Protamines: lessons learned from mouse models

in Reproduction
Authors:
Lena ArévaloDepartment of Developmental Pathology, Institute of Pathology, University Hospital Bonn, Bonn, Germany

Search for other papers by Lena Arévalo in
Current site
Google Scholar
PubMed
Close
https://orcid.org/0000-0002-0681-0392
,
Gina Esther MergesDepartment of Developmental Pathology, Institute of Pathology, University Hospital Bonn, Bonn, Germany

Search for other papers by Gina Esther Merges in
Current site
Google Scholar
PubMed
Close
https://orcid.org/0000-0002-5805-965X
,
Simon SchneiderDepartment of Developmental Pathology, Institute of Pathology, University Hospital Bonn, Bonn, Germany
Bonn Technology Campus, Core Facility ‘Gene-Editing’, University Hospital Bonn, Bonn, Germany

Search for other papers by Simon Schneider in
Current site
Google Scholar
PubMed
Close
https://orcid.org/0000-0001-9894-5608
, and
Hubert SchorleDepartment of Developmental Pathology, Institute of Pathology, University Hospital Bonn, Bonn, Germany

Search for other papers by Hubert Schorle in
Current site
Google Scholar
PubMed
Close
https://orcid.org/0000-0001-8272-0076
View More View Less

Correspondence should be addressed to L Arévalo or H Schorle; Email: Lena.lueke@gmail.com or Schorle@uni-bonn.de
Free access

In brief

Protamines package and shield the paternal DNA in the sperm nucleus and have been studied in many mouse models over decades. This review recapitulates and updates our knowledge about protamines and reveals a surprising complexity in protamine function and their interactions with other sperm nuclear proteins.

Abstract

The packaging and safeguarding of paternal DNA in the sperm cell nucleus is a critical feature of proper sperm function. Histones cannot mediate the necessary hypercondensation and shielding of chromatin required for motility and transit through the reproductive tracts. Paternal chromatin is therefore reorganized and ultimately packaged by protamines. In most mammalian species, one protamine is present in mature sperm (PRM1). In rodents and primates among others, however, mature sperm contain a second protamine (PRM2). Unlike PRM1, PRM2 is cleaved at its N-terminal end. Although protamines have been studied for decades due to their role in chromatin hypercondensation and involvement in male infertility, key aspects of their function are still unclear. This review updates and integrates our knowledge of protamines and their function based on lessons learned from mouse models and starts to answer open questions. The combined insights from recent work reveal that indeed both protamines are crucial for the production of functional sperm and indicate that the two protamines perform distinct functions beyond simple DNA compaction. Loss of one allele of PRM1 leads to subfertility whereas heterozygous loss of PRM2 does not. Unprocessed PRM2 seems to play a distinct role related to the eviction of intermediate DNA-bound proteins and the incorporation of both protamines into chromatin. For PRM1, on the other hand, heterozygous loss leads to strongly reduced sperm motility as the main phenotype, indicating that PRM1 might be important for processes ensuring correct motility, apart from DNA compaction.

Abstract

In brief

Protamines package and shield the paternal DNA in the sperm nucleus and have been studied in many mouse models over decades. This review recapitulates and updates our knowledge about protamines and reveals a surprising complexity in protamine function and their interactions with other sperm nuclear proteins.

Abstract

The packaging and safeguarding of paternal DNA in the sperm cell nucleus is a critical feature of proper sperm function. Histones cannot mediate the necessary hypercondensation and shielding of chromatin required for motility and transit through the reproductive tracts. Paternal chromatin is therefore reorganized and ultimately packaged by protamines. In most mammalian species, one protamine is present in mature sperm (PRM1). In rodents and primates among others, however, mature sperm contain a second protamine (PRM2). Unlike PRM1, PRM2 is cleaved at its N-terminal end. Although protamines have been studied for decades due to their role in chromatin hypercondensation and involvement in male infertility, key aspects of their function are still unclear. This review updates and integrates our knowledge of protamines and their function based on lessons learned from mouse models and starts to answer open questions. The combined insights from recent work reveal that indeed both protamines are crucial for the production of functional sperm and indicate that the two protamines perform distinct functions beyond simple DNA compaction. Loss of one allele of PRM1 leads to subfertility whereas heterozygous loss of PRM2 does not. Unprocessed PRM2 seems to play a distinct role related to the eviction of intermediate DNA-bound proteins and the incorporation of both protamines into chromatin. For PRM1, on the other hand, heterozygous loss leads to strongly reduced sperm motility as the main phenotype, indicating that PRM1 might be important for processes ensuring correct motility, apart from DNA compaction.

Introduction

Vertebrate reproduction relies on the successful transport of the paternal chromatin to the oocyte by motile sperm cells. The formation of functional sperm requires a unique and well-orchestrated differentiation cascade in the male gonads. Crucial for the successful transport of the paternal DNA is the elongation and condensation of the sperm cell nucleus into a hydrodynamic shape. The necessary size reduction and shielding of DNA required for motility and transit through the reproductive tracts cannot be achieved by histones. Thus, paternal chromatin is completely remodeled from nucleo-histone to nucleo-protamine during spermiogenesis. As a consequence, DNA in mammalian sperm is hypercondensed to an almost crystalline state, considerably reducing the size of nuclei and rendering mature sperm transcriptionally inactive (Chandley & Kofman-Alfaro 1971, Sega 1974, Kierszenbaum & Tres 1975, Pogany et al. 1981, Balhorn 1982). While the sperm chromatin of most mammals is packaged by one protamine, primates and most rodents, as well as subset of other placental mammals express two protamines, protamine 1 (PRM1) and protamine 2 (PRM2).

Both genes are transcribed in round spermatids (steps 7–9 in mice) (Hecht et al. 1986a, b), after which mRNA is stored in the form of cytoplasmic ribonucleoprotein particles until translation in elongating spermatids (steps 12–16) (Kleene et al. 1984, Kleene 1989). Following transcript storage, murine Prm1 seems to be translated earlier (stage 12) than Prm2 (stage 14/15) (Klaus et al. 2016). PRM1 is a small protein (50 aa in mice), containing an arginine-rich DNA binding domain, flanked by short serine- and threonine-containing sequences (Fig. 1). PRM2 is larger (106 aa in mice) than PRM1, containing an arginine-rich DNA-binding domain and serine and threonine residues throughout its sequence (Balhorn 2007, Balhorn et al. 2018) (Fig. 1). Unlike PRM1, PRM2 is expressed as a precursor protein (Yelick et al. 1987). Upon DNA-binding, the N-terminal cleaved-PRM2 domain (cP2) is sequentially cleaved off, leaving only the mature-PRM2 (mP2) domain bound to DNA. Among placental mammals, both protamines contain numerous cysteine residues allowing for the formation of disulfide bridges, stabilizing the condensed chromatin (Balhorn et al. 1992) (Fig. 1). Additionally, PRM2 is able to bind one zinc atom per molecule, allowing for the formation of zinc bridges during epididymal maturation (Bench et al. 2000).

Figure 1
Figure 1

Amino acid alignments of human and mouse PRM1 and PRM2. PRM2 domains cP2 (cleaved-PRM2) and mP2 (mature-PRM2) and DNA binding regions are marked. Amino acids are colored red through blue according to their hydrophobicity value, where red is the most hydrophobic and blue is the most hydrophilic (Geneious Prime v 2019.2.3).

Citation: Reproduction 164, 3; 10.1530/REP-22-0107

Protamines have been discovered almost 150 years ago (Miescher 1874) and their importance for male fertility has been studied extensively. Several outstanding protamine reviews were published, containing detailed information of protamine evolution, biochemistry, function and relation to infertility (Oliva & Dixon 1991, Braun 2001, Oliva 2006, Balhorn 2007, Carrell et al. 2007, Rathke et al. 2014, Balhorn et al. 2018). We refer the reader to these excellent works for details or topics we cannot include here. This review aims to provide an updated summary of our knowledge of protamines and their function as learned from the analysis of different mouse models and related in vitro studies. We will specifically emphasize the differences between PRM1 and PRM2 function.

Box 1 Evolutionary history and selective pressures

Protamines are present in the sperm of both protostomes and deuterostomes (Oliva & Dixon 1991). They most likely developed in chordates from a sperm-specific histone H1 variant via a transition from a lysine-rich histone H1 to an arginine-rich protamine (Lewis et al. 2004). It is assumed that the selective restrictions imposed by internal fertilization have driven the evolution toward an arginine-rich protamine (Kasinsky et al. 2011). Because of its guanidinium group, arginine richness results in a stronger affinity for the protein to attach to the DNA molecule as well as better binding flexibility (Cheng et al. 2003, Ausio et al. 2014). Protamines in eutherian mammals evolved to be quite small compared to other species. Eutherian protamines, unlike other species, include cysteine residues, which allow the formation of complex tertiary structures by forming disulfide bridges between and among protamines (Balhorn 1982, Oliva & Dixon 1991, Balhorn et al. 1995).

Due to their involvement in reproduction, proteins of the male and female germline are often the target of selective pressures driven by sexual selection or male–female coevolution (Swanson & Vacquier 2002, Turner et al. 2008). Comparative evolutionary studies can reveal these pressures allowing us to draw conclusions about the function and importance of reproductive genes and proteins by evaluating the naturally occurring phenotypic diversity and associating it to changes and selective pressures in gene sequences and expression rates. When compared to other sperm nuclear basic proteins, some studies suggest that protamines are the quickest evolving reproductive proteins, with immense structural heterogeneity (Oliva & Dixon 1991, Wyckoff et al. 2000). For Prm1, however, evidence of purifying selection (i.e. selection against detrimental mutations in the coding sequence) focusing on the maintenance of the high arginine content exists. The location of arginine residues appears to be variable, indicating that Prm1 may be impacted by both positive (i.e. selection favoring beneficial mutations in the coding sequence) and purifying selection (Rooney et al. 2000, Lüke et al. 2016a ). Notably, it was found that the arginine content of PRM1 correlated with sperm head size and was maintained by sexual selection across mammals (Lüke et al. 2016a ).

Prm2 on the other hand seems to evolve under less-selective constraints. Studies including primates and rodents, as well as other Prm2 expressing mammalian species surprisingly indicate that the two domains of Prm2 (cP2 and mP2) seem to evolve under different selective pressures. The cP2 domain was shown to be highly conserved, whereas the mP2 domain was found to evolve at a faster rate indicating positive selection in primates and relaxed selective constraint in rodents (Lüke et al. 2011). Interestingly in rodents, changes in the cP2 coding sequence are significantly correlated to sperm head width and elongation (Lüke et al. 2014a). A potential reason for these different selection patterns is that cP2 function needs to be conserved, while mP2, putatively being redundant to PRM1 is free to evolve under less constraint (Lüke et al. 2011, 2016a). Since the cleaving of PRM2 seems to takes place while associated with chromatin and is achieved over a period of several days (Carré-Eusèbe et al. 1991, Chauvière et al. 1992), DNA condensation coincides temporally with PRM2 cleaving and posttranslational processing (Kierszenbaum & Tres 1975, Lee et al. 1995, Brewer et al. 2002). Based on this, Lüke et al. proposed that the cP2 domain may regulate the dynamics of DNA condensation, while the mP2 domain is mainly involved in the maintenance of the hypercondensed state of the mature sperm nucleus (Lüke et al. 2016a, b).

The histone to protamine transition

Chromatin remodeling is preceded by extensive histone modification, especially histone H4 acetylation, orchestrated by the interaction between histone acetyltransferases and histone deacetylases (Candido & Dixon 1972, Meistrich et al. 1992, Hazzouri et al. 2000, Marcon & Boissonneault 2004, Shirakata et al. 2014). Further, canonical histones are replaced by testis-specific variants, such as the histone H1 variant HILS (Yan et al. 2003), the histone H2B variant TH2B (Montellier et al. 2013, Shinagawa et al. 2015) and the histone H2A variant H2A.L1/2 (Govin et al. 2007, Barral et al. 2017). Chromatin remodeling was shown to be aided by various proteins, complexes and chaperones. During the process of chromatin reorganization, topoisomerase II, for example, induces double-strand breaks, that are believed to relieve supercoiled DNA strands and are subsequently repaired (McPherson & Longo 1993, Marcon & Boissonneault 2004, Laberge & Boissonneault 2005). We refer the reader to the excellent review by Rathke et al. for an overview of proteins involved in the processes preceding protamine binding (Rathke et al. 2014). We will concentrate here on proteins that are known or believed to be directly involved with protamine loading, during the final steps of spermiogenesis (e.g., CAMK4, BRDT, TH2B, H2A.L.2 and transition proteins (TNPs)) (Fig. 2). Barral et al. demonstrated that the testis-specific histone variant H2A.L.2, in collaboration with TH2B and TNPs (TNP1 and TNP2), induces structural alterations in chromatin that allow protamines to bind DNA (Barral et al. 2017). Protamines themselves then replace histones and histone–TNP complexes are evicted (Barral et al. 2017). Phosphorylation of protamines was shown to be crucial for correct protamine binding (Pirhonen et al. 1994a, b, Wu et al. 2000, Papoutsopoulou et al. 1999). Once bound to DNA, murine protamines seem to be dephosphorylated thereby ‘locking’ them in place (Fig. 2).

Figure 2
Figure 2

Schematic representation of the histone-to protamine-transition based on current knowledge from mouse models and in vitro studies as discussed in this review. PRM1 in green and PRM2 in yellow.

Citation: Reproduction 164, 3; 10.1530/REP-22-0107

The binding of protamines results in the coiling of DNA into a toroidal structure (Ward & Coffey 1991). The current consensus model of protamine-mediated chromatin condensation states that one protamine molecule binds to each turn of DNA in the major groove, neutralizing the phosphodiester backbone charge and causing the DNA to coil into toroidal structures (Ward & Coffey 1991, Balhorn 2007, Cree et al. 2011). After histone and TNP eviction is completed, during epididymal transit, the protamine–DNA complex is further stabilized by inter- and intra-protamine disulfide bonds and zinc bridges (Calvin & Bedford 1971, Balhorn 1982, Oliva & Dixon 1991, Balhorn et al. 1995, Hutchison et al. 2017) (Fig. 2). The reorganization ultimately achieves a nearly complete silencing of transcription and efficient protection of the paternal chromatin (Oliva & Dixon 1991, Steger 1999).

Box 2 Histone retention

Even though chromatin is reorganized and histones are replaced by protamines, a small fraction of the genome remains packed by histones (1–10% in mice, 5–15% in men), conveying epigenetic information (Balhorn et al. 1977, Hammoud et al. 2009a, 2010, Lesch & Page 2014, Samans et al. 2014, Jung et al. 2017, Yamaguchi et al. 2018). Both canonical and testis-specific histone variants such as TH2B and H2A.L.2 have been shown to be retained in fully condensed sperm chromatin (Kim et al. 1987, Govin et al. 2007, Barral et al. 2017). Studies in human and mouse sperm reported histone retention to be biased toward promoters of genes expressed during early development (Gardiner-Garden et al. 1998, Arpanahi et al. 2009, Hammoud et al. 2009a, Brykczynska et al. 2010, Erkek et al. 2013, Lesch & Page 2014). Other works report that histone retention is enriched in repeat elements such as long and short interspersed nuclear elements or pericentric repeats (Pittoggi et al. 1999, Govin et al. 2007). In addition, histone retention was reported to be particularly concentrated on pericentric heterochromatin, which was proposed to be due to the ability of H2A.L.2 to interact with RNA in these chromatin regions (Hoghoughi et al. 2020). The methodological difficulty of separating protamine and histone-bound fractions from hypercondensed DNA might be the cause of the discrepancies between studies (Carone et al. 2014, Yamaguchi et al. 2018). Recent studies started to use the protamine chaperone nucleoplasmin for selective protamine removal prior to ChIPseq. Using this method, it was shown that H3K4me3 seems to be preferentially retained in CpG-rich promoters and H3K9me3 in satellite repeats (Yamaguchi et al. 2018). Overall, several studies were able to confirm the association of H3K4me3 to CpG-rich promoters of developmentally relevant genes (Hammoud et al. 2009a, Brykczynska et al. 2010, Erkek et al. 2013, Xu et al. 2018, Yamaguchi et al. 2018). Additionally, Lesch and Page described the presence of bivalent histone marks in histones retained around the promoter of developmentally important genes (Lesch & Page 2014). They propose that this epigenetic information may play a role in the preparation for totipotency of the zygote. Moreover, histone retention seems to be involved in maintaining epigenetic information in and around imprinted genes (Brykczynska et al. 2010, Hammoud et al. 2010, 2011, Erkek et al. 2013). How histone retention is regulated during chromatin reorganization is a matter of debate. Acetylation and butyrylation of H4 tails were reported to balance histone retention and removal. According to this model, the binding of BRDT to acetylated H4 allows for the removal of histones by TNPs, while butyrylated H4 escapes replacement (Goudarzi et al. 2014). On the other hand, the selective incorporation of specific histone variants, such as H2A.L.2, which creates an open nucleosome allowing for TNP loading, might determine the histone retention pattern (Govin et al. 2007, Barral et al. 2017). What seems clear is that synchronized histone removal and selective retention involve complex processes and several independent factors (Goudarzi et al. 2014).

The protamine ratio

The essentiality of both protamines for sperm chromatin rearrangement is evident from the extensive investigation using mouse models (Cho et al. 2001, 2003, Mashiko et al. 2013, Takeda et al. 2016, Schneider et al. 2020, Arévalo et al. 2022, Merges et al. 2022). However, it is still unclear why two protamines are required in some species whereas one protamine (PRM1) appears to be sufficient in others. Part of the answer might lie in studying the interaction between protamines and the protamine ratio in detail.

Exactly how protamines interact is purely based on theoretical models (Corzett et al. 2002, Cree et al. 2011). It is however very clear that maintaining the exact ratio between PRM1 and PRM2 expression and abundance in mature sperm is a key factor of proper sperm function (Balhorn et al. 1988, Cho et al. 2001, Steger et al. 2001, 2003, Aoki et al. 2005, Oliva 2006). Importantly, the protamine ratio (PRM1:PRM2) varies widely between mammalian species (0–70% PRM2) while it is almost constant within the same species (Corzett et al. 2002). In men, PRM1 and PRM2 seem to be proportioned almost equally, while PRM2 is slightly higher abundant in mice accounting for 65% of total protamine (Corzett et al. 2002). The standard procedure for the evaluation of the protamine ratio is based on acid-urea gel electrophoresis band density measurements or quantitative RT-PCR in case of measuring the protamine expression ratio. It should be noted, that the gel-based method has certain limitations, such as potential differences in staining, as well as the difficulty in discerning between different proteoforms. In line with this, Soler-Ventura et al. recently showed that the human protamine ratio measured by LC−MS intensities is significantly higher (1.76 ± 0.16) than reported in previous studies using gel-based assays (1.04 ± 0.16) (Soler-Ventura et al. 2020). This certainly warrants further investigation in different species.

Even though the cause and specific implications are still not understood, changes in species-specific protamine ratios can disrupt sperm development and seem to be associated with male infertility even more so than absolute protamine amounts in humans (Balhorn et al. 1988, Steger et al. 2001, 2003). Modifications in the protamine ratio are associated with subfertility and increased DNA fragmentation in humans (Aoki et al. 2005, Oliva 2006) (see also: Box 3). Consistent with these findings, comparative analyses including different rodent species suggest the protamine ratio to be of great importance for sperm competitiveness and the species-specific sperm head phenotype (Lüke et al. 2014b, Arévalo et al. 2021). Earlier studies indicate that species lacking PRM2 or producing very little PRM2 relative to PRM1 show stronger chromatin condensation and decondense DNA more slowly after fertilization (Perreault et al. 1988, Corzett et al. 2002, Gosálvez et al. 2011). A recent study showed that the protamine expression ratio seems to adapt toward a 1:1 proportion under the selective pressure of sperm competition in different rodent clades (Arévalo et al. 2021) (see also: Box 1). If sperm competition-driven selection prefers a clear optimum, other selective factors should be at work to produce such different protamine ratios between species. Aside from condensing the nucleus and shielding paternal DNA, protamine expression levels and the protamine expression ratio may be related to the level of histone retention. And in fact, there appears to be an inverse relationship between the protamine expression ratio and the degree of histone retention in sperm chromatin in men (Hammoud et al. 2009b). Sperm histone retention (Box 2) is important for early embryonic development and potentially for the preservation of parental imprinting (Brykczynska et al. 2010, Hammoud et al. 2011, Yamaguchi et al. 2018). Under these terms, the protamine ratio might reflect a complex interplay between protamines and histones to organize and maintain the correct, species-specific histone retention pattern.

Lessons learned from mouse models

The fact that rodent and primate sperm express two protamines, that show considerable differences in sequence composition and processing, raises four main questions:

  1. Are both protamines required for the production of functional sperm?

  2. Do they perform different functions during chromatin condensation?

  3. What is the role of PRM2 processing?

  4. Do protamines interact with TNPs and other proteins involved in the histone-to-protamine transition?

Protamine-deficient mouse models

Over the last two decades, several protamine-deficient mouse models have been generated using a variety of different approaches (Cho et al. 2001, 2003, Mashiko et al. 2013, Takeda et al. 2016, Schneider et al. 2020, Arévalo et al. 2022, Merges et al. 2022) (Table 1).

Table 1

Overview of mouse models discussed in this review.

Gene/study Genotype Fertility Phenotype
Prm1
 Cho et al. (2001) Chimera Infertile Presence of unprocessed PRM2
 Mashiko et al. (2013) Prm1+/− Infertile Fertilization incompetent sperm, reduced motility
 Takeda et al. (2016) Prm1+/− Infertile Fertilization incompetent sperm (fertilization possible when using IVF with zona-free oocytes)
 Moritz et al. (2021) Prm1K49A/K49A Subfertile Presence of unprocessed PRM2, increased histone retention, reduced motility
 Merges et al. (2022) Prm1+/− Subfertile Presence of unprocessed PRM2, impaired protamination, reduced motility
 Merges et al. (2022) Prm1−/− Infertile Inviable sperm, DNA degradation, increased histone retention, presence of unprocessed PRM2
Prm2
 Cho et al. (2001, 2003) chimera Infertile Impaired DNA condensation, presence of unprocessed PRM2
 Schneider et al. (2016, 2020) Prm2+/− Fertile
 Schneider et al. (2016, 2020) Prm2−/− Infertile Inviable sperm, DNA degradation, increased histone retention
 Itoh et al. (2019) Prm2S56A/S56A Fertile
 Arévalo et al. (2022) Prm2+/Δc Infertile Inviable sperm, DNA degradation, TNP retention, failed H2A.L.2 retention, reduced protamine content
 Arévalo et al. (2022) Prm2-/Δc Infertile Inviable sperm, DNA degradation, TNP retention, failed H2A.L.2 retention, reduced protamine content
Brdt
 Shang et al. (2007) BrdtΔBD1/ΔBD1 Infertile Abnormal sperm, increased histone retention
 Gaucher et al. (2012) BrdtΔBD1/ΔBD1 Infertile Failed TNP and protamine incorporation, increased histone retention
Th2b
 Montellier et al. (2013) Th2b−/− Fertile
 Montellier et al. (2013) Th2b+/tag Infertile Developmental arrest at condensing spermatid stage
H2A.L.2
 Barral et al. (2017) H2A.L.2−/− Infertile Impaired TNP incorporation, presence of unprocessed PRM2
Tnp1
 Yu et al. (2000) Tnp1−/− Subfertile Incomplete DNA condensation, presence of unprocessed PRM2, altered protamine ratio
Tnp2
 Yu et al. (2000) Tnp2−/− Subfertile Incomplete DNA condensation, presence of unprocessed PRM2
Tnp1/Tnp2
 Shirley et al. (2004) Tnp1+/−/Tnp2+/− Fertile
 Zhao et al. (2004) Tnp1−/−/Tnp2−/− Infertile Defects in testicular sperm release, aberrant DNA condensation, presence of unprocessed PRM2
Camk4
 Wu et al. (2000) Camk4−/− Infertile Failed PRM2 deposition, TNP retention
nGpx4
 Pfeifer et al. (2001) nGPx4−/− Fertile Delayed nuclear compaction
 Conrad et al. (2005) nGPx4−/− Fertile
 Puglisi et al. (2014) nGPx4−/− Fertile Interaction with prms

First, Cho et al. deleted Prm1 and Prm2 coding regions by classical gene targeting, i.e. the transfection of a construct encoding for PGK-neo cassette flanked by DNA sequences homologous to the Prm1 and Prm2 sequence (Cho et al. 2001, 2003). Using homologous recombination, they generated Prm1+/− and Prm2+/− embryonic stem (ES) cells, which upon blastocyst injection produced several male and female chimeras. Interestingly, none of the chimeras were able to generate ES-derived offspring. However, ES cells seemed to have contributed to the development of the germ line, since sperm derived from the ES cells was detected. Cho et al. reported that chimeras with higher amounts of ES-derived sperm displayed more morphologically abnormal sperm with lower chromatin integrity. The researchers concluded that the DNA in the protamine deficient ES cell-derived sperm was incompletely condensed and therefore labile or degraded. Chimeric Prm1 and Prm2 males with larger amounts of ES-derived sperm showed increased amounts of unprocessed PRM2 in mature sperm. Cho et al. showed that chromatin structure was altered in ES-derived sperm and hypothesized that this causes the defects in PRM2 processing. The authors concluded that haploinsufficiency of PRM1 and PRM2 renders male mice infertile (Cho et al. 2001).

More than a decade later, however, Schneider et al. were able to establish a PRM2-deficient mouse line (Schneider et al. 2016). Contradicting the earlier results based on PRM2-deficient chimeras (Cho et al. 2001, 2003), Prm2+/− mice were reported to be fertile (Schneider et al. 2016). Schneider et al. hypothesized that expression of the neo-cassette in the mouse model published by Cho et al. resulted in trans-silencing of the nearby Prm1 gene (Schneider et al. 2016). In fact, such trans silencing had been described by Olson et al. who demonstrated that depending on the orientation of the PGK-neo cassette inserted to disrupt the MRF4 locus the neighboring gene Myf5 was silenced as well (Olson et al. 1996). In the study by Cho et al., sperm of chimeric PRM2-deficient mice showed a significant reduction in PRM1 level supporting the hypothesis of trans-silencing (Cho et al. 2001, 2003). Since Schneider et al. reported Prm2+/− males to be fertile, Prm2−/− mice could be generated and analyzed.

Prm2−/− mice are infertile (Schneider et al. 2016). Observed defects in DNA hypercondensation and increased DNA fragmentation correlate well with the described function of protamines and data from infertile men with impaired protamination. Additionally, PRM2-deficient murine sperm displayed disintegrated membranes, decreased sperm head size and viability as well as acrosomal defects (Schneider et al. 2016, 2020) (Fig. 3). Additionally, Prm2−/− sperm were completely immotile. This is however clearly due to the fact that the sperm cells are inviable and degrading. Similar morphological malformations were reported for Prm2 chimeric animals (Cho et al. 2001, 2003) and in patients with abnormal protamine loading (Depa-Martynow et al. 2012, Rogenhofer et al. 2013, Iranpour 2014). Since these observed defects do not seem to be directly related to protamine function, they are most likely secondary defects (Schneider et al. 2016). Later studies revealed that Prm2−/− mature sperm showed a significant increase in histone retention (Arévalo et al. 2022, Merges et al. 2022). Proteomic profiling of PRM2-deficient sperm revealed a strong decline in reactive oxygen species (ROS) scavenger proteins SOD1 (superoxide dismutase 1) and PRDX5 (peroxiredoxin 5) (Schneider et al. 2020). Well-balanced ROS levels are essential for fertility, as low levels are required for sperm capacitation, acrosome reaction and sperm maturation. In contrast, overshooting levels of ROS result in oxidative DNA damage, lipid peroxidation and apoptosis (Tremellen 2008, Aitken et al. 2012). Thus, lower levels of SOD1 and PRDX5 result in the rise of ROS levels, causing DNA fragmentation, lipid peroxidation and the downstream secondary effects seen in PRM2-deficient mice (Schneider et al. 2020). These observations support the hypothesis that the degree of protamination might serve as a proxy for chromatin integrity assessed during epididymal transit as discussed previously (Carrell et al. 2007, Steger & Balhorn 2018). According to this hypothesis, faulty DNA protamination initiates a self-destruction cascade of sperm. This process seems to be initiated during epididymal transit in these mice since oxidative DNA damage and impaired acrosome status were not observed in PRM2-deficient testicular spermatids (Schneider et al. 2020). This opens up new possibilities for treatment regimens for infertile men if the proposed molecular processes can be transferred from the mouse model to the human situation.

Figure 3
Figure 3

Schematic summary of phenotypic effects in different protamine mouse models in relation to the observed changes in the protamine ratio. Black blocks indicate the average level of the measured parameter for the different genotypes. Fertility indicates a summary of the level of fertility from infertile to fertile taken from measures of average litter size and pregnancy rate of the genotype. DNA fragmentation indicates the average percentage of mature sperm found to have fragmented DNA according to transmission electron microscopy, agarose gel electrophoreses analysis and oxidative DNA damage marker staining of the genotype. Sperm motility indicates the average percentage of motile mature sperm for each genotype. PRM2-precursor/TNP/Histone retention summarizes the amount of PRM2 precursor/TNP/histone (H3 and H4) found in epididymal sperm in the different genotypes compared to WT levels. Data were taken from the studies indicated.

Citation: Reproduction 164, 3; 10.1530/REP-22-0107

In a CRISPR/Cas methodology-focused study, Mashiko et al. generated Prm1+/− mice, which were reported to be infertile (Mashiko et al. 2013). However, very limited fertility testing was performed in which Prm1+/− male founders were mated with superovulated females and oocytes were isolated. Formation of pronuclei, which was considered a successful fertilization event, was not detected for Prm1+/− sperm. The authors reported that sperm of Prm1+/− males show morphological abnormalities and reduced motility.

In order to overcome the infertility observed in male PRM1-deficient animals, Takeda et al. (2016) targeted Prm1 in TT2-XO ES cells and generated female chimeras which transmitted the Prm1 deletion to the germline. Prm1+/− males were infertile, with similar testis weight compared to WT, suggesting that sperm production was not affected (Takeda et al. 2016). Prm1+/− sperm showed reduced/defective protamination, unprocessed PRM2 in mature sperm, reduced DNA integrity and various head and tail deformities (Takeda et al. 2016). Tail malformations were linked to disorganization of microtubules in the sperm tail and a reduced mitochondrial membrane potential resulting in a severe reduction of sperm motility. Further, Prm1+/− sperm showed accelerated acrosome reactions, which might contribute to the reduction in motility. Comparison of the expression profiles of Prm1+/− to WT cauda epididymal sperm revealed only little differences, indicating that transcriptional silencing was not disturbed in Prm1+/− sperm. Surprisingly, however, in vitro fertilization (IVF) of zona-free oocytes with sperm from Prm1+/− males generated offspring, suggesting that infertility of Prm1+/− males was rather caused by secondary effects on acrosome integrity and sperm motility.

Contrary to previous studies (Cho et al. 2001, Mashiko et al. 2013, Takeda et al. 2016), a recent study by Merges et al. generating PRM1-deficient mice by CRISPR/Cas9-mediated gene editing showed that Prm1+/− males are able to produce offspring by natural breeding (Merges et al. 2022). Prm1+/− males were, however, subfertile showing reduced average litter sizes and pregnancy frequencies. In Prm1+/− epididymal sperm, minor ROS-mediated DNA damages were detected. Epididymal sperm appeared condensed and only slight DNA fragmentation was detected. Confirming results from previous studies, Prm1+/− sperm showed a severe reduction in motility. Additionally, Merges et al. showed that the protamine ratio was shifted in Prm1+/− sperm (from 1:2 to 1:5). In accordance with the previously reported mouse models, protamination of mature sperm seemed to be impaired and unprocessed PRM2 was abundantly detected (Takeda et al. 2016, Merges et al. 2022) (Fig. 3).

The subfertility of Prm1+/− males allowed for breeding of Prm1−/− males, which were infertile. Testis weight and morphology, as well as the number of sperm produced, were comparable in Prm1−/−, Prm1+/− and WT mice. In Prm1−/− testis, however, reactive oxygen species (ROS)-mediated DNA damages were detectable from stage 15–16 testicular spermatids, earlier than reported for Prm2−/− mice by Schneider et al. (2016, 2020). Prm1−/− sperm damage intensified during epididymal transit, and the majority of DNA isolated from Prm1−/− cauda epididymal sperm was fragmented. Prm1−/− cauda epididymal sperm head shapes were altered, being generally smaller and thinner compared to Prm1+/− and WT sperm (Fig. 3). Additionally, Prm1−/− epididymal sperm showed disrupted membranes, damaged acrosomes and were therefore completely immotile. Further, similar to Prm2−/− males, histone retention was increased in Prm1−/− mature sperm and slight TNP retention was detected (Arévalo et al. 2022, Merges et al. 2022) (Fig. 3).

These recently established protamine mouse lines will allow for the generation of Prm1+/− Prm2+/− males. In principle, sperm of mice, lacking one allele of Prm1 and Prm2, would contain only 50% total amount of protamine but could be expected to display an unaltered protamine ratio. Such animals could answer the question of whether male mice remain fertile when the total amount of protamine is reduced, while the correct protamine ratio is maintained.

Protamine-modified mouse models

Great advancements have been made in describing the role of protamines and their function, yet the role of posttranslational processing and posttranslational modification (PTM) of protamines is just starting to be understood. Several recent studies based on mouse models have started to tackle these questions (Table 1).

A mouse model carrying a deletion of the cP2 domain has been recently established by Arévalo et al. addressing the role of PRM2 processing (Arévalo et al. 2022). Due to a targeted deletion of the cP2 domain, Prm2+/Δc mice express only mP2 from the gene-edited allele in addition to full Prm2 from the WT allele. Surprisingly, in contrast to Prm2+/−, male Prm2+/Δc mice were infertile. In order to test if mice expressing only Prm2Δc would be fertile, Prm2−/Δc mice were generated. However, Prm2−/Δc males were also infertile, demonstrating that the infertility of Prm2+/Δc males was not solely caused by an aberrant interaction with full-length PRM2 expressed from the WT allele. Since Prm2+/− males were fertile, the infertility of Prm2−/Δc males originates from the lack of the cP2 domain, demonstrating its essential role in male fertility. DNA of sperm from Prm2+/Δc and Prm2−/Δc males was degraded starting in condensed testicular spermatids and PRM2 incorporation into the DNA was reduced during late spermiogenesis. Arévalo et al. hypothesize that this was caused either by the prevention of nuclear localization of PRM2 or secondary eviction from the nucleus. Using a cP2-specific antibody, it was shown that in WT, unprocessed PRM2 can be found in the cytoplasm of condensed spermatids and in residual bodies indicating that significant amounts of PRM2 are not processed in the WT (Fig. 2). cP2 staining was absent in the nuclei of condensed spermatids, while the mP2-specific signal was present in the nuclei. In Prm2+/Δc mice, however, PRM2 precursor was detected in the nuclei of condensed spermatids and in epididymal sperm (Arévalo et al. 2022). These results indicate that what is interpreted as ‘incomplete PRM2 processing’ in many of the protamine-related mouse lines might actually be failed PRM2 precursor eviction or PRM2 precursor retention. This warrants further testing.

In contrast to Prm2−/− mice, failure of transcriptional silencing was not apparent in cP2-deficient mice (Schneider et al. 2020). However, Pcgf5 (polycomb group ring finger 5) was shown to be highly expressed in cP2-deficient testes (Arévalo et al. 2022). Pcgf5 is part of the polycomb repressive complex 1 (PRC1) mediating heritable histone modifications (Piunti & Shilatifard 2021). The involvement of cP2 with PRC1 is not clear yet and merits further investigation.

Interestingly, loss of the cP2 domain lead to a strong alteration of the protamine ratio (2:1–5:1), showing a lower percentage of PRM2 (including precursors) compared to the WT. Additionally, the total amount of protamine is severely reduced in epididymal sperm. Thus, not only PRM2 but also PRM1 failed to correctly incorporate into chromatin in cP2-deficient mice. As mentioned earlier, large amounts of unprocessed PRM2 can be detected in mature sperm of PRM1-deficient mice (Merges et al. 2022). The PRM2 precursor therefore seems to be tightly connected to PRM1 incorporation, while the presence of PRM1 is a prerequisite for correct PRM2 processing or PRM2 precursor localization.

Remarkably, cP2-deficient epididymal sperm showed a strong retention of TNP1, but no generalized increase in histone retention. However, in cP2-deficient condensing spermatids, the histone variant H2A.L.2, which is critically involved in the histone-to-protamine transition, seemed to be localized in aberrantly shaped heterochromatin foci in the nucleus. In epididymal sperm of cP2-deficient males, H2A.L.2 was significantly lower abundant compared to WT. Thus, the cP2 domain seems to be directly involved in TNP removal and H2A.L.2 retention and is required for full sperm chromatin protamination (Arévalo et al. 2022) (Fig. 2). In line with these results, a recent study using a different cP2 antibody demonstrated that TNPs and the PRM2 precursor were consistently co-detected suggesting that TNPs and PRM2 do not act successively as previously assumed but simultaneously during histone-to-protamine transition (Rezaei-Gazik et al. 2022). In fact, Arévalo et al. were able to show that the cP2 domain tagged with eGFP was able to pull down untagged TNP1 when co-expressed in human embryonic kidney cells (HEK293). This suggests that cP2 might directly interact with TNP1 in vivo. Further, they proposed that the comparably low TNP retention found in PRM1-deficient sperm might be related to the strong presence of unprocessed PRM2, rather than a direct effect of PRM1 deficiency (Arévalo et al. 2022).

PTMs of histones like acetylation, methylation and phosphorylation are well-known epigenetic modifications that represent a histone code, which strongly affects chromatin structure and gene expression (Jenuwein & Allis 2001). PTMs have been also discovered on sperm histones as well as protamines. Phosphorylation was shown to be crucial for protamine DNA binding (Louie & Dixon 1972, Gusse et al. 1986, Pirhonen et al. 1994a, b). Brunner et al. discovered seven PTMs for murine PRM1 and four PTMs for murine PRM2, including serine and threonine phosphorylation, serine and lysine acetylation as well as lysine methylation marks (Brunner et al. 2014). A recent study by Soler-Ventura et al. provides a very detailed analysis of protamine PTMs and isoforms using a combined mass spectrometry-based approach. Remarkably, they were able to describe novel truncated PRM1 and unprocessed PRM2 proteoforms in humans, including a wide range of PTMs (Soler-Ventura et al. 2020). It is a matter of debate whether the sum of these modifications in protamines represents a protamine code which contributes to epigenetic intergenerational inheritance (Brunner et al. 2014, Castillo et al. 2015; reviewed in Blévec et al. 2020). The molecular function of these PTMs is largely unknown. Two recent studies revealed that dephosphorylation of PRM2 at serine 56 and acetylation of PRM1 at lysine 49 are essential for sperm maturation and function (Itoh et al. 2019, Moritz et al. 2021 (preprint)). Itoh et al. showed that serine 56 can be dephosphorylated by the testis- and sperm-specific phosphatase PPP1CC2. The phosphatase itself is a substrate of the HSP70S chaperone family. Presence of the co-chaperone HSPA4L, which acts as a nucleotide exchange factor, was proposed to support the release of PPP1CC2 from HSP70S, thereby allowing PPP1CC2 translocation to the nucleus and subsequently serine 56 dephosphorylation. Knockout of HSPA4L resulted in male infertility and impaired spermiogenesis with round-headed sperm, which were unable to initiate embryonic development following IVF or ICSI. Accordingly, the level of sperm chromatin-associated PPP1CC2 was decreased in HSPA4L-deficient sperm. However, the exchange of serine 56 to alanine (S56A) was able to rescue the phenotype of HSPA4L-deficient mice. Prm2S56A/S56A mice themselves were fertile and displayed no gross abnormalities. Thus, serine 56 and its PTM seem to be dispensable for fertility and sperm function. However, if present, the dephosphorylation of serine 56 is a prerequisite for sperm maturation during spermiogenesis (Itoh et al. 2019). Like murine protamines, human protamines undergo posttranslational phosphorylation and dephosphorylation (Pruslin et al. 1987, Chira et al. 1993). However, it remains to be investigated whether results from the mouse model can be associated with human male infertility as well.

In a recent study published as a preprint, Moritz et al. set out to investigate the significance of evolutionarily conserved non-arginine residues in protamine-DNA binding (Moritz et al. 2021 (preprint)). It is believed that the protamine-DNA binding dynamic relies primarily on passive electrostatic interaction with the arginine-rich binding domains (Gou et al. 2020). However, it was shown that PTMs of protamines, especially phosphorylation plays an additional role (Pirhonen et al. 1994b, Green 2001, Seligman et al. 2004). Moritz et al. found that these PTM sites seemed to be conserved within species but can vary between (Moritz et al. 2021 (preprint)). Focusing on PRM1 lysine 49 (P1 K49), which was found to be acetylated and conserved across rodents, mutant mice carrying a substitution of P1 K49 with alanine (K49A) were generated and shown to be subfertile in the homozygous state. While spermatogenesis and sperm count seemed to be unaffected, mature sperm show head morphology and midpiece morphological abnormalities and progressive motility of sperm seemed to be severely reduced. The mutation did not affect the levels of PRM1 and PRM2 or the protamine ratio, however unprocessed PRM2 was accumulated in mature sperm. Mutant sperm show histone, but no TNP retention. They concluded that PRM1 K49 acetylation might be necessary for the proper binding of PRM1 and successful eviction of nucleosomes. They hypothesize that the acetyl-lysine reader BRDT might be involved in assessing the lysine acetylation levels of protamines, including PRM1 K49 acetylation, and thereby in modulating the chromatin reorganization (Moritz et al. 2021 (preprint)).

Of note, the deletion of Prm1, Prm2 or the cP2 domain resulted in deviations from the WT protamine ratio of 1:2 (Schneider et al. 2016, 2020, Arévalo et al. 2022, Merges et al. 2022). In Fig. 3, we summarize the main phenotypes of the protamine mouse models described above in relation to the observed protamine ratio. Taken together, the phenotypes that were described for these different mouse models indicate that a shift in the ratio toward PRM2 might be more easily tolerated than a relative excess of PRM1 in mice (Fig. 3).

Protamine-interactor mouse models

During the post-meiotic phases, global chromatin hyperacetylation initiates the reorganization of paternal chromatin. The testis-specific bromodomain-containing protein BRDT not only plays an important role during the meiotic stages of spermatogenesis but is also involved in the initiation of chromatin hypercondensation (Fig. 2). BRDT’s bromo-domain 1 (BD1) reads acetyl-lysine marks and binds to acetylated histone H4. Thereby, it is proposed to drive the genome-wide replacement of histones with TNPs (Pivot-Pajot et al. 2003, Govin et al. 2004, Dhar et al. 2012). Different mouse models have been established to study the function of BRDT, but the deletion of BD1 specifically demonstrated its role during chromatin reorganization. In male mice lacking BD1 (Shang et al. 2007, Gaucher et al. 2012), both TNPs and protamines while detectable seem not to be incorporated into the chromatin of spermatids (Gaucher et al. 2012). Thus, BD1 seems to be instrumental for the eviction and replacement of histones.

During chromatin reorganization, canonical histones are replaced by testis-specific variants. Two of these variants, the histone H2B variant TH2B (Montellier et al. 2013, Shinagawa et al. 2015) and the histone H2A variant H2A.L.2 (Govin et al. 2007, Barral et al. 2017) play important roles during histone-to-protamine transition (Fig. 2). It was shown that H2A.L.2 and TH2B are prone to forming dimers (Govin et al. 2007).

TH2B, first described in 1975 (Branson et al. 1975, Shires et al. 1975), is expressed in meiotic spermatogonia where it gradually and completely replaces canonical H2B (Montellier et al. 2013). The role and function of TH2B were investigated using KO mouse models as well as a model in which TH2B was C-terminally tagged (Montellier et al. 2013). Strikingly, TH2B-deficient mice were fertile and other H2B variants were upregulated to compensate for the lack of TH2B. Although the loss of TH2B causes dramatic changes in core histone marks, it did not seem to affect litter sizes or health of the offspring. The mice harboring the C-terminally tagged TH2B on the other hand were infertile. In these mice, germ cell development was arrested in condensing spermatids, at a time when TNPs and protamines should have completely displaced TH2B. This suggests that correct interaction of TH2B, or compensatory H2B variants with H2A.L.2, TNPs and/or protamines is necessary during the final stages of spermiogenesis.

H2A.L.2, described by Govin et al., is expressed and incorporated into DNA during late spermiogenesis, at the time of histone replacement and detected at the same time as TNPs (Govin et al. 2007, Barral et al. 2017). H2A.L.2-deficient mice are infertile but showed normal testis histology and sperm count. TNPs and protamines seemed to be present at normal levels. However, H2A.L.2 deficiency leads to incomplete loading of TNPs and similar to TNP deficiency accumulation of PRM2 precursor. Co-immunoprecipitation analyses further support an interaction between H2A.L.2, TH2B, TNP2 and PRM2. Thus, they conclude that H2A.L.2 integration into chromatin initiates a series of transitional phases eventually leading to the incorporation of protamines and the hypercondensation of the paternal DNA.

TNPs are likely to closely interact with protamines during protamine loading and final histone eviction. They are arginine- and lysine-rich with a strong affinity to binding DNA (Kistler et al. 1975, Heidaran & Kistler 1987, Kleene & Flynn 1987, Brewer et al. 2002). The function of these proteins is still not completely clear. It is believed that TNPs cooperate with topoisomerase II, rendering the DNA more flexible (Akama et al. 1998, Lévesque et al. 1998, Singh & Rao 1988, Meistrich et al. 2003), condense chromatin (Baskaran & Rao 1990, Brewer et al. 2002) and stimulate the repair of DNA breaks (Caron et al. 2001). Both TNP1- and TNP2-deficient mouse models have been generated and analyzed (Yu et al. 2000). Tnp1−/− and Tnp2−/− male mice were subfertile. TNPs seem to be able to partly compensate for each other since in Tnp1−/− and Tnp2−/− males only slight sperm abnormalities were described (Yu et al. 2000, Zhao et al. 2001). Both, Tnp1−/− and Tnp2−/− males showed abnormal chromatin condensation starting in step 11 spermatids, forming condensed foci. Additionally, chromatin condensation was incomplete in epididymal Tnp2−/− sperm. While histone eviction was not disrupted in Tnp1−/− or Tnp2−/− sperm, unprocessed PRM2 was detected in mature sperm. In epididymal sperm of Tnp1−/− males, the relative protamine ratio was shifted to 1:3 (including unprocessed PRM2) but remains unchanged in Tnp2−/− sperm. Strikingly, in Tnp1−/− spermatids, PRM2 seemed to be translated prematurely in part. If this premature PRM2 translation represents a compensatory mechanism is still unclear.

Tnp1−/−Tnp2−/− double KO mice were infertile, with reduced epididymal sperm counts correlating with defects in testicular sperm release (Zhao et al. 2004). The majority of Tnp1−/−Tnp2−/− sperm were inviable and immotile showing a severe head and tail deformations and aberrant chromatin condensation. Again, histone removal was unaffected by TNP loss. Both protamines were effectively deposited, but large amounts of PRM2 precursor were detected in epididymal sperm. Haploinsufficiency of one TNP while the other is absent caused similar phenotypical aberrations as the double KO (Shirley et al. 2004). Mice heterozygous for both TNPs, on the other hand, were fertile showing no obvious sperm defects again demonstrating the compensation.

So, in general, TNPs are not required for the deposition of protamines but the loss of TNPs seems to be connected to the accumulation of unprocessed PRM2 in epididymal sperm. In fact, Barral et al. were able to show that TNP2 is able to pull down both processed and unprocessed PRM2 in vitro (Barral et al. 2017). Together with the results from cP2-deficient mice, this suggests that the PRM2 precursor is closely connected to transition protein turnover. Unprocessed PRM2 being necessary for transition protein eviction while the presence of correct levels of TNPs is needed to prevent the accumulation of PRM2 precursor in the nucleus. The same relationship can be observed between PRM1 presence/incorporation and the PRM2 precursor.

Ca2+/calmodulin-dependent protein kinase IV (CAMK4) is a serine/threonine protein kinase, which is expressed in spermatids and associated with chromatin and the nuclear matrix (Wu et al. 2000). CAMK4−/− mice were infertile with spermiogenesis being arrested during nuclear condensation (stage IV). The transcript levels of TNPs and protamines and translational timings were not affected in CAMK4-deficient sperm. However, while PRM1 was detected throughout spermiogenesis, PRM2 was undetectable from step 15 spermatids onward. TNP2 on the other hand showed prolonged retention. This is congruent with the results from PRM2 and cP2-deficient mice indicating that PRM2 is required for TNP eviction. Wu et al. showed that CAMK4 phosphorylates PRM2 (in vitro) during or after proteolytic processing. Thus, CAMK4 seems to be involved in PRM2 phosphorylation, proper PRM2 deposition and consequently TNP removal (Wu et al. 2000).

Following histone and TNP eviction, during epididymal maturation, the sperm chromatin complex is stabilized by disulfide bonds and zinc bridges (Calvin & Bedford 1971, Balhorn 1982, Oliva & Dixon 1991, Balhorn et al. 1995). The sperm nuclear isoform of glutathione peroxidase 4 (nGPx4) contains an N-terminal arginine-rich sequence, which mediates nuclear localization and chromatin binding (Pfeifer et al. 2001, Moreno et al. 2003). Pfeifer et al. (2001) proposed nGpx4 to exhibit protamine thiol peroxidase activity that aids in stabilizing the hypercondensed chromatin by cross-linking protamine cysteine residues (Fig. 2). Conrad et al. showed that nGPx4-deficient mice were fertile but exhibit a delay in the completion of epididymal nuclear compaction (Conrad et al. 2005). Thus, while nGPx4 did seem to be involved in cross-linking protamine disulfides, the lack of this variant was eventually compensated for, possibly by other glutathione peroxidases or canonical GPx4 (Conrad et al. 2005). Further analyses by Puglisi et al. revealed that sperm from nGPx4-deficient mice decondensed faster than WT sperm in vitro, further suggesting that nGPx4 is involved in sperm chromatin compaction. They later showed that nGPx4 localized to the nuclear matrix and seemed to directly interact with protamines (Puglisi et al. 2014).

Many of the main factors orchestrating the histone-to-protamine exchange have been identified and studied (Table 1). The detailed interaction dynamics and the role of these complexes in regulating histone retention and modification, however, are still unclear and require further investigation (Fig. 2).

Questions answered

The extensive study of protamines and their interactors over decades allowed us to at least partly answer the four key questions posed at the beginning of this chapter.

  1. Are both protamines required for the production of functional sperm?

    Indeed, both protamines are required for the production of functional sperm in mice. Prm1−/− and Prm2−/− mice showed strongly aberrant sperm phenotypes and infertility. In both cases, DNA was fragmented and oxidative damage was strongly apparent rendering sperm inviable.

  2. Do they perform different functions during chromatin condensation?

    The results gathered from the different protamine-deficient mouse models do indicate that PRM1 and PRM2 perform distinct functions during spermiogenesis. First, strong TNP retention was only observed in PRM2 mouse models, while none of the published PRM1 mouse models reported more than weak TNP retention. Thus PRM2, specifically its cleaved domain, seems to be instrumental in TNP eviction from DNA/chromatin of sperm. Additionally, loss of one allele of Prm1 leads to subfertility whereas loss of one allele of Prm2 did not, and strikingly Prm1+/− mice showed a significant reduction in sperm motility but only slight DNA damage. Considering that the translation of Prm1 precedes Prm2, different hypotheses on the roles of these proteins can be posted. First, PRM1 could act as a vanguard preparing the field for PRM2. If PRM1 levels are low, this does not happen properly, allowing for an PRM2 ‘excess’ which might cause further problems, such as binding to DNA in regions it should not. Second, PRM1 might have a unique yet unknown function other than binding and condensing DNA affecting sperm motility.

  3. What is the role of PRM2 processing?

    While the role of PRM2 processing and the cP2 domain is still not completely clear, the picture that emerges from studying cP2-deficient mouse lines points toward a role in orchestrating the eviction of TNPs, possibly their transport out of the nucleus and the regulation of the H2A.L.2 nucleosome. Recent results indicate that not all PRM2 is processed and that the precursor itself might perform a different function than the processed form. Additionally, these results lead to the hypothesis that only PRM2 that is bound to DNA is processed while the PRM2 precursor engaged in other processes is not cleaved and remains in cytoplasm and residual bodies. PRM2 processing or PRM2 precursor localization seems to depend on the tightly regulated balance of various proteins (H2A.L.2, TNP1, TNP2, PRM1 and PRM2 itself).

  4. Do protamines interact with TNPs and other proteins involved in the histone-to-protamine transition?

    Cell culture assays revealed that TNP2 was able to interact with processed and unprocessed PRM2 via the mP2 domain, but interestingly not the cP2 domain. Additionally, initial results indicated that the cP2 domain was able to interact with TNP1. nGPX4 was shown to directly interact with protamines during chromatin cross-linking. Further interactions remain to be elucidated.

Lessons learned from in vitro studies

The complexity of spermiogenesis and the difficulty of working with and disentangling the different components of the hypercondensed chromatin restricts the experimental approaches for a detailed analysis of protamine function in vivo. Therefore, in vitro-based methods are increasingly helpful. Key questions that can be answered in vitro are the following:

  1. Is an interaction with TNPs and testis-specific histones necessary for DNA condensation by protamines?

  2. Does DNA binding affinity and DNA condensation efficiency differ between PRM1 and processed and unprocessed PRM2?

Iuso et al. demonstrated in sheep fibroblasts that mouse PRM1 binds to DNA and condenses nuclei when expressed in vitro (Iuso et al. 2015). PRM1 accumulated as nuclear speckles, which grow in size until the nucleus is condensed into a stick-like shape. Of note, they were able to inject these spermatid-like (2N) nuclei into enucleated oocytes where the DNA underwent decondensation, maternal histones were incorporated and development to blastocyst stage was reported (Iuso et al. 2015, Czernik et al. 2016). The ability to condense somatic nuclei in vitro has since been shown for mouse PRM2 and the mP2 domain as well (Arévalo et al. 2022).

In their recent preprint, Moritz et al. extensively evaluated differences between mouse PRM1 and PRM2, as well as unprocessed PRM2 and mutated PRM1 (K49A) in DNA binding affinity and compaction efficiency in vitro using electrophoretic mobility shift assays and DNA curtain assays (Moritz et al. 2021 (preprint)). Their results indicated that the mutated PRM1 had a significantly lower DNA binding affinity. While WT PRM1 selectively bound mP2 rather than the unprocessed PRM2, the mutated PRM1 lost this selectivity. Additionally, they were able to show that mP2 had a higher DNA binding affinity and compacted DNA strands faster than unprocessed PRM2 in these assays. DNA binding affinity was similar between PRM1 and mP2, but mP2 compacted DNA significantly faster and to a denser state than PRM1. Jointly and in the WT ratio of 1:2, PRM1 and PRM2 bound DNA more efficiently. However, PRM1 together with unprocessed PRM2 showed a much lower binding affinity (Moritz et al. 2021 (preprint)). This argues for a role of cP2 in reducing the pace of DNA condensation. This could serve to minimize the DNA damage occurring due to torsional stress. PRM2 processing-based slowing down of the condensation rate could in principle allow the double-strand repair system more time to fix DNA damage.

Earlier studies, mimicking DNA compaction by salmon protamines and arginine-rich peptides in vitro, showed that the rate of DNA condensation depended on the amount of protamine binding to the chromatin (Brewer et al. 1999) and that the stability of toroid structures formed was dependent on the arginine content of the protamine or peptide used (Brewer et al. 1999, Balhorn et al. 2000). Additionally, in vitro experiments using Syrian hamster protamines determined that the condensation and decondensation rates of PRM1 and PRM2 were similar (Brewer et al. 2002). Addition of zinc, however, increased the DNA condensation rate of PRM2 nearly three-fold, while it did not influence PRM1-mediated DNA condensation. Additionally, Björndahl and Kvist showed that disruption of zinc salt bridges was easier than dissolving disulfide bridges, which could result in faster decondensation of paternal chromatin after fertilization (Bjorndahl & Kvist 2014). Thus, the presence of zinc bridges in species expressing both protamines could be an explanation for the differences in the extent of DNA compaction between species expressing PRM1 and PRM2 as opposed to PRM1 only (Gosálvez et al. 2011).

Questions answered

Although in vitro-based studies of protamines and the histone-to-protamine transition are still ongoing and have great potential for enhancing our knowledge of these processes in the future, the questions posed can be partly answered:

  1. Is an interaction with TNPs and testis-specific histones necessary for functional DNA condensation by protamines?

    At least for mouse PRM1 in sheep fibroblasts the answer seems to be no, PRM1 does not need to interact with TNPs or sperm-specific histones in order to condense DNA into a state that is apt for inducing embryo development until the blastocyst stage. Although mouse PRM2 and mP2 are able to condense free DNA molecules and somatic cell chromatin it has not been shown yet if the DNA is sufficiently shielded and intact to support embryo development.

  2. Does DNA binding affinity and DNA condensation efficiency differ between PRM1 and processed and unprocessed PRM2?

    Depending on the species and experimental conditions (presence of zinc) DNA binding affinity and condensation efficiency does seem to differ between protamines and processed or unprocessed PRM2. Processed PRM2 seemed to condense DNA faster than PRM1. Unprocessed PRM2 seemed to have a comparably lower binding affinity to DNA and condensed DNA at a slower pace compared to processed PRM2. Thus, the cP2 domain could be necessary for controlling the pace of DNA compaction. Interestingly, processed PRM2 seemed to also have a stronger affinity to binding PRM1. Considering that both PRM1 and PRM2 abundance is reduced in cP2-deficient sperm this could indicate that mP2 and PRM1 intercept each other before binding DNA. This would suggest that cP2 might be able to block premature interaction with PRM1, congruent with the hypothesis that PRM2 is only processed when bound to DNA.

Box 3 Protamines and infertility in humans

Protamines have been studied extensively in light of their involvement in human male sub- and infertility (Oliva 2006, Carrell et al. 2007). Aberrant sperm protamine content affected sperm motility, morphology and sperm number in men (Aoki et al. 2005). The presence of PRM2 precursor in epididymal sperm was associated with decreased DNA integrity and sperm dysfunction (Yebra 1998, Torregrosa et al. 2006). DNA damage, in turn, leads to decreased survival of offspring (Ruiz-López et al. 2010). Several studies have reported sub- or infertility related to SNPs in protamine sequences (Tanaka et al. 2003, Aoki et al. 2006a, Ravel et al. 2007, Gázquez et al. 2008, Imken et al. 2009, Tüttelmann et al. 2010, Jodar et al. 2011, Yu et al. 2012). Mutations or polymorphisms of protamines were reported to induce conformational changes and affect incorporation into chromatin (Jodar et al. 2011, He et al. 2012). A meta-analysis concluded that the two most commonly found SNPs (PRM1 rs2301365 (190C>A) and PRM2 rs1646022 (298 G>C)) were likely to play a role in human fertility problems, but that more research is needed to support the clear association with male infertility (Jiang et al. 2018). The SNP rs2301365 locates to the 5′-UTR of PRM1, potentially affecting the expression of PRM1 (Jodar et al. 2011, He et al. 2012, Jiang et al. 2015, 2018). However, the strongest association with sub- and infertility in men has been reported for deviations of the protamine ratio (Balhorn et al. 1988, Belokopytova et al. 1993, Yebra & Oliva 1993, Khara et al. 1997, Bench et al. 1998a, Yebra 1998, Carrell & Liu 2001, Mengual et al. 2003, Oliva 2006, Torregrosa et al. 2006, Rogenhofer et al. 2013, Francis et al. 2014, Ni et al. 2016). Hence, the protamine ratio has been constituted as a biomarker for male fertility (Steger et al. 2008, Ni et al. 2016). A meta-analysis confirmed the association between aberrant sperm protamine ratio and subfertility (Ni et al. 2016). In addition, it seemed that PRM2 deregulation rather than PRM1 deregulation was the main cause of alterations in the protamine ratio, while a lower protamine ratio was associated with stronger DNA damage than higher ratios (Yebra & Oliva 1993, Carrell & Liu 2001, Aoki et al. 2005). However, alteration in the protamine ratio of mature sperm is not necessarily related to SNPs in protamine sequences or regulatory regions, but can also be caused by a mutation in interactors and potentially altered zinc levels. These factors should be considered in patients with altered protamine expression levels. Additionally, conventional ART techniques like IVF or ICSI with epididymal sperm show lower success rates for patients with impaired protamination (Nasr-Esfahani et al. 2004, Aoki et al. 2006b, Rogenhofer et al. 2013). ICSI of oocytes using testicular sperm from Prm2−/− mice induced the development of blastocyst stage embryos (Schneider et al. 2020), suggesting that this technique is a promising alternative treatment approach for men with protamine-related alterations.

Perspectives

Protamines are unique proteins, explicitly dedicated to packaging and protecting the paternal chromatin on its way to the fertilization site. Extensive investigation over the course of decades not only confirmed their essentiality many times but also strikingly highlighted that the role and function of protamines go beyond the simple packaging and shielding of DNA. The mere existence of a second protamine (Protamine 2) that has to be expressed in a certain ratio to the ubiquitous protamine 1 points toward high complexity in interaction dynamics and chromatin regulation. We suggest that by continuing to investigate the differences in protamine 1 and protamine 2 function, the role of the protamine ratio and protamine 2 processing, we will be able to unravel the whole range of protamine function. Their tightly restricted expression pattern and function together with recent technological advances, such as CRISPR/Cas, allow for relatively easy establishment and analysis of animal models. However, the tight compaction of sperm DNA presents restrictions in investigating interactions and protein complexes involved in chromatin compaction on a molecular level. Here in vitro studies will be indispensable. Taking what we learn from mouse models and applying it in in vitro settings, which will again inform further in vivo studies, will allow us to get a comprehensive picture of sperm chromatin dynamics. Further, comparative evolutionary studies allowed for conclusions about the essentiality of protamines, their domains and the protamine ratio as well as their function by carefully evaluating the naturally occurring phenotypic diversity and associating it to changes and selective pressures in gene sequences and expression rates. Thus, evolutionary-comparative approaches can significantly aid in our efforts to fully comprehend protamines. We would like to emphasize that only a transdisciplinary scheme, including biomedical, medical, cell molecular, biochemical, biophysics and evolutionary approaches, will lead us to fully understand how paternal chromatin is reorganized and how the fundamental genetic and epigenetic information is maintained for the next generation.

Lessons learned

  1. PRM1 and PRM2 DNA binding dynamics differ.While in vitro PRM1 and PRM2 had similar DNA binding affinities, processed PRM2 condensed DNA at a faster pace. Unprocessed PRM2 had a lower binding affinity to DNA and condensed DNA at a lower rate. In mice, Prm1 translation precedes Prm2 translation, suggesting distinct roles of both protamines in DNA binding rather than competition for DNA binding sites. PRM2 is processed upon binding to DNA in vivo, potentially slowing down DNA condensation and/or avoiding premature protamine–protamine interactions.

  2. PRM1 and PRM2 are both essential for fertility in mice. Shifts in the protamine ratio and changes in the protamine coding sequence have been related to male sub- and infertility both in mice and men. In mice, heterozygous deletion of Prm1 caused sperm defects and subfertility. One allele of Prm2, on the other hand, was sufficient to retain normal fertility. Loss of either protamine or the cP2 domain, however, lead to severe sperm damages and infertility.

  3. PRM1 and PRM2 have common and distinct functions.Both protamines participate in sperm DNA condensation. Unprocessed PRM2, however, carrying essential information in its N-terminus, seems to be involved in TNP-to-protamine turnover and H2A.L.2 retention. The process of PRM2 cleaving on the other hand might be essential to slow down DNA condensation. PRM1 is translated prior to PRM2 and seems to not only bind and condense DNA but also execute an additional function essential for sperm motility.

  4. PRM1 and PRM2 perform their functions in a complex interacting network.Prior to protamine loading, the stage is prepared by replacing canonical histones with specialized variants and global histone acetylation. This allows for binding of TNPs and the formation of specialized nucleosome complexes (H2A.L.2-TH2B-TNP). Protamines then come in to replace these complexes and possibly to regulate their eviction and the retention of a subset of nucleosomes (see Fig. 2). Studies have shown that when any of these key components are missing, chromatin condensation is negatively affected.

Declaration of interest

The authors declare that there is no conflict of interest that could be perceived as prejudicing the impartiality of the research reported.

Funding

Our work was supported by DFG grants to L A (AR 1221/1-1), H S (Scho 503/15-2; Scho 503/23-1) and S S (Schn 1668/1-1).

Author contribution statement

L A, G M, S S and H S wrote, discussed and edited the manuscript.

Acknowledgements

The authors are very grateful to their colleagues whose excellent research was included in this review, as well as to their colleagues for writing the outstanding reviews that were published before this one on similar topics. We apologize to colleagues whose publications are not referenced here, due to space limits. We thank the anonymous reviewers for their constructive criticism and excellent suggestions, which considerably improved this manuscript.

References

  • Aitken RJ, Jones KT & Robertson SA 2012 Reactive oxygen species and sperm function – in sickness and in health. Journal of Andrology 33 10961106. (https://doi.org/10.2164/jandrol.112.016535)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Akama K, Sato H, Hasegawa S, Shimada I & Nakano M 1998 Transition protein 1 from boar late spermatid nuclei having DNA-melting activity is a dimeric protein. Biochemistry and Molecular Biology International 44 315323. (https://doi.org/10.1080/15216549800201332)

    • PubMed
    • Search Google Scholar
    • Export Citation
  • Aoki VW, Moskovtsev SI, Willis J, Liu L, Mullen JB & Carrell DT 2005 DNA integrity is compromised in protamine-deficient human sperm. Journal of Andrology 26 741748. (https://doi.org/10.2164/jandrol.05063)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Aoki VW, Liu L & Carrell DT 2006a A novel mechanism of protamine expression deregulation highlighted by abnormal protamine transcript retention in infertile human males with sperm protamine deficiency. Molecular Human Reproduction 12 4150. (https://doi.org/10.1093/molehr/gah258)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Aoki VW, Liu L, Jones KP, Hatasaka HH, Gibson M, Peterson CM & Carrell DT 2006b Sperm protamine 1/protamine 2 ratios are related to in vitro fertilization pregnancy rates and predictive of fertilization ability. Fertility and Sterility 86 14081415. (https://doi.org/10.1016/j.fertnstert.2006.04.024)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Arévalo L, Tourmente M, Varea-Sánchez M, Ortiz-García D & Roldan ERS 2021 Sexual selection towards a protamine expression ratio optimum in two rodent groups? Evolution; International Journal of Organic Evolution 75 21242131. (https://doi.org/10.1111/evo.14305)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Arévalo L, Merges GE, Schneider S, Oben FE, Neumann IS & Schorle H 2022 Loss of the cleaved-protamine 2 domain leads to incomplete histone-to-protamine exchange and infertility in mice. PLoS Genetics 18 e1010272. (https://doi.org/10.1371/journal.pgen.1010272)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Arpanahi A, Brinkworth M, Iles D, Krawetz SA, Paradowska A, Platts AE, Saida M, Steger K, Tedder P & Miller D 2009 Endonuclease-sensitive regions of human spermatozoal chromatin are highly enriched in promoter and CTCF binding sequences. Genome Research 19 13381349. (https://doi.org/10.1101/gr.094953.109)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Ausio J, Gonzalez-Romero R & Woodcock CL 2014 Comparative structure of vertebrate sperm chromatin. Journal of Structural Biology 188 142155. (https://doi.org/10.1016/j.jsb.2014.09.004)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Balhorn R 1982 A model for the structure of chromatin in mammalian sperm. Journal of Cell Biology 93 298305. (https://doi.org/10.1083/jcb.93.2.298)

  • Balhorn R 2007 The protamine family of sperm nuclear proteins. Genome Biology 8 227. (https://doi.org/10.1186/gb-2007-8-9-227)

  • Balhorn R, Gledhill BL & Wyrobek AJ 1977 Mouse sperm chromatin proteins: quantitative isolation and partial characterization. Biochemistry 16 40744080. (https://doi.org/10.1021/bi00637a021)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Balhorn R, Reed S & Tanphaichitr N 1988 Aberrant protamine 1/protamine 2 ratios in sperm of infertile human males. Experientia 44 5255. (https://doi.org/10.1007/BF01960243)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Balhorn R, Corzett M & Mazrimas JA 1992 Formation of intraprotamine disulfides in vitro. Archives of Biochemistry and Biophysics 296 384393. (https://doi.org/10.1016/0003-9861(9290588-n)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Balhorn R, Hud NV, Corzett M & Mazrimas J 1995 Importance of Protamine Phosphorylation to Histone Displacement in Spermatids: Can the Disruption of This Process Be Used for Male Contraception. Livermore, CA (United States): Lawrence Livermore National Lab(LLNL).

    • Search Google Scholar
    • Export Citation
  • Balhorn R, Brewer L & Corzett M 2000 DNA condensation by protamine and arginine-rich peptides: analysis of toroid stability using single DNA molecules. Molecular Reproduction and Development 56 (Supplement) 230234. (https://doi.org/10.1002/(SICI)1098-2795(200006)56:2+<230::AID-MRD3>3.0.CO;2-V)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Balhorn R, Steger K, Bergmann M, Schuppe HC, Neuhauser S & Balhorn MC 2018 New monoclonal antibodies specific for mammalian protamines P1 and P2. Systems Biology in Reproductive Medicine 64 424447. (https://doi.org/10.1080/19396368.2018.1510063)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Barral S, Morozumi Y, Tanaka H, Montellier E, Govin J, Dieuleveult de M, Charbonnier G, Couté Y, Puthier D & Buchou T et al.2017 Histone variant HA.L.2 guides transition protein-dependent protamine assembly in male germ cells. Molecular Cell 66 89 .e8101.e8. (https://doi.org/10.1016/j.molcel.2017.02.025)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Baskaran R & Rao MR 1990 Interaction of spermatid-specific protein TP2 with nucleic acids, in vitro. A comparative study with TP1. Journal of Biological Chemistry 265 2103921047. (https://doi.org/10.1016/S0021-9258(1745324-5)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Belokopytova IA, Kostyleva EI, Tomilin AN & Vorob’ev VI 1993 Human male infertility may be due to a decrease of the protamine P2 content in sperm chromatin. Molecular Reproduction and Development 34 5357. (https://doi.org/10.1002/mrd.1080340109)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Bench G, Corzett MH, Yebra de L, Oliva R & Balhorn R 1998 Protein and DNA contents in sperm from an infertile human male possessing protamine defects that vary over time. Molecular Reproduction and Development 50 345353. (https://doi.org/10.1002/(SICI)1098-2795(199807)50:3<345::AID-MRD11>3.0.CO;2-3)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Bench G, Corzett MH, Kramer CE, Grant PG & Balhorn R 2000 Zinc is sufficiently abundant within mammalian sperm nuclei to bind stoichiometrically with protamine 2. Molecular Reproduction and Development 56 512519. (https://doi.org/10.1002/1098-2795(200008)56:4<512::AID-MRD9>3.0.CO;2-M)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Bjorndahl L & Kvist U 2014 Structure of chromatin in spermatozoa. Advances in Experimental Medicine and Biology 791 111. (https://doi.org/10.1007/978-1-4614-7783-9_1)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Blévec EL, Muroňová J, Ray PF & Arnoult C 2020 Paternal epigenetics: mammalian sperm provide much more than DNA at fertilization. Molecular and Cellular Endocrinology 518 110964. (https://doi.org/10.1016/j.mce.2020.110964)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Branson RE, Grimes SR, Yonuschot G & Irvin JL 1975 The histones of rat testis. Archives of Biochemistry and Biophysics 168 403412. (https://doi.org/10.1016/0003-9861(7590269-6)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Braun RE 2001 Packaging paternal chromosomes with protamine. Nature Genetics 28 1012. (https://doi.org/10.1038/ng0501-10)

  • Brewer LR, Corzett M & Balhorn R 1999 Protamine-induced condensation and decondensation of the same DNA molecule. Science 286 120123. (https://doi.org/10.1126/science.286.5437.120)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Brewer L, Corzett M & Balhorn R 2002 Condensation of DNA by spermatid basic nuclear proteins. Journal of Biological Chemistry 277 3889538900. (https://doi.org/10.1074/jbc.M204755200)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Brunner AM, Nanni P & Mansuy IM 2014 Epigenetic marking of sperm by post-translational modification of histones and protamines. Epigenetics and Chromatin 7 2. (https://doi.org/10.1186/1756-8935-7-2)

    • PubMed
    • Search Google Scholar
    • Export Citation
  • Brykczynska U, Hisano M, Erkek S, Ramos L, Oakeley EJ, Roloff TC, Beisel C, Schübeler D, Stadler MB & Peters AHFM 2010 Repressive and active histone methylation mark distinct promoters in human and mouse spermatozoa. Nature Structural and Molecular Biology 17 679687. (https://doi.org/10.1038/nsmb.1821)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Calvin HI & Bedford JM 1971 Formation of disulphide bonds in the nucleus and accessory structures of mammalian spermatozoa during maturation in the epididymis. Journal of Reproduction and Fertility: Supplement 13 (Supplement 13) 6575.

    • Search Google Scholar
    • Export Citation
  • Candido EPM & Dixon GH 1972 Trout testis cells. 3. Acetylation of histones in different cell types from developing trout testis. Journal of Biological Chemistry 247 55065510. (https://doi.org/10.1016/S0021-9258(2081134-X)

    • PubMed
    • Search Google Scholar
    • Export Citation
  • Caron N, Veilleux S & Boissonneault G 2001 Stimulation of DNA repair by the spermatidal TP1 protein. Molecular Reproduction and Development 58 437443. (https://doi.org/10.1002/1098-2795(20010401)58:4<437::AID-MRD12>3.0.CO;2-Q)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Carone BR, Hung JH, Hainer SJ, Chou MT, Carone DM, Weng Z, Fazzio TG & Rando OJ 2014 High-resolution mapping of chromatin packaging in mouse embryonic stem cells and sperm. Developmental Cell 30 1122. (https://doi.org/10.1016/j.devcel.2014.05.024)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Carré-Eusèbe D, Lederer F, KHD & Elsevier SM 1991 Processing of the precursor of protamine P2 in mouse. Peptide mapping and N -terminal sequence analysis of intermediates. Biochemical Journal 277 3945. (https://doi.org/10.1042/bj2770039)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Carrell DT & Liu L 2001 Altered protamine 2 expression is uncommon in donors of known fertility, but common among men with poor fertilizing capacity, and may reflect other abnormalities of spermiogenesis. Journal of Andrology 22 604610.

    • Search Google Scholar
    • Export Citation
  • Carrell DT, Emery BR & Hammoud S 2007 Altered protamine expression and diminished spermatogenesis: what is the link? Human Reproduction Update 13 313327. (https://doi.org/10.1093/humupd/dml057)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Castillo J, Estanyol JM, Ballesca JL & Oliva R 2015 Human sperm chromatin epigenetic potential: genomics, proteomics, and male infertility. Asian Journal of Andrology 17 601609. (https://doi.org/10.4103/1008-682X.153302)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Chandley AC & Kofman-Alfaro S 1971 ‘Unscheduled’ DNA synthesis in human germ cells following UV irradiation. Experimental Cell Research 69 4548. (https://doi.org/10.1016/0014-4827(7190308-9)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Chauvière M, Martinage A, Debarle M, Sautière P & Chevaillier P 1992 Molecular characterization of six intermediate proteins in the processing of mouse protamine P2 precursor. European Journal of Biochemistry 204 759765. (https://doi.org/10.1111/j.1432-1033.1992.tb16691.x)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Cheng AC, Chen WW, Fuhrmann CN & Frankel AD 2003 Recognition of nucleic acid bases and base-pairs by hydrogen bonding to amino acid side-chains. Journal of Molecular Biology 327 781796. (https://doi.org/10.1016/s0022-2836(0300091-3)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Chira F, Arkhis A, Martinage A, Jaquinod M, Chevaillier P & Sautière P 1993 Phosphorylation of human sperm protamines HP1 and HP2: identification of phosphorylation sites. Biochimica et Biophysica Acta 1203 109114. (https://doi.org/10.1016/0167-4838(9390043-q)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Cho C, Willis WD, Goulding EH, Jung-Ha H, Choi YC, Hecht NB & Eddy EM 2001 Haploinsufficiency of protamine-1 or -2 causes infertility in mice. Nature Genetics 28 8286. (https://doi.org/10.1038/ng0501-82)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Cho C, Jung-Ha H, Willis WD, Goulding EH, Stein P, Xu Z, Schultz RM, Hecht NB & Eddy EM 2003 Protamine 2 deficiency leads to sperm DNA damage and embryo death in mice. Biology of Reproduction 69 211217. (https://doi.org/10.1095/biolreprod.102.015115)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Conrad M, Moreno SG, Sinowatz F, Ursini F, Kölle S, Roveri A, Brielmeier M, Wurst W, Maiorino M & Bornkamm GW 2005 The nuclear form of phospholipid hydroperoxide glutathione peroxidase is a protein thiol peroxidase contributing to sperm chromatin stability. Molecular and Cellular Biology 25 76377644. (https://doi.org/10.1128/MCB.25.17.7637-7644.2005)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Corzett M, Mazrimas J & Balhorn R 2002 Protamine 1: protamine 2 stoichiometry in the sperm of eutherian mammals. Molecular Reproduction and Development 61 519527. (https://doi.org/10.1002/mrd.10105)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Cree LH, Balhorn R & Brewer LR 2011 Single molecule studies of DNA-protamine interactions. Protein and Peptide Letters 18 802810. (https://doi.org/10.2174/092986611795713943)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Czernik M, Iuso D, Toschi P, Khochbin S & Loi P 2016 Remodeling somatic nuclei via exogenous expression of protamine 1 to create spermatid-like structures for somatic nuclear transfer. Nature Protocols 11 21702188. (https://doi.org/10.1038/nprot.2016.130)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Depa-Martynow M, Kempisty B, Jagodziński PP, Pawelczyk L & Jedrzejczak P 2012 Impact of protamine transcripts and their proteins on the quality and fertilization ability of sperm and the development of preimplantation embryos. Reproductive Biology 12 5772. (https://doi.org/10.1016/s1642-431x(1260077-1)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Dhar S, Thota A & Rao MRS 2012 Insights into role of bromodomain, testis-specific (Brdt) in acetylated histone H4-dependent chromatin remodeling in mammalian spermiogenesis. Journal of Biological Chemistry 287 63876405. (https://doi.org/10.1074/jbc.M111.288167)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Erkek S, Hisano M, Liang CY, Gill M, Murr R, Dieker J, Schübeler D, Vlag van der J, Stadler MB & Peters AHFM 2013 Molecular determinants of nucleosome retention at CpG-rich sequences in mouse spermatozoa. Nature Structural and Molecular Biology 20 868875. (https://doi.org/10.1038/nsmb.2599)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Francis S, Yelumalai S, Jones C & Coward K 2014 Aberrant protamine content in sperm and consequential implications for infertility treatment. Human Fertility 17 8089. (https://doi.org/10.3109/14647273.2014.915347)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Gardiner-Garden M, Ballesteros M, Gordon M & Tam PPL 1998 Histone- and protamine-DNA association: conservation of different patterns within the β-globin domain in human sperm. Molecular and Cellular Biology 18 33503356. (https://doi.org/10.1128/MCB.18.6.3350)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Gaucher J, Boussouar F, Montellier E, Curtet S, Buchou T, Bertrand S, Hery P, Jounier S, Depaux A & Vitte AL et al.2012 Bromodomain-dependent stage-specific male genome programming by Brdt. EMBO Journal 31 38093820. (https://doi.org/10.1038/emboj.2012.233)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Gázquez C, Oriola J, Mateo S, Vidal‐Taboada JM, Ballescà JL & Oliva R 2008 A common protamine 1 promoter polymorphism (−190 C→A) correlates With abnormal sperm morphology and increased protamine P1/P2 ratio in infertile patients. Journal of Andrology 29 540548. (https://doi.org/10.2164/jandrol.107.004390)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Gosálvez J, López-Fernández C, Fernández JL, Gouraud A & Holt WV 2011 Relationships between the dynamics of iatrogenic DNA damage and genomic design in mammalian spermatozoa from eleven species. Molecular Reproduction and Development 78 951961. (https://doi.org/10.1002/mrd.21394)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Gou LT, Lim DH, Ma W, Aubol BE, Hao Y, Wang X, Zhao J, Liang Z, Shao C & Zhang X et al.2020 Initiation of parental genome reprogramming in fertilized oocyte by splicing kinase SRPK1-catalyzed protamine phosphorylation. Cell 180 1212 .e141227.e14. (https://doi.org/10.1016/j.cell.2020.02.020)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Goudarzi A, Shiota H, Rousseaux S & Khochbin S 2014 Genome-scale acetylation-dependent histone eviction during spermatogenesis. Journal of Molecular Biology 426 33423349. (https://doi.org/10.1016/j.jmb.2014.02.023)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Govin J, Caron C, Lestrat C, Rousseaux S & Khochbin S 2004 The role of histones in chromatin remodelling during mammalian spermiogenesis. European Journal of Biochemistry 271 34593469. (https://doi.org/10.1111/j.1432-1033.2004.04266.x)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Govin J, Escoffier E, Rousseaux S, Kuhn L, Ferro M, Thévenon J, Catena R, Davidson I, Garin J & Khochbin S et al.2007 Pericentric heterochromatin reprogramming by new histone variants during mouse spermiogenesis. Journal of Cell Biology 176 283294. (https://doi.org/10.1083/jcb.200604141)

    • Crossref
    • PubMed
    • Search Google Scholar
    • Export Citation
  • Green GR 2001 Phosphorylation of histone variant regions in chromatin: unlocking the linker? Biochemistry and Cell Biology 79 275287.

  • Gusse M, Sautière P, Bélaiche D, Martinage A, Roux C, Dadoune JP & Chevaillier P 1986 Purification and characterization of nuclear basic proteins of human sperm. Biochimica et Biophysica Acta 884 124134. (https://doi.org/10.1016/0304-4165(8690235-7)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Hammoud SS, Nix DA, Zhang H, Purwar J, Carrell DT & Cairns BR 2009a Distinctive chromatin in human sperm packages genes for embryo development. Nature 460 473478. (https://doi.org/10.1038/nature08162)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Hammoud S, Liu L & Carrell DT 2009b Protamine ratio and the level of histone retention in sperm selected from a density gradient preparation. Andrologia 41 8894. (https://doi.org/10.1111/j.1439-0272.2008.00890.x)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Hammoud SS, Purwar J, Pflueger C, Cairns BR & Carrell DT 2010 Alterations in sperm DNA methylation patterns at imprinted loci in two classes of infertility. Fertility and Sterility 94 17281733. (https://doi.org/10.1016/j.fertnstert.2009.09.010)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Hammoud SS, Nix DA, Hammoud AO, Gibson M, Cairns BR & Carrell DT 2011 Genome-wide analysis identifies changes in histone retention and epigenetic modifications at developmental and imprinted gene loci in the sperm of infertile men. Human Reproduction 26 25582569. (https://doi.org/10.1093/humrep/der192)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Hazzouri M, Rousseaux S, Mongelard F, Usson Y, Pelletier R, Faure AK, Vourc’h C & Sle B 2000 Genome organization in the human sperm nucleus studied by FISH and confocal microscopy. Molecular Reproduction and Development 55 307315. (https://doi.org/10.1002/(SICI)1098-2795(200003)55:3<307::AID-MRD9>3.0.CO;2-P)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • He XJ, Ruan J, Du WD, Chen G, Zhou Y, Xu S, Zuo XB, Cao YX & Zhang XJ 2012 PRM1 variant rs35576928 (ArgSer) is associated with defective spermatogenesis in the Chinese Han population. Reproductive Biomedicine Online 25 627634. (https://doi.org/10.1016/j.rbmo.2012.09.005)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Hecht NB, Bower PA, Waters SH, Yelick PC & Distel RJ 1986a Evidence for haploid expression of mouse testicular genes. Experimental Cell Research 164 183190. (https://doi.org/10.1016/0014-4827(8690465-9)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Hecht NB, Kleene KC, Yelick PC, Johnson PA, Pravtcheva DD & Ruddle FH 1986b Mapping of haploid expressed genes: genes for both mouse protamines are located on chromosome 16. Somatic Cell and Molecular Genetics 12 203208. (https://doi.org/10.1007/BF01560667)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Heidaran MA & Kistler WS 1987 Transcriptional and translational control of the message for transition protein 1, a major chromosomal protein of mammalian spermatids. Journal of Biological Chemistry 262 1330913315. (https://doi.org/10.1016/S0021-9258(1845202-7)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Hoghoughi N, Barral S, Curtet S, Chuffart F, Charbonnier G, Puthier D, Buchou T, Rousseaux S & Khochbin S 2020 RNA-guided genomic localization of H2A.L.2 histone variant. Cells 9 474. (https://doi.org/10.3390/cells9020474)

    • Search Google Scholar
    • Export Citation
  • Hutchison JM, Rau DC & DeRouchey JE 2017 Role of disulfide bonds on DNA packaging forces in bull sperm chromatin. Biophysical Journal 113 19251933. (https://doi.org/10.1016/j.bpj.2017.08.050)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Imken L, Rouba H, Houate BE, Louanjli N, Barakat A, Chafik A & McElreavey K 2009 Mutations in the protamine locus: association with spermatogenic failure? Molecular Human Reproduction 15 733738. (https://doi.org/10.1093/molehr/gap056)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Iranpour FG 2014 The effects of protamine deficiency on ultrastructure of human sperm nucleus. Advanced Biomedical Research 3 24. (https://doi.org/10.4103/2277-9175.124666)

    • Search Google Scholar
    • Export Citation
  • Itoh K, Kondoh G, Miyachi H, Sugai M, Kaneko Y, Kitano S, Watanabe H, Maeda R, Imura A & Liu Y et al.2019 Dephosphorylation of protamine 2 at serine 56 is crucial for murine sperm maturation in vivo. Science Signaling 12 eaao7232. (https://doi.org/10.1126/scisignal.aao7232)

    • Search Google Scholar
    • Export Citation
  • Iuso D, Czernik M, Toschi P, Fidanza A, Zacchini F, Feil R, Curtet S, Buchou T, Shiota H & Khochbin S et al.2015 Exogenous expression of human protamine 1 (hPrm1) remodels fibroblast nuclei into spermatid-like structures. Cell Reports 13 17651771. (https://doi.org/10.1016/j.celrep.2015.10.066)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Jenuwein T & Allis CD 2001 Translating the histone code. Science 293 10741080. (https://doi.org/10.1126/science.1063127)

  • Jiang W, Sun H, Zhang J, Zhou Q, Wu Q, Li T, Zhang C, Li W, Zhang M & Xia X 2015 Polymorphisms in protamine 1 and protamine 2 predict the risk of male infertility: a meta-analysis. Scientific Reports 5 15300. (https://doi.org/10.1038/srep15300)

    • Search Google Scholar
    • Export Citation
  • Jiang W, Shi L, Liu H, Cao J, Zhu P, Zhang J, Yu M, Guo Y, Cui Y & Xia X 2018 Systematic review and meta-analysis of the genetic association between protamine polymorphism and male infertility. Andrologia 50 e12990. (https://doi.org/10.1111/and.12990)

    • Search Google Scholar
    • Export Citation
  • Jodar M, Oriola J, Mestre G, Castillo J, Giwercman A, Vidal‐Taboada JM, Ballescà JL & Oliva R 2011 Polymorphisms, haplotypes and mutations in the protamine 1 and 2 genes. International Journal of Andrology 34 470485. (https://doi.org/10.1111/j.1365-2605.2010.01115.x)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Jung YH, Sauria MEG, Lyu X, Cheema MS, Ausio J, Taylor J & Corces VG 2017 Chromatin states in mouse sperm correlate with embryonic and adult regulatory landscapes. Cell Reports 18 13661382.

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Kasinsky HE, Eirin-Lopez JM & Ausio J 2011 Protamines: structural complexity, evolution and chromatin patterning. Protein and Peptide Letters 18 755771. (https://doi.org/10.2174/092986611795713989)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Khara KK, Vlad M, Griffiths M & Kennedy CR 1997 Human protamines and male infertility. Journal of Assisted Reproduction and Genetics 14 282290. (https://doi.org/10.1007/BF02765830)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Kierszenbaum AL & Tres LL 1975 Structural and transcriptional features of the mouse spermatid genome. Journal of Cell Biology 65 258270. (https://doi.org/10.1083/jcb.65.2.258)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Kim YJ, Hwang I, Tres LL, Kierszenbaum AL & Chae CB 1987 Molecular cloning and differential expression of somatic and testis-specific H2B histone genes during rat spermatogenesis. Developmental Biology 124 2334. (https://doi.org/10.1016/0012-1606(8790455-6)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Kistler WS, Noyes C, Hsu R & Heinrikson RL 1975 The amino acid sequence of a testis-specific basic protein that is associated with spermatogenesis. Journal of Biological Chemistry 250 18471853. (https://doi.org/10.1016/S0021-9258(1941772-9)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Klaus ES, Gonzalez NH, Bergmann M, Bartkuhn M, Weidner W, Kliesch S & Rathke C 2016 Murine and human spermatids are characterized by numerous, newly synthesized and differentially expressed transcription factors and bromodomain-containing proteins. Biology of Reproduction 95 4. (https://doi.org/10.1095/biolreprod.115.137620)

    • Search Google Scholar
    • Export Citation
  • Kleene KC 1989 Poly(A) shortening accompanies the activation of translation of five mRNAs during spermiogenesis in the mouse. Development 106 367373. (https://doi.org/10.1242/dev.106.2.367)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Kleene KC & Flynn JF 1987 Characterization of a cDNA clone encoding a basic protein, TP2, involved in chromatin condensation during spermiogenesis in the mouse. Journal of Biological Chemistry 262 1727217277. (https://doi.org/10.1016/S0021-9258(1845373-2)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Kleene KC, Distel RJ & Hecht NB 1984 Translational regulation and deadenylation of a protamine mRNA during spermiogenesis in the mouse. Developmental Biology 105 7179. (https://doi.org/10.1016/0012-1606(8490262-8)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Laberge RM & Boissonneault G 2005 Chromatin remodeling in spermatids: a sensitive step for the genetic integrity of the male gamete. Archives of Andrology 51 125133. (https://doi.org/10.1080/014850190518134)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Lee K, Haugen HS, Clegg CH & Braun RE 1995 Premature translation of protamine 1 mRNA causes precocious nuclear condensation and arrests spermatid differentiation in mice. PNAS 92 1245112455. (https://doi.org/10.1073/pnas.92.26.12451)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Lesch BJ & Page DC 2014 Poised chromatin in the mammalian germ line. Development 141 36193626. (https://doi.org/10.1242/dev.113027)

  • Lévesque D, Veilleux S, Caron N & Boissonneault G 1998 Architectural DNA-binding properties of the spermatidal transition proteins 1 and 2. Biochemical and Biophysical Research Communications 252 602609. (https://doi.org/10.1006/bbrc.1998.9687)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Lewis JD, Saperas N, Song Y, Zamora MJ, Chiva M & Ausió J 2004 Histone H1 and the origin of protamines. PNAS 101 41484152. (https://doi.org/10.1073/pnas.0308721101)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Louie AJ & Dixon GH 1972 Kinetics of enzymatic modification of the protamines and a proposal for their binding to chromatin. Journal of Biological Chemistry 247 79627968. (https://doi.org/10.1016/S0021-9258(2081796-7)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Lüke L, Vicens A, Serra F, Luque-Larena JJ, Dopazo H, Roldan ERS & Gomendio M 2011 Sexual selection halts the relaxation of protamine 2 among rodents. PLoS ONE 6 e29247. (https://doi.org/10.1371/journal.pone.0029247)

    • Search Google Scholar
    • Export Citation
  • Lüke L, Vicens A, Tourmente M & Roldan ER 2014a Evolution of protamine genes and changes in sperm head phenotype in rodents. Biology of Reproduction 90 67. (https://doi.org/10.1095/biolreprod.113.115956)

    • Search Google Scholar
    • Export Citation
  • Lüke L, Campbell P, Sánchez MV, Nachman MW & Roldan ERS 2014b Sexual selection on protamine and transition nuclear protein expression in mouse species. Proceedings of the Royal Society of London: Series B 281 359. (https://doi.org/10.1098/rspb.2013.3359)

    • Search Google Scholar
    • Export Citation
  • Lüke L, Tourmente M, Dopazo H, Serra F & Roldan ERS 2016a Selective constraints on protamine 2 in primates and rodents. BMC Evolutionary Biology 16 21. (https://doi.org/10.1186/s12862-016-0588-1)

    • Search Google Scholar
    • Export Citation
  • Lüke L, Tourmente M & Roldan ERS 2016b Sexual selection of protamine 1 in mammals. Molecular Biology and Evolution 33 174184. (https://doi.org/10.1093/molbev/msv209)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Marcon L & Boissonneault G 2004 Transient DNA strand breaks during mouse and human spermiogenesis: new insights in stage specificity and link to chromatin remodeling. Biology of Reproduction 70 910918. (https://doi.org/10.1095/biolreprod.103.022541)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Mashiko D, Fujihara Y, Satouh Y, Miyata H, Isotani A & Ikawa M 2013 Generation of mutant mice by pronuclear injection of circular plasmid expressing Cas9 and single guided RNA. Scientific Reports 3 3355. (https://doi.org/10.1038/srep03355)

    • Search Google Scholar
    • Export Citation
  • McPherson SM & Longo FJ 1993 Nicking of rat spermatid and spermatozoa DNA: possible involvement of DNA topoisomerase II. Developmental Biology 158 122130. (https://doi.org/10.1006/dbio.1993.1173)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Meistrich ML, Trostle-Weige PK, Lin R, Bhatnagar YM & Allis CD 1992 Highly acetylated H4 is associated with histone displacement in rat spermatids. Molecular Reproduction and Development 31 170181. (https://doi.org/10.1002/mrd.1080310303)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Meistrich ML, Mohapatra B, Shirley CR & Zhao M 2003 Roles of transition nuclear proteins in spermiogenesis. Chromosoma 111 483488. (https://doi.org/10.1007/s00412-002-0227-z)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Mengual L, Ballescá JL, Ascaso C & Oliva R 2003 Marked differences in protamine content and P1/P2 ratios in sperm cells from Percoll fractions between patients and controls. Journal of Andrology 24 438447. (https://doi.org/10.1002/j.1939-4640.2003.tb02692.x)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Merges GE, Meier J, Schneider S, Kruse A, Fröbius AC, Kirfel G, Steger K, Arévalo L & Schorle H 2022 Loss of Prm1 leads to defective chromatin protamination, impaired PRM2 processing, reduced sperm motility and subfertility in male mice. Development 149 dev200330. (https://doi.org/10.1242/dev.200330)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Miescher T 1874 Das protamin – eine neue organische Basis aus den Samenfäden des Rheinlachses. Zeitschrift für Analytische Chemie 13 325326. (https://doi.org/10.1007/BF01302341)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Montellier E, Boussouar F, Rousseaux S, Zhang K, Buchou T, Fenaille F, Shiota H, Debernardi A, Hery P & Curtet S et al.2013 Chromatin-to-nucleoprotamine transition is controlled by the histone H2B variant TH2B. Genes and Development 27 16801692. (https://doi.org/10.1101/gad.220095.113)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Moreno SG, Laux G, Brielmeier M, Bornkamm GW & Conrad M 2003 Testis-specific expression of the nuclear form of phospholipid hydroperoxide glutathione peroxidase (PHGPx). Biological Chemistry 384 635643. (https://doi.org/10.1515/BC.2003.070)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Moritz L, Schon SB, Rabbani M, Sheng Y, Pendlebury DF, Agrawal R, Sultan C, Jorgensen K, Zheng X & Diehl A et al.2021 Single residue substitution in protamine 1 disrupts sperm genome packaging and embryonic development in mice. bioRxiv. (https://doi.org/10.1101/2021.09.16.460631)

    • Search Google Scholar
    • Export Citation
  • Nasr-Esfahani MH, Salehi M, Razavi S, Mardani M, Bahramian H, Steger K & Oreizi F 2004 Effect of protamine-2 deficiency on ICSI outcome. Reproductive Biomedicine Online 9 652658. (https://doi.org/10.1016/s1472-6483(1061776-2)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Ni K, Spiess AN, Schuppe HC & Steger K 2016 The impact of sperm protamine deficiency and sperm DNA damage on human male fertility: a systematic review and meta-analysis. Andrology 4 789799. (https://doi.org/10.1111/andr.12216)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Oliva R 2006 Protamines and male infertility. Human Reproduction Update 12 417435. (https://doi.org/10.1093/humupd/dml009)

  • Oliva R & Dixon GH 1991 Vertebrate protamine genes and the histone-to-protamine replacement reaction. Progress in Nucleic Acid Research and Molecular Biology 40 2594. (https://doi.org/10.1016/s0079-6603(0860839-9)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Olson EN, Arnold HH, Rigby PW & Wold BJ 1996 Know your neighbors: three phenotypes in null mutants of the myogenic bHLH gene MRF4. Cell 85 14. (https://doi.org/10.1016/s0092-8674(0081073-9)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Papoutsopoulou S, Nikolakaki E, Chalepakis G, Kruft V, Chevaillier P & Giannakouros T 1999 SR protein-specific kinase 1 is highly expressed in testis and phosphorylates protamine 1. Nucleic Acids Research 27 29722980. (https://doi.org/10.1093/nar/27.14.2972)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Perreault SD, Barbee RR, Elstein KH, Zucker RM & Keefer CL 1988 Interspecies differences in the stability of mammalian sperm nuclei assessed in vivo by sperm microinjection and in vitro by flow cytometry. Biology of Reproduction 39 157167. (https://doi.org/10.1095/biolreprod39.1.157)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Pfeifer H, Conrad M, Roethlein D, Kyriakopoulos A, Brielmeier M, Bornkamm GW & Behne D 2001 Identification of a specific sperm nuclei selenoenzyme necessary for protamine thiol cross-linking during sperm maturation. FASEB Journal 15 12361238. (https://doi.org/10.1096/fj.00-0655fje)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Pirhonen A, Linnala-Kankkunen A & Maenpaa PH 1994a Identification of phosphoseryl residues in protamines from mature mammalian spermatozoa. Biology of Reproduction 50 981986. (https://doi.org/10.1095/biolreprod50.5.981)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Pirhonen A, Linnala-Kankkunen A & Mënpää PH 1994b P2 protamines are phosphorylated in vitro by protein kinase C, whereas P1 protamines prefer cAMP-dependent protein kinase. A comparative study of five mammalian species. European Journal of Biochemistry 223 165169. (https://doi.org/10.1111/j.1432-1033.1994.tb18979.x)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Pittoggi C, Renzi L, Zaccagnini G, Cimini D, Degrassi F, Giordano R, Magnano AR, Lorenzini R, Lavia P & Spadafora C 1999 A fraction of mouse sperm chromatin is organized in nucleosomal hypersensitive domains enriched in retroposon DNA. Journal of Cell Science 112 35373548. (https://doi.org/10.1242/jcs.112.20.3537)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Piunti A & Shilatifard A 2021 The roles of Polycomb repressive complexes in mammalian development and cancer. Nature Reviews: Molecular Cell Biology 22 326345. (https://doi.org/10.1038/s41580-021-00341-1)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Pivot-Pajot C, Caron C, Govin J, Vion A, Rousseaux S & Khochbin S 2003 Acetylation-dependent chromatin reorganization by BRDT, a testis-specific bromodomain-containing protein. Molecular and Cellular Biology 23 53545365. (https://doi.org/10.1128/MCB.23.15.5354-5365.2003)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Pogany GC, Corzett M, Weston S & Balhorn R 1981 DNA and protein content of mouse sperm Implications regarding sperm chromatin structure. Experimental Cell Research 136 127136. (https://doi.org/10.1016/0014-4827(8190044-6)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Pruslin FH, Imesch E, Winston R & Rodman TC 1987 Phosphorylation state of protamines 1 and 2 in human spermatids and spermatozoa. Gamete Research 18 179190. (https://doi.org/10.1002/mrd.1120180208)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Puglisi R, Maccari I, Pipolo S, Mangia F & Boitani C 2014 The nuclear form of glutathione peroxidase 4 colocalizes and directly interacts with protamines in the nuclear matrix during mouse sperm chromatin assembly. Spermatogenesis 4 e28460. (https://doi.org/10.4161/spmg.28460)

    • Search Google Scholar
    • Export Citation
  • Rathke C, Baarends WM, Awe S & Renkawitz-Pohl R 2014 Chromatin dynamics during spermiogenesis. Biochimica et Biophysica Acta 1839 155168. (https://doi.org/10.1016/j.bbagrm.2013.08.004)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Ravel C, Chantot-Bastaraud S, Houate BE, Berthaut I, Verstraete L, Larouziere de V, Lourenço D, Dumaine A, Antoine JM & Mandelbaum J et al.2007 Mutations in the protamine 1 gene associated with male infertility. Molecular Human Reproduction 13 461464. (https://doi.org/10.1093/molehr/gam031)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Rezaei-Gazik M, Vargas A, Amiri-Yekta A, Vitte AL, Akbari A, Barral S, Esmaeili V, Chuffart F, Sadighi-Gilani MA & Couté Y et al.2022 Direct visualization of pre-protamine 2 detects protamine assembly failures and predicts ICSI success. Molecular Human Reproduction 28 gaac004. (https://doi.org/10.1093/molehr/gaac004)

    • Search Google Scholar
    • Export Citation
  • Rogenhofer N, Dansranjavin T, Schorsch M, Spiess A, Wang H, Schönfeldt von V, Cappallo-Obermann H, Baukloh V, Yang H & Paradowska A et al.2013 The sperm protamine mRNA ratio as a clinical parameter to estimate the fertilizing potential of men taking part in an ART programme. Human Reproduction 28 969978. (https://doi.org/10.1093/humrep/des471)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Rooney AP, Zhang J & Nei M 2000 An unusual form of purifying selection in a sperm protein. Molecular Biology and Evolution 17 278283. (https://doi.org/10.1093/oxfordjournals.molbev.a026307)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Ruiz-López MJ, Espeso G, Evenson DP, Roldan ERS & Gomendio M 2010 Paternal levels of DNA damage in spermatozoa and maternal parity influence offspring mortality in an endangered ungulate. Proceedings: Biological Sciences 277 25412546. (https://doi.org/10.1098/rspb.2010.0333)

    • Search Google Scholar
    • Export Citation
  • Samans B, Yang Y, Krebs S, Sarode GV, Blum H, Reichenbach M, Wolf E, Steger K, Dansranjavin T & Schagdarsurengin U 2014 Uniformity of nucleosome preservation pattern in mammalian sperm and its connection to repetitive DNA elements. Developmental Cell 1 113.

    • Search Google Scholar
    • Export Citation
  • Schneider S, Balbach M, Jan FJ, Fietz D, Nettersheim D, Jostes S, Schmidt R, Kressin M, Bergmann M & Wachten D et al.2016 Re-visiting the protamine-2 locus: deletion, but not haploinsufficiency, renders male mice infertile. Scientific Reports 6 36764. (https://doi.org/10.1038/srep36764)

    • Search Google Scholar
    • Export Citation
  • Schneider S, Shakeri F, Trötschel C, Arévalo L, Kruse A, Buness A, Poetsch A, Steger K & Schorle H 2020 Protamine-2 deficiency initiates a reactive oxygen species (ROS)-mediated destruction cascade during epididymal sperm maturation in mice. Cells 9 1789. (https://doi.org/10.3390/cells9081789)

    • Search Google Scholar
    • Export Citation
  • Sega GA 1974 Unscheduled DNA synthesis in the germ cells of male mice exposed in vivo to the chemical mutagen ethyl methanesulfonate. PNAS 71 49554959. (https://doi.org/10.1073/pnas.71.12.4955)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Seligman J, Zipser Y & Kosower NS 2004 Tyrosine phosphorylation, thiol status, and protein tyrosine phosphatase in rat epididymal spermatozoa. Biology of Reproduction 71 10091015. (https://doi.org/10.1095/biolreprod.104.028035)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Shang E, Nickerson HD, Wen D, Wang X & Wolgemuth DJ 2007 The first bromodomain of Brdt, a testis-specific member of the BET sub-family of double-bromodomain-containing proteins, is essential for male germ cell differentiation. Development 134 35073515. (https://doi.org/10.1242/dev.004481)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Shinagawa T, Huynh LM, Takagi T, Tsukamoto D, Tomaru C, Kwak HG, Dohmae N, Noguchi J & Ishii S 2015 Disruption of Th2a and Th2b genes causes defects in spermatogenesis. Development 142 12871292. (https://doi.org/10.1242/dev.121830)

    • Search Google Scholar
    • Export Citation
  • Shirakata Y, Hiradate Y, Inoue H, Sato E & Tanemura K 2014 Histone H4 modification during mouse spermatogenesis. Journal of Reproduction and Development 60 383387. (https://doi.org/10.1262/jrd.2014-018)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Shires A, Carpenter MP & Chalkley R 1975 New histones found in mature mammalian testes. PNAS 72 27142718. (https://doi.org/10.1073/pnas.72.7.2714)

  • Shirley CR, Hayashi S, Mounsey S, Yanagimachi R & Meistrich ML 2004 Abnormalities and reduced reproductive potential of sperm from Tnp1- and Tnp2-null double mutant mice. Biology of Reproduction 71 12201229. (https://doi.org/10.1095/biolreprod.104.029363)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Singh J & Rao MR 1988 Interaction of rat testis protein, TP, with nucleosome core particle. Biochemistry International 17 701710.

  • Soler-Ventura A, Gay M, Jodar M, Vilanova M, Castillo J, Arauz-Garofalo G, Villarreal L, Ballescà JL, Vilaseca M & Oliva R 2020 Characterization of human sperm protamine proteoforms through a combination of top-down and bottom-up mass spectrometry approaches. Journal of Proteome Research 19 221237. (https://doi.org/10.1021/acs.jproteome.9b00499)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Steger K 1999 Transcriptional and translational regulation of gene expression in haploid spermatids. Anatomy and Embryology 199 471487. (https://doi.org/10.1007/s004290050245)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Steger K & Balhorn R 2018 Sperm nuclear protamines: a checkpoint to control sperm chromatin quality. Anatomia, Histologia, Embryologia 47 273279. (https://doi.org/10.1111/ahe.12361)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Steger K, Failing K, Klonisch T, Behre HM, Manning M, Weidner W, Hertle L, Bergmann M & Kliesch S 2001 Round spermatids from infertile men exhibit decreased protamine-1 and -2 mRNA. Human Reproduction 16 709716. (https://doi.org/10.1093/humrep/16.4.709)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Steger K, Fink L, Failing K, Bohle RM, Kliesch S, Weidner W & Bergmann M 2003 Decreased protamine-1 transcript levels in testes from infertile men. Molecular Human Reproduction 9 331336. (https://doi.org/10.1093/molehr/gag041)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Steger K, Wilhelm J, Konrad L, Stalf T, Greb R, Diemer T, Kliesch S, Bergmann M & Weidner W 2008 Both protamine-1 to protamine-2 mRNA ratio and Bcl2 mRNA content in testicular spermatids and ejaculated spermatozoa discriminate between fertile and infertile men. Human Reproduction 23 1116. (https://doi.org/10.1093/humrep/dem363)

    • Search Google Scholar
    • Export Citation
  • Swanson WJ & Vacquier VD 2002 The rapid evolution of reproductive proteins. Nature Reviews: Genetics 3 137144. (https://doi.org/10.1038/nrg733)

  • Takeda N, Yoshinaga K, Furushima K, Takamune K, Li Z, Abe S, Aizawa S & Yamamura K 2016 Viable offspring obtained from Prm1-deficient sperm in mice. Scientific Reports 6 27409. (https://doi.org/10.1038/srep27409)

    • Search Google Scholar
    • Export Citation
  • Tanaka H, Iguchi N, Carvalho de CE, Tadokoro Y, Yomogida K & Nishimune Y 2003 Novel actin-like proteins T-actin 1 and T-actin 2 are differentially expressed in the cytoplasm and nucleus of mouse haploid germ cells. Biology of Reproduction 69 475482. (https://doi.org/10.1095/biolreprod.103.015867)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Torregrosa N, Domínguez-Fandos D, Camejo MI, Shirley CR, Meistrich ML, Ballescà JL & Oliva R 2006 Protamine 2 precursors, protamine 1/protamine 2 ratio, DNA integrity and other sperm parameters in infertile patients. Human Reproduction 21 20842089. (https://doi.org/10.1093/humrep/del114)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Tremellen K 2008 Oxidative stress and male infertility – a clinical perspective. Human Reproduction Update 14 243258. (https://doi.org/10.1093/humupd/dmn004)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Turner LM, Chuong EB & Hoekstra HE 2008 Comparative analysis of testis protein evolution in rodents. Genetics 179 20752089. (https://doi.org/10.1534/genetics.107.085902)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Tüttelmann F, Krenková P, Römer S, Nestorovic AR, Ljujic M, Stambergová A, Macek M, Macek M, Nieschlag E & Gromoll J et al.2010 A common haplotype of protamine 1 and 2 genes is associated with higher sperm counts. International Journal of Andrology 33 e240e248. (https://doi.org/10.1111/j.1365-2605.2009.01003.x)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Ward WS & Coffey DS 1991 DNA packaging and organization in mammalian spermatozoa: comparison with somatic cells. Biology of Reproduction 44 569574. (https://doi.org/10.1095/biolreprod44.4.569)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Wu JY, Ribar TJ, Cummings DE, Burton KA, McKnight GS & Means AR 2000 Spermiogenesis and exchange of basic nuclear proteins are impaired in male germ cells lacking Camk4. Nature Genetics 25 448452. (https://doi.org/10.1038/78153)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Wyckoff GJ, Wang W & Wu CI 2000 Rapid evolution of male reproductive genes in the descent of man. Nature 403 304309. (https://doi.org/10.1038/35002070)

  • Xu H, Medina-Sanchez M, Magdanz V, Schwarz L, Hebenstreit F & Schmidt OG 2018 Sperm-hybrid micromotor for targeted drug delivery. ACS Nano 12 327337. (https://doi.org/10.1021/acsnano.7b06398)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Yamaguchi K, Hada M, Fukuda Y, Inoue E, Makino Y, Katou Y, Shirahige K & Okada Y 2018 Re-evaluating the localization of sperm-retained histones revealed the modification-dependent accumulation in specific genome regions. Cell Reports 23 39203932. (https://doi.org/10.1016/j.celrep.2018.05.094)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Yan W, Ma L, Burns KH & Matzuk MM 2003 HILS1 is a spermatid-specific linker histone H1-like protein implicated in chromatin remodeling during mammalian spermiogenesis. PNAS 100 1054610551. (https://doi.org/10.1073/pnas.1837812100)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Yebra L & Oliva R 1993 Rapid analysis of mammalian sperm nuclear proteins. Analytical Biochemistry 209 201203. (https://doi.org/10.1006/abio.1993.1104)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Yebra de L, Ballescá JL, Vanrell JA, Corzett M, Balhorn R & Oliva R 1998 Detection of P2 precursors in the sperm cells of infertile patients who have reduced protamine P2 levels. Fertility and Sterility 69 755759. (https://doi.org/10.1016/s0015-0282(9800012-0)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Yelick PC, Balhorn R, Johnson PA, Corzett M, Mazrimas JA, Kleene KC & Hecht NB 1987 Mouse protamine 2 is synthesized as a precursor whereas mouse protamine 1 is not. Molecular and Cellular Biology 7 21732179. (https://doi.org/10.1128/mcb.7.6.2173-2179.1987)

    • Search Google Scholar
    • Export Citation
  • Yu YE, Zhang Y, Unni E, Shirley CR, Deng JM, Russell LD, Weil MM, Behringer RR & Meistrich ML 2000 Abnormal spermatogenesis and reduced fertility in transition nuclear protein 1-deficient mice. PNAS 97 46834688. (https://doi.org/10.1073/pnas.97.9.4683)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Yu QF, Yang XX, Li FX, Ye LW, Wu YS & Mao XM 2012 Association of PRM1-190C- > a polymorphism with teratozoospermia. Zhonghua Nan Ke Xue 18 3 14317.

  • Zhao M, Shirley CR, Yu YE, Mohapatra B, Zhang Y, Unni E, Deng JM, Arango NA, Terry NHA & Weil MM et al.2001 Targeted disruption of the transition protein 2 gene affects sperm chromatin structure and reduces fertility in mice. Molecular and Cellular Biology 21 72437255. (https://doi.org/10.1128/MCB.21.21.7243-7255.2001)

    • Crossref
    • Search Google Scholar
    • Export Citation
  • Zhao M, Shirley CR, Hayashi S, Marcon L, Mohapatra B, Suganuma R, Behringer RR, Boissonneault G, Yanagimachi R & Meistrich ML 2004 Transition nuclear proteins are required for normal chromatin condensation and functional sperm development. Genesis 38 200213. (https://doi.org/10.1002/gene.20019)

    • Crossref
    • Search Google Scholar