.5 volume of 7.5 M ammonium acetate and 2.5 volumes of cold 100% ethanol. After washing with 70% ethanol and air-drying in the dark, the pellet was dissolved in 15 μl hybridization buffer.

Mo-2 oligonucleotide probes were synthesized by Integrated DNA Technologies, with 6-FAM or TYE 665 fluorophores added to both 5′ and 3′ ends of the oligonucleotide. The DNA sequence was designed on the basis of the previously defined consensus sequence24, and the probe was used at a final concentration of 10 pmol μl−1 in hybridization buffer without Cot-1 DNA. The Y chromosome paint probe was purchased from ID Labs and used at 1:30 dilution in hybridization buffer without Cot-1 DNA.

EdU incorporation

Seminiferous tubules were incubated in Dulbecco’s modified Eagle’s medium (DMEM) with 10% FCS and 10 μM 5-ethynyl-3′-deoxyuridine (EdU) at 37 °C for 1 h for in vitro labelling. EdU incorporation was detected using the Click-iT EdU Alexa Fluor 647 imaging kit (Invitrogen) according to the manufacturer’s instructions.

Image acquisition

Images of spread spermatocytes were acquired on a Zeiss Axio Observer Z1 Marianas Workstation, equipped with an ORCA-Flash 4.0 camera and DAPI, CFP, FITC, TEXAS red and Cy5 filter sets, illuminated by an X-Cite 120 PC-Q light source, with either a 63×/1.4 NA oil immersion objective or a 100×/1.4 NA oil-immersion objective. Marianas Slidebook 5.0 (Intelligent Imaging Innovations) software was used for acquisition.

Structured illumination microscopy (3D-SIM) was performed at the Bio-Imaging Resource Center in Rockefeller University using an OMX Blaze 3D-SIM super-resolution microscope (Applied Precision), equipped with 405-nm, 488-nm and 568- nm lasers, and a 100×/1.40 NA UPLSAPO oil objective (Olympus). Image stacks of several-μm thickness were taken with 0.125-μm z-steps, and were reconstructed in Deltavision softWoRx 6.1.1 software with a Wiener filter of 0.002 using wavelength-specific experimentally determined optical transfer functions. Slides were prepared and stained as described above, except that chromosomes were spread only on the central portion of the slides, and the slides were mounted using 18 × 18-mm coverslips (Zeiss).

Image analysis

The 3D-SIM images are shown either as a z-stack using the sum slices function in Fiji, or as a unique slice. The X and/or Y chromosomes were cropped, rotated and further cropped for best display. For montage display, the X and Y chromosome images were positioned on a black background using Adobe Illustrator 2020 (v.24.1). In the instances in which the axes of the X and Y chromosomes were cropped, the area of cropping was labelled with a light grey dotted line. Loop and axis measurements, foci counts and fluorescence intensity quantification were only performed on images from conventional microscopy using the original, unmodified data.

To measure the colocalization between RMMAI proteins, we costained for SYCP3 and ANKRD31 along with MEI4, REC114 or IHO1, and manually counted the number of ANKRD31 foci that overlapped with SYCP3 and colocalized (or not) with MEI4, REC114 or IHO1. These counts were performed in 16 spermatocytes from leptonema to early or mid-zygonema.

To quantify the total number of RPA2, MEI4, REC114, ANKRD31 and IHO1 foci, single cells were manually cropped and analysed with semi-automated scripts in Fiji37 (v.2.0.0-rc-69/1.52p) as previously described in detail20. In brief, images were auto-thresholded on SYCP3 staining, which was used as a mask to use ‘Find Maxima’ to determine the number of foci. Images were manually inspected to determine that there were no obvious defects in determining SYCP3 axes, that no axes from neighbouring cells were counted, that no artefacts were present, and that no foci were missed by the script.

To test for colocalization between RPA2 and mo-2 FISH signals, we manually scored the percentage of mo-2 FISH signals colocalizing at least partly with RPA2. Depending on the progression of synapsis during prophase I, between eight and four discrete mo-2 FISH signals could be detected, corresponding to (with increasing signal intensity) chromosome 4, chromosome 13, chromosome 9 and the PAR (two signals for each when unpaired, or a single signal for each after homologous pairing and synapsis). Notably, the RPA2 focus was most often found in a slightly more centromere-proximal position compared to the bulk of mo-2 FISH signals, and therefore colocalized partly with mo-2 FISH signals. In the case of the PAR, this position corresponds closely to the region of the PAR boundary (PARb probe). A similar trend was observed on autosomal mo-2 clusters.

For estimates of chromatin extension, we measured the maximal axis-orthogonal distance between the FISH signal and the centre of the PAR axis, or the centromere-distal axis for chromosome 9 stained by SYCP3. In mutant mice defective for RMMAI protein recruitment in the mo-2 regions, the PAR axis was defined as the nearest SYCP3 segment adjacent to the telomeric SYCP3 signal.

For quantification of RPA2, ANKRD31, REC8 and mo-2 signal intensity in B6 × MSM and MSM × B6 F1 hybrids, late-zygotene spermatocytes with at least one RPA2 focus on the X or Y PAR were analysed. We used the elliptic selection tool in Fiji to define a region of interest around the largest signal in the PAR, and the same selection tool was then positioned on the other PAR axis for comparison. The fluorescence intensity was measured as the integrated density with background subtraction.

Prophase I sub-staging and identification of the PAR

Nuclei were staged according to the dynamic behaviour of the autosome and sex-chromosome axes during prophase I, using SYCP3 staining. Leptonema was defined as having short stretches of SYCP3 but no evidence of synapsis; early or mid-zygonema as having longer stretches of SYCP3 staining and some synapsis; and late zygonema as having fully assembled chromosome axes and substantial (more than 70%) synapsis. The X and Y chromosomes generally can be identified at this stage, and the PAR axis is distinguishable because it appears thicker than the centromeric end—particularly near the end of zygonema, when autosomes are almost fully synapsed. Early pachynema was defined as complete autosomal synapsis, whereas the X and Y chromosomes could display various configuration: i) unsynapsed, with thickened PAR axes; ii) engaged in PAR synapsis; and iii) synapsed in the PAR and non-homologously synapsed along the full (or nearly full) Y chromosome axis. Mid-pachynema was defined as showing bright signal from autosome axes, with desynapsing X and Y axes remaining synapsed only in the PAR and a short PAR axis. During this stage, the autosomes and the non-PAR X and Y chromosome axes are initially short and thick, and progressively become longer and thinner. Late pachynema was defined as brighter autosome axes with a characteristic thickening of all autosome ends. The X and Y chromosome non-PAR axes are then long and thin and show excrescence of axial elements. Diplonema was defined as brighter axes and desynapsing autosome, associated with prominent thickening of the autosome ends, particularly the centromeric ends. In early diplonema, the non-PAR axes of X and Y chromosomes are still long and thin and progressively condense to form bright axes, associated with bulges. Most experiments were conducted using SYCP3 in combination with a RMMAI protein, which allows easier distinction between synapsing and desynapsing X and Y chromosomes.

By using only SYCP3 staining, the PARs can only be identified unambiguously from the late zygonema-to-early pachynema transition through to diplonema. From pre-leptonema to mid-or late zygonema, the PARs were identified as the two brightest RMMAI signals, the two brightest mo-2 FISH signals, the two brightest PARb FISH signals or the two FISH signals from the PARd probe. The Y chromosome PAR could be distinguished from the X chromosome PAR using the PARb probe, as this probe also weakly stains the chromatin of the non-PAR portion of the Y chromosome.

Measurements of the PAR chromatin loops and axis length in oocytes were performed on two 14.5–15.5 dpc female fetuses (enriched for leptotene and zygotene oocytes) and two 18.5 dpc female fetuses (enriched for pachytene oocytes).

We found notable variability in the length of the X or Y chromosome PAR axis between different mice in our mouse colony maintained in a C57BL/6J congenic background, and even between different C57BL/6J males obtained directly from the Jackson Laboratory. This is in agreement with previous reports about the hypervariable nature of the mo-2 minisatellite and its involvement in unequal crossing over in the mouse6,24,36,38,39 (mo-2 was also named DXYmov15 or Mov15 flanking sequences). However, the intensity and elongation of the RMMAI signal and the length of the PAR axis were always correlated with mo-2 FISH signal intensity. Despite this variability, mo-2 and RMMAI proteins were enriched in the PAR and autosome ends of all mice analysed.

Analysis of SSDS data

SSDS sequencing data were from previously described studies7,20,31 and are all available at the Gene Expression Omnibus (GEO) under the accession numbers GSE35498, GSE99921 and GSE118913. To define the enrichment values presented in Extended Data Fig. 8b, the SSDS coverage was summed across the indicated coordinates adjacent to the mo-2 repeats. A chromosomal mean and standard deviation for chromosome 9 was estimated by dividing the chromosome into 4-kb bins, summing the SSDS coverage in each bin and calculating the mean and standard deviation after excluding those bins that overlapped a DSB hotspot. The enrichment score was then defined as the difference between the coverage in the mo-2-adjacent region and the mean coverage for chromosome 9, divided by the chromosome 9 standard deviation.

Immunoprecipitation and mass spectrometry

Immunoprecipitations were carried out on samples from wild-type and Ankrd31−/− mice using two separate polyclonal anti-ANKRD31 antibodies raised in rabbit and guinea pig20 (four samples total). Two additional immunoprecipitations were performed using an anti-cyclin B3 antibody on either wild-type or Ccnb3-knockout testes40,41; these samples serve as additional negative controls for the ANKRD31 interaction screen. For each sample, protein extracts were prepared from testes of three 12-day-old mice in 1 ml RIPA buffer (50 mM Tris-HCl, 150 mM NaCl, 0.1% SDS, 0.5% sodium deoxycholate, 1% NP-40, 10 mM MgCl2 and 100 units of benzonase for 1h at 4 °C. After centrifugation at 13,000 rpm for 20 min at 4 °C, the lysate was pre-cleared using 30 μl of a slurry of protein A/G Dynabeads for 1 h at 4 °C. Next, 50 μl of protein A/G beads coupled for 30 min with 10 μg of anti-ANKRD31 or anti-cyclin B3 antibody (monoclonal antibody 5 from ref. 40) were added and the solution was incubated overnight at 4 °C on a rotating rack. Beads were washed 3 times in 1 ml RIPA buffer and once with 1 ml of 50 mM ammonium bicarbonate. Samples were then digested overnight with 2 μg trypsin in 80 μl of 50 mM ammonium bicarbonate at 37 °C on a thermo mixer (850 rpm). Peptides were desalted using C18 zip tips, and then dried by vacuum centrifugation. Each sample was reconstituted in 10 μl 0.1% (v/v) formic acid and 4 μl was analysed by microcapillary liquid chromatography with tandem mass spectrometry (MS/MS) using the NanoAcquity (Waters) with an ACQUITY UPLC BEH C18 Column (Waters) configured with an ACQUITY UPLC M-Class Symmetry C18 trap column (Waters) coupled to a QExactive Plus mass spectrometer (Thermo Fisher Scientific). Peptides were eluted with a linear gradient of 0–35% acetonitrile (0.1% formic acid) in water (0.1% formic acid) over 150 min with a flow rate of 300 nl per min. The QE Plus was operated in automatic, data-dependent MS/MS acquisition mode with one MS full scan (380–1,800 m/z) at 70,000 mass resolution and up to ten concurrent MS/MS scans for the ten most intense peaks that were selected from each survey scan. Survey scans were acquired in profile mode and MS/MS scans were acquired in centroid mode at 17,500 resolution with an isolation window of 1.5 amu and normalized collision energy of 27. Automatic gain control (AGC) was set to 1 × 10 for MS1 and 5 × 10 and 100 ms IT for MS2. Charge state exclusion rejected ions having unassigned charge states or having a charge state above 6, with a dynamic exclusion of 15 s. All MS/MS samples were analysed using MaxQuant v.1.5.3.3 (Max Planck Institute of Biochemistry) at default settings with a few modifications.

Yeast two-hybrid assay

Mouse testis cDNAs for Ptip (also known as Paxip1), Zmym3 and Ankrd31 were amplified and cloned in vectors to generate fusion proteins with the Gal4 DNA-binding domain (Gal4BD) or activation domain (Gal4AD). Assays were conducted according to the manufacturer’s instructions (Clontech).In brief, Y2HGold and Y187 (Clontech) yeast haploid strains were transformed with constructs encoding Gal4BD and Gal4AD fusion proteins. After mating on YPD plates, diploid cells expressing Gal4BD and Gal4AD fusion proteins were selected on double dropout medium lacking leucine and tryptophan. Protein interactions were assayed by spotting diploid-cell suspensions on selective medium lacking leucine, tryptophan, histidine and adenine (quadruple dropout), and quadruple dropout medium containing X-α-gal (5-bromo-4-chloro-3-indolyl α-d-galactopyranoside) and aureobasidin A, and growing for 3 days at 30 °C.

Statistical analysis

All statistical tests were performed in R (v.3.4.4)42 and RStudio (v.1.1.442). Negative binomial regression was calculated using the glm.nb function from the MASS package (v.7.3-49)43.

Statistics and reproducibility

The pictures shown in this article are representative images that aim to illustrate the findings in the clearest manner. Any conclusion or statement regarding the results that is not associated with explicit quantification is based on the imaging and analysis of at least 20 cells, sometimes hundreds, usually from multiple mice. Details for the main figures are as follows.

Fig. 1a: The thickening of the PAR axis (using SYCP3 staining) and the elongation of the RMMAI signal along the PAR axis were observed in more than three different mice in hundreds of late-zygotene spermatocytes, mostly using our homemade antibodies against REC114 and ANKRD31. Other antibodies such as anti-SYCP2 and anti-HORMAD2 were used to confirm the PAR axis thickening, and anti-MEI1, anti-MEI4 and anti-IHO1 were used to confirm the elongation of the REC114 and ANKRD31 signal along the PAR axis, in more than 20 spermatocytes for each antibody. Fig. 1b: The PAR axis splitting, the extension of the RMMAI signal and the collapse of the PAR structure during X–Y chromosome synapsis were observed by SIM in more than 60 spermatocytes in more than 3 different mice. Fig. 2b: The colocalization between REC114 blobs (or RMMAI blobs in general) and mo-2 FISH signals was observed in all spermatocytes analysed (more than 200), from leptotene to early pachytene in more than 3 different mice. Fig. 3c: Axis splitting on the Y chromosome PAR was observed by SIM in more than 100 late-zygotene spermatocytes and in more than 20 zygotene-like spermatocytes from Hormad1−/− mice. The fork-shaped PAR structure in Rec8−/− mice was observed in more than 20 spermatocytes. The absence of PAR differentiation and decompaction of mo-2-containing chromatin was observed in more than 30 Ankrd31−/− spermatocytes and 20 Mei4−/− spermatocytes. This specific pattern was confirmed in at least three different mice of each genotype using conventional microscopy. The differentiation of the PAR axis becomes hardly detectable in Hormad1−/− mutants at a later stage in some pachytene-like spermatocytes as cells enter apoptosis, similar to Spo11−/− mutants. Fig. 4a: The differentiation of the non-centromeric end of chromosome 9 was observed in 6 spermatocytes by SIM and in more than 20 late-zygotene spermatocytes by conventional microscopy in 3 different mice.

Reporting summary

Further information on research design is available in the Nature Research Reporting Summary linked to this paper.