Advertisement
Research| Volume 12, ISSUE 11, P3221-3236, November 2013

Download started.

Ok

Linking Spermatid Ribonucleic Acid (RNA) Binding Protein and Retrogene Diversity to Reproductive Success*

  • Karen M. Chapman
    Affiliations
    Departments of Pharmacology, University of Texas Southwestern Medical Center in Dallas, Texas

    Cecil H & Ida Green Center for Reproductive Biology Sciences, University of Texas Southwestern Medical Center in Dallas, Texas
    Search for articles by this author
  • Heather M. Powell
    Affiliations
    Departments of Pharmacology, University of Texas Southwestern Medical Center in Dallas, Texas

    Cecil H & Ida Green Center for Reproductive Biology Sciences, University of Texas Southwestern Medical Center in Dallas, Texas
    Search for articles by this author
  • Jaideep Chaudhary
    Affiliations
    Departments of Pharmacology, University of Texas Southwestern Medical Center in Dallas, Texas

    Cecil H & Ida Green Center for Reproductive Biology Sciences, University of Texas Southwestern Medical Center in Dallas, Texas
    Search for articles by this author
  • John M. Shelton
    Affiliations
    Departments of Internal Medicine - Division of Cardiology, University of Texas Southwestern Medical Center in Dallas, Texas
    Search for articles by this author
  • James A. Richardson
    Affiliations
    Departments of Pathology, University of Texas Southwestern Medical Center in Dallas, Texas

    Departments of Molecular Biology, University of Texas Southwestern Medical Center in Dallas, Texas
    Search for articles by this author
  • Timothy E. Richardson
    Affiliations
    Departments of Pharmacology, University of Texas Southwestern Medical Center in Dallas, Texas

    Cecil H & Ida Green Center for Reproductive Biology Sciences, University of Texas Southwestern Medical Center in Dallas, Texas
    Search for articles by this author
  • F. Kent Hamra
    Correspondence
    To whom correspondence should be addressed: Department of Pharmacology and Cecil H and Ida Green Center for Reproductive Biology Sciences, UT Southwestern Medical Center, 6001 Forest Park Road, Dallas, TX 75390. Tel.: (214) 645-6279; Fax: (214) 645-6276;
    Affiliations
    Departments of Pharmacology, University of Texas Southwestern Medical Center in Dallas, Texas

    Cecil H & Ida Green Center for Reproductive Biology Sciences, University of Texas Southwestern Medical Center in Dallas, Texas
    Search for articles by this author
  • Author Footnotes
    * This work was supported by NIH grants RO1HD036022 and RO1HD053889 from the National Institute of Child Health and Human Development; R24RR03232601 from the National Center for Research Resources; R24OD011108 from the Office of the Director.
    This article contains supplemental Figs S1 to S3, Tables S1 and S2, Data S1 to S4, and Discussion.
Open AccessPublished:August 12, 2013DOI:https://doi.org/10.1074/mcp.M113.030585
      Spermiogenesis is a postmeiotic process that drives development of round spermatids into fully elongated spermatozoa. Spermatid elongation is largely controlled post-transcriptionally after global silencing of mRNA synthesis from the haploid genome. Here, rats that differentially express EGFP from a lentiviral transgene during early and late steps of spermiogenesis were used to flow sort fractions of round and elongating spermatids. Mass-spectral analysis of 2D gel protein spots enriched >3-fold in each fraction revealed a heterogeneous RNA binding proteome (hnRNPA2/b1, hnRNPA3, hnRPDL, hnRNPK, hnRNPL, hnRNPM, PABPC1, PABPC4, PCBP1, PCBP3, PTBP2, PSIP1, RGSL1, RUVBL2, SARNP2, TDRD6, TDRD7) abundantly expressed in round spermatids prior to their elongation. Notably, each protein within this ontology cluster regulates alternative splicing, sub-cellular transport, degradation and/or translational repression of mRNAs. In contrast, elongating spermatid fractions were enriched with glycolytic enzymes, redox enzymes and protein synthesis factors. Retrogene-encoded proteins were over-represented among the most abundant elongating spermatid factors identified. Consistent with these biochemical activities, plus corresponding histological profiles, the identified RNA processing factors are predicted to collectively drive post-transcriptional expression of an alternative exome that fuels finishing steps of sperm maturation and fitness.
      Post-transcriptional regulation of gene expression is essential for cells to transition into and out of distinct developmental, physiological, and pathological states (
      • David C.J.
      • Chen M.
      • Assanah M.
      • Canoll P.
      • Manley J.L.
      HnRNP proteins controlled by c-Myc deregulate pyruvate kinase mRNA splicing in cancer.
      ,
      • Kir S.
      • Beddow S.A.
      • Samuel V.T.
      • Miller P.
      • Previs S.F.
      • Suino-Powell K.
      • Xu H.E.
      • Shulman G.I.
      • Kliewer S.A.
      • Mangelsdorf D.J.
      FGF19 as a postprandial, insulin-independent activator of hepatic protein and glycogen synthesis.
      ,
      • Licatalosi D.D.
      • Yano M.
      • Fak J.J.
      • Mele A.
      • Grabinski S.E.
      • Zhang C.
      • Darnell R.B.
      Ptbp2 represses adult-specific splicing to regulate the generation of neuronal precursors in the embryonic brain.
      ,
      • Zhu H.
      • Shyh-Chang N.
      • Segre A.V.
      • Shinoda G.
      • Shah S.P.
      • Einhorn W.S.
      • Takeuchi A.
      • Engreitz J.M.
      • Hagan J.P.
      • Kharas M.G.
      • Urbach A.
      • Thornton J.E.
      • Triboulet R.
      • Gregory R.I.
      • Altshuler D.
      • Daley G.Q.
      The Lin28/let-7 axis regulates glucose metabolism.
      ). Accordingly, post-transcriptional control of gene expression plays essential roles during gametogenesis and embryo development by helping cells undergo dynamic changes in fate and function (
      • Rouget C.
      • Papin C.
      • Boureux A.
      • Meunier A.C.
      • Franco B.
      • Robine N.
      • Lai E.C.
      • Pelisson A.
      • Simonelig M.
      Maternal mRNA deadenylation and decay by the piRNA pathway in the early Drosophila embryo.
      ,
      • Ghosh S.
      • Marchand V.
      • Gaspar I.
      • Ephrussi A.
      Control of RNP motility and localization by a splicing-dependent structure in oskar mRNA.
      ,
      • Kadyrova L.Y.
      • Habara Y.
      • Lee T.H.
      • Wharton R.P.
      Translational control of maternal Cyclin B mRNA by Nanos in the Drosophila germline.
      ,
      • Li Y.
      • Maines J.Z.
      • Tastan O.Y.
      • McKearin D.M.
      • Buszczak M.
      Mei-P26 regulates the maintenance of ovarian germline stem cells by promoting BMP signaling.
      ,
      • Henderson M.A.
      • Cronland E.
      • Dunkelbarger S.
      • Contreras V.
      • Strome S.
      • Keiper B.D.
      A germline-specific isoform of eIF4E (IFE-1) is required for efficient translation of stored mRNAs and maturation of both oocytes and sperm.
      ,
      • Flemr M.
      • Ma J.
      • Schultz R.M.
      • Svoboda P.
      P-body loss is concomitant with formation of a messenger RNA storage domain in mouse oocytes.
      ). In a classic model, transcriptionally inactive oocytes store large reserves of de-adenylated transcripts in a translationally repressed state (
      • Jackson R.J.
      • Standart N.
      Do the poly(A) tail and 3′ untranslated region control mRNA translation?.
      ). Then, in response to meiotic maturation, polyadenylation of the stored mRNAs signals their translation into maternal proteins required for early embryogenesis (
      • Jackson R.J.
      • Standart N.
      Do the poly(A) tail and 3′ untranslated region control mRNA translation?.
      ). Spermatozoan development is also known for its diverse post-transcriptional modes of gene expression (
      • Schmidt E.E.
      • Hanson E.S.
      • Capecchi M.R.
      Sequence-independent assembly of spermatid mRNAs into messenger ribonucleoprotein particles.
      ,
      • Kleene K.C.
      Patterns of translational regulation in the mammalian testis.
      ,
      • Yang J.
      • Medvedev S.
      • Reddi P.P.
      • Schultz R.M.
      • Hecht N.B.
      The DNA/RNA-binding protein MSY2 marks specific transcripts for cytoplasmic storage in mouse male germ cells.
      ,
      • Vemuganti S.A.
      • de Villena F.P.
      • O'Brien D.A.
      Frequent and recent retrotransposition of orthologous genes plays a role in the evolution of sperm glycolytic enzymes.
      ,
      • He C
      • Zuo Z.
      • Chen H.
      • Zhang L.
      • Zhou F.
      • Cheng H.
      • Zhou R.
      Genome-wide detection of testis- and testicular cancer-specific alternative splicing.
      ,
      • Harr B.
      • Turner L.M.
      Genome-wide analysis of alternative splicing evolution among Mus subspecies.
      ). However, in contrast to many oocyte mRNAs, translational activation in developing spermatozoa is commonly associated with poly-A tail shortening, rather than polyadenylation (
      • Kleene K.C.
      Poly(A) shortening accompanies the activation of translation of five mRNAs during spermiogenesis in the mouse.
      ,
      • Yanagiya A.
      • Delbes G.
      • Svitkin Y.V.
      • Robaire B.
      • Sonenberg N.
      The poly(A)-binding protein partner Paip2a controls translation during late spermiogenesis in mice.
      ,
      • Kleene K.C.
      • Distel R.J.
      • Hecht N.B.
      Translational regulation and deadenylation of a protamine mRNA during spermiogenesis in the mouse.
      ).
      In the typical fertile human male, energy is expended for net biosynthesis of >35 million new spermatozoa each day (
      • Amann R.P.
      The cycle of the seminiferous epithelium in humans: a need to revisit?.
      ). This equates to >25,000 spermatozoa generated/male/minute throughout a 65 year reproductive life span to help parent an average family of ∼3 children (
      • Amann R.P.
      The cycle of the seminiferous epithelium in humans: a need to revisit?.
      ). In adult males, haploid gametes that form spermatozoa are continuously being produced from spermatogonial stem cells in the testes through the developmental process of spermatogenesis (
      • Amann R.P.
      The cycle of the seminiferous epithelium in humans: a need to revisit?.
      ,
      • Oatley J.M.
      • Brinster R.L.
      The germline stem cell niche unit in mammalian testes.
      ). During spermatogenesis, a subset of spermatogonial stem cells within the basal compartment of seminiferous tubules give rise to differentiating spermatogonia that amplify in number through a series of mitotic divisions. Differentiating spermatogonia then initiate meiosis and traverse the Sertoli cell blood-testis barrier to enter the seminiferous tubules as newly formed spermatocytes (
      • Dym M.
      • Fawcett D.W.
      The blood-testis barrier in the rat and the physiological compartmentation of the seminiferous epithelium.
      ). Once inside the adluminal compartment of the seminiferous epithelium, spermatocytes complete meiosis to generate nascent haploid germ cells, termed round spermatids (
      • Clermont Y.
      Kinetics of spermatogenesis in mammals: seminiferous epithelium cycle and spermatogonial renewal.
      ). As requisite for round spermatids to mature into spermatozoa, they must dramatically transform anatomically in size, shape, and organelle composition through the post-meiotic process of spermiogenesis (
      • Clermont Y.
      Kinetics of spermatogenesis in mammals: seminiferous epithelium cycle and spermatogonial renewal.
      ).
      In rodents, newly formed round spermatids undergo up to 19 well-defined steps of spermiogenesis before being shed into the lumen of the seminiferous epithelium as fully elongated, yet functionally immature spermatozoa (reviewed by Yves Clermont) (
      • Clermont Y.
      Kinetics of spermatogenesis in mammals: seminiferous epithelium cycle and spermatogonial renewal.
      ). Acrosome biogenesis is a sperm-specific process adapted from the Golgi complex, and is commonly used to classify spermatids at distinct steps of spermiogenesis as they mature during progressive stages of the seminiferous epithelium cycle (
      • Clermont Y.
      Kinetics of spermatogenesis in mammals: seminiferous epithelium cycle and spermatogonial renewal.
      ). The periodic acid-Schiff's (PAS) staining method differentially highlights step-specific morphological changes to the acrosome and nucleus of developing spermatids (
      • Clermont Y.
      Kinetics of spermatogenesis in mammals: seminiferous epithelium cycle and spermatogonial renewal.
      ). It should be stressed that linear “steps” in germ cell development are different from the “stages” of the seminiferous epithelium cycle (
      • Clermont Y.
      Kinetics of spermatogenesis in mammals: seminiferous epithelium cycle and spermatogonial renewal.
      ). This is because the epithelial stages are defined by physical associations formed between different testis cell types during an epithelial cycle (see Supplemental Fig. 1 for review). Each unique cellular association defining a respective stage is organized vertically in space within a tubular segment by successive generations of spermatogenic cell types (
      • Clermont Y.
      Kinetics of spermatogenesis in mammals: seminiferous epithelium cycle and spermatogonial renewal.
      ). Developmental gaps between each generation of germ cells comprising a given stage is defined by the time taken to complete one cycle of the seminiferous epithelium (i.e. ∼12.9 days/cycle in rats) (
      • Clermont Y.
      Kinetics of spermatogenesis in mammals: seminiferous epithelium cycle and spermatogonial renewal.
      ). As such, each epithelial stage merely represents subsequent snap-shots in cycle time within the same seminiferous tubule segment (Supplemental Fig. 1).
      In rodents, it is estimated that >5% of mRNAs are specifically expressed to support the meiotic and post-meiotic processes of sperm development and fertilization (
      • Eddy E.M.
      Male germ cell gene expression.
      ,
      • Schultz N.
      • Hamra F.K.
      • Garbers D.L.
      A multitude of genes expressed solely in meiotic or postmeiotic spermatogenic cells offers a myriad of contraceptive targets.
      ,
      • Johnston D.S.
      • Wright W.W.
      • Dicandeloro P.
      • Wilson E.
      • Kopf G.S.
      • Jelinsky S.A.
      Stage-specific gene expression is a fundamental characteristic of rat spermatogenic cells and Sertoli cells.
      ). This includes numerous testis-specific isoforms of metabolic enzymes required for sperm function (
      • Miki K.
      • Qu W.
      • Goulding E.H.
      • Willis W.D.
      • Bunch D.O.
      • Strader L.F.
      • Perreault S.D.
      • Eddy E.M.
      • O'Brien D.A.
      Glyceraldehyde 3-phosphate dehydrogenase-S, a sperm-specific glycolytic enzyme, is required for sperm motility and male fertility.
      ,
      • Danshina P.V.
      • Geyer C.B.
      • Dai Q.
      • Goulding E.H.
      • Willis W.D.
      • Kitto G.B.
      • McCarrey J.R.
      • Eddy E.M.
      • O'Brien D.A.
      Phosphoglycerate kinase 2 (PGK2) is essential for sperm function and male fertility in mice.
      ,
      • Odet F.
      • Duan C.
      • Willis W.D.
      • Goulding E.H.
      • Kung A.
      • Eddy E.M.
      • Goldberg E.
      Expression of the gene for mouse lactate dehydrogenase C (Ldhc) is required for male fertility.
      ). Spermatogenic cells also express an unusually high percentage of retrogenes, a subset of which encode glycolytic enzymes hypothesized to have been selected by enhancing sperm fitness (
      • Vemuganti S.A.
      • de Villena F.P.
      • O'Brien D.A.
      Frequent and recent retrotransposition of orthologous genes plays a role in the evolution of sperm glycolytic enzymes.
      ). Additionally, spermatogenic cells express a most diverse array of alternatively processed mRNAs unique to the germline (
      • He C
      • Zuo Z.
      • Chen H.
      • Zhang L.
      • Zhou F.
      • Cheng H.
      • Zhou R.
      Genome-wide detection of testis- and testicular cancer-specific alternative splicing.
      ,
      • Harr B.
      • Turner L.M.
      Genome-wide analysis of alternative splicing evolution among Mus subspecies.
      ). Still, global silencing of transcription occurs during spermatid elongation as haploid nuclei remodel their chromatin into a hyper-compacted state within the spermatozoan head piece (
      • Monesi V.
      Ribonucleic Acid Synthesis during Mitosis and Meiosis in the Mouse Testis.
      ,
      • Kierszenbaum A.L.
      • Tres L.L.
      Structural and transcriptional features of the mouse spermatid genome.
      ). As a result, major fractions of “silenced” transcripts are stored for up to a week in messenger ribonucleic acid particles (mRNPs) until factors in the elongation phase of spermiogenesis trigger their translation to support final steps in spermatozoan development and fertilization (
      • Schmidt E.E.
      • Hanson E.S.
      • Capecchi M.R.
      Sequence-independent assembly of spermatid mRNAs into messenger ribonucleoprotein particles.
      ,
      • Kleene K.C.
      Patterns of translational regulation in the mammalian testis.
      ,
      • Yang J.
      • Medvedev S.
      • Reddi P.P.
      • Schultz R.M.
      • Hecht N.B.
      The DNA/RNA-binding protein MSY2 marks specific transcripts for cytoplasmic storage in mouse male germ cells.
      ). Thus, sperm development provides an extraordinary system to study molecular events that regulate post-transcriptional gene expression.
      Here, we present mass spectral analysis of proteins distinctly detected in fractions of round and elongating spermatids collected by flow cytometry using Green Elongating Spermatid (GESptd1)
      The abbreviations used are:GESptd1 rats
      green elongating spermatid rats
      EGFP
      enhanced green fluorescent protein
      CMV
      cytomegalovirus
      hnRNP
      heterogeneous nuclear ribonucleoprotein
      mRNP
      messenger ribonucleoprotein particle
      LTR
      long terminal repeat
      UTR
      untranslated region.
      transgenic rats. This cell sorting approach was made possible because GESptd1 rats differentially express EGFP post-transcriptionally from viral elements by mechanisms that mimic many spermatid genes. Protein profiles revealed by these studies demonstrate global down-regulation of an abundant RNA binding proteome expressed prior to up-regulation of metabolic proteins during spermatid elongation. Biochemical activities reported for these RNA-processing factors shed light on how the multiplicity of germline specific mRNA variants are uniquely generated, stored and then expressed to drive spermatozoan biology following transcriptional silencing of the male haploid genome.

      MATERIALS AND METHODS

      Complete methodologies used for spermatid isolation, Western blot, Northern blot, immunofluorescence, histological staining, in situ hybridization, and sucrose gradient fractionation are detailed in Supplementary Information.

      Animal Protocols

      Protocols for use of rats in this study were approved by the Institutional Animal Care and Use Committee (IACUC) at UT-Southwestern Medical Center in Dallas, as certified by the Association for Assessment and Accreditation of Laboratory Animal Care International (AALAC).

      Transgenic Rats

      GESptd1–12 transgenic rat strains [i.e. S.D.-Tg(ESptd-EGFP)Fkh1−12] were produced in Sprague-Dawley (Harlan, Inc.) backgrounds using freshly isolated laminin-binding spermatogonia that were treated overnight in culture with lentiviral particles (produced using plasmids pHR'-CMV-EGFP-SIN18, pBH10ψE, and pLVSV-G) prior to transplantation into recipient testes, as described (
      • Hamra F.K.
      • Gatlin J.
      • Chapman K.M.
      • Grellhesl D.M.
      • Garcia J.V.
      • Hammer R.E.
      • Garbers D.L.
      Production of transgenic rats by lentiviral transduction of male germ-line stem cells.
      ). Lentiviral transgene copy number/rat strain was verified by Southern blots (
      • Hamra F.K.
      • Gatlin J.
      • Chapman K.M.
      • Grellhesl D.M.
      • Garcia J.V.
      • Hammer R.E.
      • Garbers D.L.
      Production of transgenic rats by lentiviral transduction of male germ-line stem cells.
      ). Lentiviral integration sites for GESptd1 and GESptd2 rat strains were identified by splinkerette PCR using genomic DNA digested with HaeIII and ligated to a splinker adapter. Two rounds of PCR were performed using lentiviral vector-specific primers to either the 5′ LTR or the 3′ LTR in combination with linker-specific primers. 5′-LTR-specific primers: “outer” 5′-ctctcgcacccatctctctc-3′; “inner” 5′-tttttggcgtactcaccagtc-3′; 3′-LTR-specific primers: “outer” 5′-gaattcgagctcggtaccttt-3′; “inner” 5′-caatgacttacaaggcagctgta-3′.

      Mass Spectrometry and Analyses

      Round and elongating spermatid fractions were collected from adult rats between 165–240 days of age by flow cytometry, as described. A total ∼2 × 107 cells for each respective, EGFP Dim and Bright fraction were collected from five cumulative sorts. Post-sort, each fraction was washed twice with PBS and centrifugation at ∼200 × g and pellets were flash frozen in liquid nitrogen and then stored at −80 °C. Frozen pellets were shipped on dry ice to Applied Biomics, Hayward, CA, for 2D DIGE expression profiling. Protein from pooled round spermatid fractions were labeled with Cy3, whereas proteins from pooled elongating spermatid fractions were labeled with Cy5 prior to fractionation. 2D gels were scanned with a Typhoon image scanner (G.E. Healthcare Life Sciences, Inc.) and images were generated with ImageQuant software (G.E. Healthcare Life Sciences, Inc.). Differential protein expression profiles were analyzed with Decyder software (G.E. Healthcare Life Sciences, Inc.). Spots of interest were picked with the Ettan Spot Picker (G.E. Healthcare Life Sciences, Inc.), treated with trypsin and identified by MALDI TOF/TOF mass spectrometry (Applied Biosystems, Foster City, CA) using the non-redundant NCBI database to search 111,467 entries (October, 2010) allowing 1 missing cleavage/entry including the following variable modifications: carbamidomethylation of cysteine and oxidation of methionine. Mass tolerance for precursor ions and fragment ions were set at 100ppm and 0.5 Da, respectively; Cutoff score/expectation value for accepting individual MS/MS spectra = 20; noise threshold <5%. Accession numbers of identified proteins were imported into the Database for Annotation, Visualization, and Integrated Discovery (DAVID) for pathway analysis (http://david.abcc.ncifcrf.gov/). Peaklist-generating software and release version was GPS Explorer™ v3.6; search engine and release version MASCOT 2.0 (Matrix Science).

      RESULTS

      Spermiogenic Expression of EGFP in Rats

      Novel strains of transgenic rats were generated by transducing donor spermatogonial stem cells with a self-inactivating, lentiviral vector (
      • Hamra F.K.
      • Gatlin J.
      • Chapman K.M.
      • Grellhesl D.M.
      • Garcia J.V.
      • Hammer R.E.
      • Garbers D.L.
      Production of transgenic rats by lentiviral transduction of male germ-line stem cells.
      ), and evaluated as animal models for studying sperm development. The lentiviral vector used was designed to express EGFP from an internal cytomegalovirus (CMV) promoter (Fig. 1A) (
      • Hamra F.K.
      • Gatlin J.
      • Chapman K.M.
      • Grellhesl D.M.
      • Garcia J.V.
      • Hammer R.E.
      • Garbers D.L.
      Production of transgenic rats by lentiviral transduction of male germ-line stem cells.
      ). Consistent with reports in mice (
      • Dyck M.K.
      • Ouellet M.
      • Gagn M.
      • Petitclerc D.
      • Sirard M.A.
      • Pothier F.
      Testes-specific transgene expression in insulin-like growth factor-I transgenic mice.
      ,
      • Mehta A.K.
      • Majumdar S.S.
      • Alam P.
      • Gulati N.
      • Brahmachari V.
      Epigenetic regulation of cytomegalovirus major immediate-early promoter activity in transgenic mice.
      ), testicular expression of EGFP was visualized by fluorescence microscopy in 7 of 12 (∼58%) rat strains generated (Fig. 1B). Genomic sites where the lentiviral vector integrated were defined for two of the single-copy strains expressing highest relative levels of EGFP in testes (i.e. GESptd1 and GESptd2 rats) (Fig. 1C). In all 7 strains where transgene expression could be visualized microscopically, a consistent pattern of fluorescence was observed in spermatogenic cells nearest to the luminal centers of seminiferous tubules (Figs. 2A, 2B). Northern and Western blot analyses of GESptd1 rats demonstrated selective expression of the reporter transgene in testes (Fig. 2C, 2D). The same Western blot profile was detected for EGFP in GESptd2 rat tissues (not shown). This localization pattern indicated that EGFP was being expressed in the germline during late steps of spermiogenesis, thus, coining the name “Green Elongating Spermatid rats.”
      Figure thumbnail gr1
      Fig. 1Transgenic rats produced with lentiviral reporter vector. A, Diagram of self-inactivating lentiviral vector, pHR'CMV-EGFP-SIN18 (
      • Zufferey R.
      • Dull T.
      • Mandel R.J.
      • Bukovsky A.
      • Quiroz D.
      • Naldini L.
      • Trono D.
      Self-inactivating lentivirus vector for safe and efficient in vivo gene delivery.
      ,
      • Gatlin J.
      • Padgett A.
      • Melkus M.W.
      • Kelly P.F.
      • Garcia J.V.
      Long-term engraftment of nonobese diabetic/severe combined immunodeficient mice with human CD34+ cells transduced by a self-inactivating human immunodeficiency virus type 1 vector.
      ), used to generate transgenic rat lines by in vitro transduction of donor spermatogonia. Depicted are the U3, R, and U5 regions of the 5′ and 3′ long terminal repeat regions (LTRs); surface glycoprotein element (G); rev response element (RRE); splice acceptor sequence (SA); splice donor sequence (S.D.); cytomegalovirus promoter (CMV); enhanced green fluorescent protein (EGFP). Also shown is the region deleted from the 3′ LTR (ΔU3), which helps to disrupt viral replication (
      • Zufferey R.
      • Dull T.
      • Mandel R.J.
      • Bukovsky A.
      • Quiroz D.
      • Naldini L.
      • Trono D.
      Self-inactivating lentivirus vector for safe and efficient in vivo gene delivery.
      ). B, Fluorescence microscopy of EGFP expression (green) in seminiferous tubules of transgenic rat strains GESptd1, GESptd2, and GESptd3 hemizygous for a single copy of the lentiviral construct. Wild-type rat seminiferous tubules illustrate background fluorescence. Right panels show respective bright field images of seminiferous tubules from each rat. Scale bar, 1 mm. C, Lentiviral transgenes integrated between the Trmp6 and Rorb genes in 1q43 of chromosomes 1 in GESptd1 rats, and between the Sema5a and LOC100360282 genes in 2q23 on Chromosome 2 in GESptd2 rats.
      Figure thumbnail gr2
      Fig. 2GESptd rats express EGFP specifically in spermatids. A, Phase contrast and green fluorescence microscopy images identify EGFP expression (green) near the lumen of seminiferous tubules in transgenic rat strains GESptd1 and GESptd2. Scale bars, 200 μm. B, Histological cross-section of seminiferous tubules form GESptd1 rats illustrates EGFP expression in elongating spermatids (green). Image overlay shows nuclear with Hoechst 33342 dye in all testis cells (blue). Scale Bar, 200 μm. C, Top: Northern analysis of EGFP expression in tissues of transgenic rat strain, GESptd1. T, testis; M, skeletal muscle; I, small intestine; K, kidney; Lv, liver; Lg, lung; H, heart. Germ Cell Specific EGFP transgenic rat testis (GCS) (
      • Cronkhite J.T.
      • Norlander C.
      • Furth J.K.
      • Levan G.
      • Garbers D.L.
      • Hammer R.E.
      Male and female germline specific expression of an EGFP reporter gene in a unique strain of transgenic rats.
      ); Wildtype rat testis (WT). Total RNA isolated from day 45 homozygous transgenic and wildtype littermates. Bottom: Same blot after stripping and re-probing for Rn18S (◀ ∼1.9 kb). D, Western analysis of EGFP (◀ ∼27 kDa) and GAPDH (◀ ∼55 kDa) expression in tissues of the transgenic (GESptd1 and GCS) and wildtype rats used in panel “A”.
      To more precisely define steps of spermiogenesis expressing EGFP in GESptd1 rats, PAS-Feulgen staining and immunofluorescence were performed on parallel cross-sections prepared from their testes. Expression of EGFP was most abundant in the cytoplasm of step 12–19 spermatids during stages XI to VIII of the seminiferous epithelium cycle (Fig. 3). By comparison, the spermatocyte and round spermatid transcription factor, CREM-Τ (
      • Delmas V.
      • van der Hoorn F.
      • Mellstrom B.
      • Jegou B.
      • Sassone-Corsi P.
      Induction of CREM activator proteins in spermatids: down-stream targets and implications for haploid germ cell differentiation.
      ), was most abundant in late pachytene spermatocytes during stages XI to XIV, and in step 1 to 11 spermatids during stages I to XI of the epithelial cycle (Fig. 3). However, an antisense probe to EGFP specifically hybridized to RNA in both meiotic and post-meiotic spermatogenic cells of GESptd1 rat testes (Fig 4A). Silver grains produced from the EGFP antisense probe became apparent over pachytene spermatocytes after stage IV of the epithelial cycle and then most densely labeled subsequent steps in spermatocyte and spermatid development (Fig. 4B). Thus, in adult rats, EGFP expression appeared to be up-regulated at the transcriptional level during meiotic prophase-I (Fig. 4C; blue gradient bar) and at the translational level during elongation steps of spermiogenesis (Fig. 4C; green gradient bar).
      Figure thumbnail gr3
      Fig. 3GESptd rats robustly express EGFP in elongating spermatids. Testis cross sections from GESptd1 rats show EGFP expressed during elongating steps 11–19 of spermiogenesis (green fluorescence); by comparison, immunolabeling detects the transcription factor, Crem-tau, in round spermatids during Steps 1–11 of spermiogenesis (red fluorescence). Right panel shows respective images of PAS-Feulgen staining in adjacent histological sections to define stages of the seminiferous epithelium cycle (Roman numerals). Scale Bar, 100 μm.
      Figure thumbnail gr4
      Fig. 4GESptd1 rats initiate transcription of EGFP in spermatocytes. A, In Situ hybridization of a radio-labeled, EGFP antisense probe to testis cells from GESptd1 rats, but not wildtype rats. Scale Bar, 200 μm. B, Autoradiographic detection of silver grains developing from the EGFP antisense probe most densely formed over late prophase-I spermatocytes (pachytene-secondary) and step 1–17 spermatids. Sertoli Cell (SC); zygotene spermatocyte (Z); pachytene spermatocyte (P); secondary spermatocyte (2nd); round spermatids (RS); elongating spermatids (ES). Scale Bar, 30 μm. C, Illustration of GESptd1 rat transgene expression in spermatozoan progenitor cell types based on in situ hybridization studies in adult rats (top blue and green gradient bars). Note: EGFP transcripts are depicted expressed in GESptd1 rats several days prior to EGFP. As a comparison, arrows at the bottom show relative postnatal days (D) when first generation spermatogenic cell types are initially formed in rat testes. D, Northern analysis of EGFP (◀ ∼1.2 kb), DAZL (◀ ∼3 kb), and Rn18S (◀ ∼1.9 kb) transcripts in GESptd1 rat testes on postnatal days 10, 21, 25, 30, 35, 40, 45; and in D45 testes of wild-type rats (WT). E, Western analysis of EGFP (◀ ∼27 kDa), DAZL (◀ ∼36 kDa) and TUBA1a (◀ ∼55 kDa) in GESptd1 rat testes on postnatal days 21, 25, 30, 35, 40, 45; and in D45 testes of wild-type rats (WT).
      The relative timing of EGFP expression in GESptd1 rats was analyzed during the onset of spermiogenesis, which initiates in rats after ∼25 days of age when round spermatids are initially produced through meiotic divisions of first-generation spermatocytes (Fig. 4C; arrows). During postnatal development, EGFP transcripts were first detectable by Northern analysis in testes of 30-day-old rats, which then steadily increased in relative abundance as a function of age (Fig. 4D). Subsequent to detection of EGFP mRNA, EGFP was initially detected by Western blot on day 40, after development of the first elongating spermatids, and proceeded to increase in abundance by day 45 (Fig. 4E). By comparison, detection of testicular transcripts and protein encoded by the deleted in azoospermia-like (dazl) gene each peaked by postnatal day 21, consistent with expression of DAZL in spermatogonia and spermatocytes (Figs. 4D, 4E) (
      • Reijo R.A.
      • Dorfman D.M.
      • Slee R.
      • Renshaw A.A.
      • Loughlin K.R.
      • Cooke H.
      • Page D.C.
      DAZ family proteins exist throughout male germ cell development and transit from nucleus to cytoplasm at meiosis in humans and mice.
      ). Accordingly, transcription and translation of the GESptd1 transgene appeared uncoupled postnatally during sperm development.
      However, because mature pachytene spermatocytes are present in 21–25-day-old GESptd1 rats, signals for EGFP detection by Northern blot in total testis lysates at these ages did not correlate precisely with in situ hybridization silver grains formed over pachytene spermatocytes in adult rat testes (Fig. 4B). This could reflect differences in detection thresholds for Northern versus in situ hybridization assays at these respective ages because 1st generation spermatocyte and spermatid populations take time to accumulate in abundance relative to other testis cell types (see supplemental Materials and Methods), and/or, actual differences in the onset of EGFP expression by pachytene spermatocytes in adolescent versus adult rats.

      Packaging EGFP into mRNPs

      To help explain the presumptive delay between expression of EGFP transcripts and EGFP during spermiogenesis, lysates prepared from GESptd1 rat testes were fractionated over sucrose density gradients (Fig. 5). Most EGFP transcripts ranging from 0.8–1.5 kb were detected in fractions near the top of the gradient (Fractions 1–3), whereas, lower molecular weight EGFP transcripts of ∼1 kb migrated into EDTA-sensitive polysome-like fractions at the bottom of the gradient (Fractions 9–11) (Fig. 5A). Similarly, the mRNP marker protein, MSY2 (
      • Yang J.
      • Medvedev S.
      • Reddi P.P.
      • Schultz R.M.
      • Hecht N.B.
      The DNA/RNA-binding protein MSY2 marks specific transcripts for cytoplasmic storage in mouse male germ cells.
      ), comigrated predominantly with a majority of EGFP transcripts in fraction 2 and 3, but was also detected in fractions throughout the sucrose gradient (Fig. 5B). By contrast, a majority of transcripts encoding DAZL migrated in Fractions 9–11 of the sucrose gradient, and thus, were enriched in the polysome-like fractions (Fig. 5C). Hybridization of total RNA from testes of 45-day-old rats with oligo-dt18−24 and treatment with RNase-H increased the relative mobility of EGFP transcripts, shifting a larger portion into bands migrating closer in size to, or less than the ∼1 kb transcript within the polysome-like fractions (Fig. 5A). RNase-H treatment also increased the relative mobility of DAZL transcripts (Fig. 5C). Thus, larger EGFP transcripts potentially represent longer poly-adenylated species, which like many endogenous spermatid genes, could be translationally repressed in mRNPs of spermatocytes and/or round spermatids prior to de-adenylation and incorporation into polysomes for translation in elongating spermatids (
      • Kleene K.C.
      Patterns of translational regulation in the mammalian testis.
      ,
      • Kleene K.C.
      Poly(A) shortening accompanies the activation of translation of five mRNAs during spermiogenesis in the mouse.
      ,
      • Kleene K.C.
      • Distel R.J.
      • Hecht N.B.
      Translational regulation and deadenylation of a protamine mRNA during spermiogenesis in the mouse.
      ). However, such modifications to GESptd1 EGFP poly-A tails in nonpolysome and polysome fractions will require verification by sequence analysis.
      Figure thumbnail gr5
      Fig. 5EGFP transcripts localize to mRNP-like particles in GESptd1 rats. A, Northern blot of EGFP transcripts in whole testis lysates from an adult GESptd1 rat after centrifugation into a sucrose density gradient. RNase-H (RH) treated samples after hybridizing with Oligo(dt). Note: a majority of EGFP transcripts are detected in low density mRNP-like particles. Controls (Ct) = testis lysates prepared from GCS-EGFP transgenic (+) and wildtype (−) rats. B, Western blot of MSY2 protein in whole testis lysates (T) from an adult GESptd1 rat after centrifugation into a sucrose density gradient. C, Northern blot of DAZL transcripts in whole testis lysates from an adult GESptd1 rat after centrifugation into a sucrose density gradient. Note: a majority of DAZL transcripts are detected in polysome-like particles. Controls (Ct) = testis lysates prepared from tgGCS-EGFP transgenic (+) and wildtype (−) rats.

      Flow Sorting Round and Elongating Spermatids

      Based on robust expression of EGFP in GESptd1 rat spermatids, we postulated that enriched fractions of elongating spermatids could be isolated from their testes by flow cytometry to facilitate biochemical analyses on spermiogenesis. Testis cells from GESptd1 rats were enzymatically disaggregated, and then two major peaks of EGFP-positive cells were identified by flow cytometry and collected by cell sorting (Figs. 6A, 6B). A dimmer peak of EGFP+ cells expressed ∼eightfold lower levels of EGFP when compared with the second, brighter peak of EGFP+ cells (Figs. 6A, 6B). However, only 34.7 ± 3.3% (S.E.; n = 3) of collected EGFP-Bright cells were nucleated; whereas, 92 ± 3.6% (S.E.; n = 3) of collected EGFP-Dim cells were nucleated when scored post-sorting using Hoechst 33342 dye. Based on transgene expression (Fig. 2, Fig. 3, Fig. 4), nuclear morphologies (Fig. 6C; supplemental Fig. S2) and testis cell markers (supplemental Figs. S3A, S3B; supplemental Tables S1, S2), the EGFP-Dim fraction was enriched with round spermatids; whereas, the EGFP-Bright fraction was enriched in both elongating spermatid cytoplasts (also termed residual bodies), in addition to intact elongating spermatids. Both spermatid fractions were depleted of somatic testis cells (supplemental Fig. S3A, S3B).
      Figure thumbnail gr6
      Fig. 6Flow sorting round and elongating spermatids from GESptd1 rats. A, Flow cytometry analysis of EGFP+ testis cells sorted from GESptd1 rats. Left: Total EGFP+ testis cell population gated within region 1 (R1) of forward and side scatter plot (R1 = 31.9 ± 5.7% total) Center: Background green fluorescence gates G1 and G2 set using a wild-type rat littermate; Right: Dominant peaks of EGFP+ GESptd1 rat testis cells sorted from gates G1 and G2 (G1 = 27.8 ± 3.4% R1; G2 = 31.6 ± 0.9% R1); ±S.D. n = 5 sorts. B, Relative purities of “EGFP-Dim” and “EGFP-Bright” GESptd1 rat cells post-sorted from gates G1 and G2, respectively. C, Top: Images of testis cells isolated from wildtype and GESptd1 rats before and after sorting into EGFP-Bright and EGFP-Dim fractions. Bottom: Hoechst 33342 nuclear labeling (blue) of cells in top panels. Scale bars, 8 μm.

      Mass Spectral Analysis of Spermatid Proteins

      Due to the relative purity and abundance of each sorted population, we conducted mass-spectral analysis to identify proteins differentially sorted into the EGFP-Dim and EGFP-Bright spermatid fractions after purification by 2D-electrophoresis (Fig. 7). Respective EGFP-Dim and EGFP-Bright spermatid fractions from five total sorts were pooled for the analysis (∼20 million cells/fraction). A total of 88 differentially expressed (≥twofold) protein-like factors were submitted for analysis (Fig. 7A) (supplemental Data Files S1–S3). Consistent with histological and cytometric fluorescence signals in round and elongating spermatids, EGFP was found by mass spectrometry to be ∼10-fold more abundant in the EGFP-Bright fraction versus the EGFP-Dim fraction (Fig. 7B). Similarly, the EGFP signal was 9.4-fold higher in the EGFP-Bright fraction than the EGFP-Dim fraction by Western blot (supplemental Fig. S3A; supplemental Table S2); in contrast, EGFP transcript levels were similar, but slightly more abundant in the EGFP-Dim fraction than in the EGFP-Bright fraction based on qtPCR (supplemental Table S1). Thus, the GESptd1 rat transgene provided an internal standard for analyzing the relative expression of round and elongating spermatid proteins (Fig. 7B).
      Figure thumbnail gr7
      Fig. 7Analysis of differentially expressed rat spermatid proteins. A, Differential display of molecules in round and elongating spermatids from GESptd1 rats after separation by 2D gel electrophoresis. Round spermatids in EGFP Dim fractions were prelabeled with a green fluorescent dye; Elongating spermatids in the EGFP Bright fractions were prelabeled with a red dye. The circled molecules were submitted for sequence analysis by mass spectrometry. B, Relative abundance of fluorescently labeled molecules identified by mass spectrometry in fractions of round spermatids (EGFP Dim Spermatids; green), and elongating spermatids (EGFP Bright Spermatids; red) from GESptd1 rats. Ratios of elongating spermatid/round spermatid molecular fluorescence signal intensities (ES/RS) were plotted for positively identified rat proteins (symbols) changing ≥threefold.
      Sequences for 20 proteins in spots enriched ≥threefold in the EGFP-Dim fraction (Table I), and 28 proteins in spots enriched ≥3-fold in the EGFP-Bright fraction were positively identified (Table II). Histological localization studies for 26 of the 50 non-EGFP proteins listed in Table I, Table II were previously reported in round and elongating spermatids of rats or mice (
      • Kimura M.
      • Ishida K.
      • Kashiwabara S.
      • Baba T.
      Characterization of two cytoplasmic poly(A)-binding proteins, PABPC1 and PABPC2, in mouse spermatogenic cells.
      ,
      • Matsui M.
      • Horiguchi H.
      • Kamma H.
      • Fujiwara M.
      • Ohtsubo R.
      • Ogata T.
      Testis- and developmental stage-specific expression of hnRNP A2/B1 splicing isoforms, B0a/b.
      ,
      • Kamma H.
      • Horiguchi H.
      • Wan L.
      • Matsui M.
      • Fujiwara M.
      • Fujimoto M.
      • Yazawa T.
      • Dreyfuss G.
      Molecular characterization of the hnRNP A2/B1 proteins: tissue-specific expression and novel isoforms.
      ,
      • Guillaume E.
      • Evrard B.
      • Com E.
      • Moertz E.
      • Jegou B.
      • Pineau C.
      Proteome analysis of rat spermatogonia: reinvestigation of stathmin spatio-temporal expression within the testis.
      ,
      • Xu M.
      • Hecht N.B.
      Polypyrimidine tract binding protein 2 stabilizes phosphoglycerate kinase 2 mRNA in murine male germ cells by binding to its 3′UTR.
      ,
      • Hosokawa M.
      • Shoji M.
      • Kitamura K.
      • Tanaka T.
      • Noce T.
      • Chuma S.
      • Nakatsuji N.
      Tudor-related proteins TDRD1/MTR-1, TDRD6 and TDRD7/TRAP: domain composition, intracellular localization, and function in male germ cells in mice.
      ,
      • Tanaka T.
      • Hosokawa M.
      • Vagin V.V.
      • Reuter M.
      • Hayashi E.
      • Mochizuki A.L.
      • Kitamura K.
      • Yamanaka H.
      • Kondoh G.
      • Okawa K.
      • Kuramochi-Miyagawa S.
      • Nakano T.
      • Sachidanandam R.
      • Hannon G.J.
      • Pillai R.S.
      • Nakatsuji N.
      • Chuma S.
      Tudor domain containing 7 (Tdrd7) is essential for dynamic ribonucleoprotein (RNP) remodeling of chromatoid bodies during spermatogenesis.
      ,
      • Seidah N.G.
      • Day R.
      • Hamelin J.
      • Gaspar A.
      • Collard M.W.
      • Chretien M.
      Testicular expression of PC4 in the rat: molecular diversity of a novel germ cell-specific Kex2/subtilisin-like proprotein convertase.
      ,
      • Vasileva A.
      • Tiedau D.
      • Firooznia A.
      • Muller-Reichert T.
      • Jessberger. R
      Tdrd6 is required for spermiogenesis, chromatoid body architecture, and regulation of miRNA expression.
      ,
      • Yu Y.
      • Zhao C.
      • Lv Z.
      • Chen W.
      • Tong M.
      • Guo X.
      • Wang L.
      • Liu J.
      • Zhou Z.
      • Zhu H.
      • Zhou Q.
      • Sha J.
      Microinjection manipulation resulted in the increased apoptosis of spermatocytes in testes from intracytoplasmic sperm injection (ICSI) derived mice.
      ,
      • Weber P.
      • Cammas F.
      • Gerard C.
      • Metzger D.
      • Chambon P.
      • Losson R.
      • Mark M.
      Germ cell expression of the transcriptional co-repressor TIF1beta is required for the maintenance of spermatogenesis in the mouse.
      ,
      • Tanaka H.
      • Iguchi N.
      • Egydio de Carvalho C.
      • Tadokoro Y.
      • Yomogida K.
      • Nishimune Y.
      Novel actin-like proteins T-ACTIN 1 and T-ACTIN 2 are differentially expressed in the cytoplasm and nucleus of mouse haploid germ cells.
      ,
      • McCarrey J.R.
      • Berg W.M.
      • Paragioudakis S.J.
      • Zhang P.L.
      • Dilworth D.D.
      • Arnold B.L.
      • Rossi J.J.
      Differential transcription of Pgk genes during spermatogenesis in the mouse.
      ,
      • Tsunekawa N.
      • Matsumoto M.
      • Tone S.
      • Nishida T.
      • Fujimoto H.
      The Hsp70 homolog gene, Hsc70t, is expressed under translational control during mouse spermiogenesis.
      ,
      • Koga M.
      • Tanaka H.
      • Yomogida K.
      • Nozaki M.
      • Tsuchida J.
      • Ohta H.
      • Nakamura Y.
      • Masai K.
      • Yoshimura Y.
      • Yamanaka M.
      • Iguchi N.
      • Nojima H.
      • Matsumiya K.
      • Okuyama A.
      • Nishimune Y.
      Isolation and characterization of a haploid germ cell-specific novel complementary deoxyribonucleic acid; testis-specific homologue of succinyl CoA:3-Oxo acid CoA transferase.
      ,
      • Vemuganti S.A.
      • Bell T.A.
      • Scarlett C.O.
      • Parker C.E.
      • de Villena F.P.
      • O'Brien D.A.
      Three male germline-specific aldolase A isozymes are generated by alternative splicing and retrotransposition.
      ,
      • Shi H.J.
      • Wu A.Z.
      • Santos M.
      • Feng Z.M.
      • Huang L.
      • Chen Y.M.
      • Zhu K.
      • Chen C.L.
      Cloning and characterization of rat spermatid protein SSP411: a thioredoxin-like protein.
      ,
      • Sasagawa I.
      • Matsuki S.
      • Suzuki Y.
      • Iuchi Y.
      • Tohya K.
      • Kimura M.
      • Nakada T.
      • Fujii J.
      Possible involvement of the membrane-bound form of peroxiredoxin 4 in acrosome formation during spermiogenesis of rats.
      ,
      • Swiston J.
      • Hubberstey A.
      • Yu G.
      • Young D.
      Differential expression of CAP and CAP2 in adult rat tissues.
      ,
      • Fundele R.
      • Winking H.
      • Illmensee K.
      • Jagerbauer E.M.
      Developmental activation of phosphoglycerate mutase-2 in the testis of the mouse.
      ,
      • Lie-Venema H.
      • de Boer P.A.
      • Moorman A.F.
      • Lamers W.H.
      Role of the 5′ enhancer of the glutamine synthetase gene in its organ-specific expression.
      ,
      • Russell D.L.
      • Kim K.H.
      Expression of triosephosphate isomerase transcripts in rat testis: evidence for retinol regulation and a novel germ cell transcript.
      ,
      • Iida H.
      • Doiguchi M.
      • Yamashita H.
      • Sugimachi S.
      • Ichinose J.
      • Mori T.
      • Shibata Y.
      Spermatid-specific expression of Iba1, an ionized calcium binding adapter molecule-1, in rat testis.
      ,
      • Bush L.A.
      • Herr J.C.
      • Wolkowicz M.
      • Sherman N.E.
      • Shore A.
      • Flickinger C.J.
      A novel asparaginase-like protein is a sperm autoantigen in rats.
      ,
      • Nakamura N.
      • Mori C.
      • Eddy E.M.
      Molecular complex of three testis-specific isozymes associated with the mouse sperm fibrous sheath: hexokinase 1, phosphofructokinase M, and glutathione S-transferase mu class 5.
      ,
      • van Lith M.
      • Karala A.R.
      • Bown D.
      • Gatehouse J.A.
      • Ruddock L.W.
      • Saunders P.T.
      • Benham A.M.
      A developmentally regulated chaperone complex for the endoplasmic reticulum of male haploid germ cells.
      ,
      • Gitlits V.M.
      • Toh B.H.
      • Loveland K.L.
      • Sentry J.W.
      The glycolytic enzyme enolase is present in sperm tail and displays nucleotide-dependent association with microtubules.
      ,
      • Maywood E.S.
      • Chahad-Ehlers S.
      • Garabette M.L.
      • Pritchard C.
      • Underhill P.
      • Greenfield A.
      • Ebling F.J.
      • Akhtar R.A.
      • Kyriacou C.P.
      • Hastings M.H.
      • Reddy A.B.
      Differential testicular gene expression in seasonal fertility.
      ,
      • Held T.
      • Paprotta I.
      • Khulan J.
      • Hemmerlein B.
      • Binder L.
      • Wolf S.
      • Schubert S.
      • Meinhardt A.
      • Engel W.
      • Adham I.M.
      Hspa4l-deficient mice display increased incidence of male infertility and hydronephrosis development.
      ,
      • Moffit J.S.
      • Boekelheide K.
      • Sedivy J.M.
      • Klysik J.
      Mice lacking Raf kinase inhibitor protein-1 (RKIP-1) have altered sperm capacitation and reduced reproduction rates with a normal response to testicular injury.
      ). In all reported cases (i.e. 26 of 26 reports), immunolocalization of the identified proteins during spermiogenesis within testicular cross-sections and isolated spermatids were in agreement with results of this study (Table I, II). As additional standards, immunolabeling rat testis sections with antibodies to three of these proteins confirmed the mass-spectral results in this study by matching the reported localization of hnRNPK and GLUL in mice; and STMN in rats (Fig. 8) (
      • Guillaume E.
      • Evrard B.
      • Com E.
      • Moertz E.
      • Jegou B.
      • Pineau C.
      Proteome analysis of rat spermatogonia: reinvestigation of stathmin spatio-temporal expression within the testis.
      ,
      • Yu Y.
      • Zhao C.
      • Lv Z.
      • Chen W.
      • Tong M.
      • Guo X.
      • Wang L.
      • Liu J.
      • Zhou Z.
      • Zhu H.
      • Zhou Q.
      • Sha J.
      Microinjection manipulation resulted in the increased apoptosis of spermatocytes in testes from intracytoplasmic sperm injection (ICSI) derived mice.
      ,
      • Lie-Venema H.
      • de Boer P.A.
      • Moorman A.F.
      • Lamers W.H.
      Role of the 5′ enhancer of the glutamine synthetase gene in its organ-specific expression.
      ). To our knowledge, up to ∼56% (13 of 23) and ∼29% (8 of 27; excluding EGFP) of proteins in spots that differed by ≥threefold in relative abundance between respective round and elongating spermatid fractions represent newly reported spermiogenic factors (Table I, II). Moreover, in rat testis sections, PCBP1 (Fig. 9A) and EE2F (Fig. 9B) were selectively expressed in round and elongating spermatids, respectively. This was also consistent with mass-spectrometry results (Tables I, II). Interestingly, PCBP1 antibody labeling demonstrated a diffuse granular patter throughout the cytoplasm and in the nucleus, but was also concentrated in distinct germinal granule-like foci in spermatocytes and round spermatids; PCBP1 was also highly localized to a relatively large type of cytoplasmic granule in elongating spermatids (Fig. 9A). EE2F immunolabeling was most prominent in differentiating spermatogonia, preleptotene spermatocytes, step 14–17 elongating spermatids and small nuclear granules in Sertoli cells (Fig. 9B).
      Table IProteins enriched in round spermatid fraction. (Common Acronym); ES, Elongating Spermatid; RS, Round Spermatid; rg, retrogene-like containing 0–1 introns; p-rg, Proteins encoded by parent gene and retrogene identified in same protein spot. Supplemental Data File S1 contains corresponding protein accession numbers and mass spec data. Supplemental Data File S2 contains additional gene information
      Protein symbolRatio ES/RSProtein nameFunctionSpermatid localization
      Reported relative abundance observed histologically during spermiogenesis.
      Ref.
      c.r., current report.
      hnRNPA3
      RGSL1 also identified in same protein spot (c.r.); RGSL1 contains sequence identical to hnRNPA3.
      −25.9Heterogeneous nuclear ribonucleoprotein A3RNA Binding ProteinUnknownc.r.
      Pabpc1
      hnRNPM and PABPC4 also identified in same spot (c.r.).
      −20.5Poly(A) binding protein, cytoplasmic 1RNA Binding ProteinRS
      • Kimura M.
      • Ishida K.
      • Kashiwabara S.
      • Baba T.
      Characterization of two cytoplasmic poly(A)-binding proteins, PABPC1 and PABPC2, in mouse spermatogenic cells.
      hnRNPA2b1−15.1Heterogeneous nuclear ribonucleoprotein A2/B1RNA Binding ProteinRS
      • Matsui M.
      • Horiguchi H.
      • Kamma H.
      • Fujiwara M.
      • Ohtsubo R.
      • Ogata T.
      Testis- and developmental stage-specific expression of hnRNP A2/B1 splicing isoforms, B0a/b.
      ,
      • Kamma H.
      • Horiguchi H.
      • Wan L.
      • Matsui M.
      • Fujiwara M.
      • Fujimoto M.
      • Yazawa T.
      • Dreyfuss G.
      Molecular characterization of the hnRNP A2/B1 proteins: tissue-specific expression and novel isoforms.
      Stmn1−12.7Stathmin 1Signal TransductionRS
      • Guillaume E.
      • Evrard B.
      • Com E.
      • Moertz E.
      • Jegou B.
      • Pineau C.
      Proteome analysis of rat spermatogonia: reinvestigation of stathmin spatio-temporal expression within the testis.
      Ptbp2 (hnRNPI)−10.4Polypyrimidine tract-binding protein-like protein 2RNA Binding ProteinRS
      • Xu M.
      • Hecht N.B.
      Polypyrimidine tract binding protein 2 stabilizes phosphoglycerate kinase 2 mRNA in murine male germ cells by binding to its 3′UTR.
      Tdrd7−8.1Tudor domain-containing protein 7RNA ProcessingRS
      • Hosokawa M.
      • Shoji M.
      • Kitamura K.
      • Tanaka T.
      • Noce T.
      • Chuma S.
      • Nakatsuji N.
      Tudor-related proteins TDRD1/MTR-1, TDRD6 and TDRD7/TRAP: domain composition, intracellular localization, and function in male germ cells in mice.
      ,
      • Tanaka T.
      • Hosokawa M.
      • Vagin V.V.
      • Reuter M.
      • Hayashi E.
      • Mochizuki A.L.
      • Kitamura K.
      • Yamanaka H.
      • Kondoh G.
      • Okawa K.
      • Kuramochi-Miyagawa S.
      • Nakano T.
      • Sachidanandam R.
      • Hannon G.J.
      • Pillai R.S.
      • Nakatsuji N.
      • Chuma S.
      Tudor domain containing 7 (Tdrd7) is essential for dynamic ribonucleoprotein (RNP) remodeling of chromatoid bodies during spermatogenesis.
      Sarnp (CIP29)−7.9SAP domain-containing ribonucleoproteinRNA Binding ProteinUnknownc.r.
      Psip1−7.1PC4 and SFRS1 interacting protein 1DNA Binding & RNA ProcessingRS
      • Seidah N.G.
      • Day R.
      • Hamelin J.
      • Gaspar A.
      • Collard M.W.
      • Chretien M.
      Testicular expression of PC4 in the rat: molecular diversity of a novel germ cell-specific Kex2/subtilisin-like proprotein convertase.
      Tdrd6−7.0Tudor domain-containing protein 6RNA ProcessingRS
      • Hosokawa M.
      • Shoji M.
      • Kitamura K.
      • Tanaka T.
      • Noce T.
      • Chuma S.
      • Nakatsuji N.
      Tudor-related proteins TDRD1/MTR-1, TDRD6 and TDRD7/TRAP: domain composition, intracellular localization, and function in male germ cells in mice.
      ,
      • Vasileva A.
      • Tiedau D.
      • Firooznia A.
      • Muller-Reichert T.
      • Jessberger. R
      Tdrd6 is required for spermiogenesis, chromatoid body architecture, and regulation of miRNA expression.
      hnRNPK p-rg−6.5Heterogeneous nuclear ribonucleoprotein KRNA Binding ProteinRS
      • Yu Y.
      • Zhao C.
      • Lv Z.
      • Chen W.
      • Tong M.
      • Guo X.
      • Wang L.
      • Liu J.
      • Zhou Z.
      • Zhu H.
      • Zhou Q.
      • Sha J.
      Microinjection manipulation resulted in the increased apoptosis of spermatocytes in testes from intracytoplasmic sperm injection (ICSI) derived mice.
      Fam107b−6.2Hypothetical ProteinUnknownUnknownc.r.
      Pcbp3 (hnRNPE3)−6.0Poly(rC)-binding protein 3RNA Binding ProteinUnknownc.r.
      Pcbp1 (hnRNPE1)−5.7Poly(rC)-binding protein 1RNA Binding ProteinUnknownc.r.
      Ruvbl2−4.3RuvB-like 2DNA Helicase; RNA ProcessingUnknownc.r.
      hnRNPL−4.1Heterogeneous nuclear ribonucleoprotein LRNA Binding ProteinUnknownc.r.
      hnRPDL (JKTBP)−4.0Heterogeneous nuclear ribonucleoprotein d-LikeRNA Binding ProteinUnknownc.r.
      Eif2s1−3.8Eukaryotic translation initiation factor 2 subunit 1Translation FactorUnknownc.r.
      Trim28−3.6Transcription intermediary factor 1-betaTranscription FactorRS
      • Weber P.
      • Cammas F.
      • Gerard C.
      • Metzger D.
      • Chambon P.
      • Losson R.
      • Mark M.
      Germ cell expression of the transcriptional co-repressor TIF1beta is required for the maintenance of spermatogenesis in the mouse.
      Actl7b rg−3.6Actin-like protein 7BCytoskeletonRS
      • Tanaka H.
      • Iguchi N.
      • Egydio de Carvalho C.
      • Tadokoro Y.
      • Yomogida K.
      • Nishimune Y.
      Novel actin-like proteins T-ACTIN 1 and T-ACTIN 2 are differentially expressed in the cytoplasm and nucleus of mouse haploid germ cells.
      Ccdc104−3.1Coiled-coil domain-containing protein 104UnknownUnknownc.r.
      a Reported relative abundance observed histologically during spermiogenesis.
      b c.r., current report.
      c RGSL1 also identified in same protein spot (c.r.); RGSL1 contains sequence identical to hnRNPA3.
      d hnRNPM and PABPC4 also identified in same spot (c.r.).
      Table IIProteins enriched in elongating spermatid fraction. ES, Elongating Spermatid; RS, Round Spermatid; rg, retrogene-like containing 0–1 introns; p-rg, Proteins encoded by parent gene and retrogene identified in same protein spot. Supplemental Data File S1 contains corresponding protein accession numbers and mass spec data. Supplemental Data File S3 contains additional gene information
      Rat symbolRatio ES/RSProtein nameFunctionSpermatid localization
      Reported relative abundance observed histologically or by western blot during spermiogenesis.
      Ref.
      c.r., current report.
      RGD1306195 rg32.3Similar to RANUnknownUnknownc.r.
      Fam216a
      Confidence Score <100.
      19.21500011H22Rik-LikeUnknownUnknownc.r.
      Dnaja418.6DnaJ (Hsp40) homolog, subfamily A, member 4Protein StabilityUnknownc.r.
      Pgk2 rg16.6Phosphoglycerate kinase 2GlycolysisES
      • McCarrey J.R.
      • Berg W.M.
      • Paragioudakis S.J.
      • Zhang P.L.
      • Dilworth D.D.
      • Arnold B.L.
      • Rossi J.J.
      Differential transcription of Pgk genes during spermatogenesis in the mouse.
      Hspa1l rg
      Proteins encoded by 3 highly identical heat shock protein 70 retrogenes (HSPA1a, HSPA1L, HSPA1b) identified in same protein spot.
      14.8Heat shock 70 protein 1-LikeProtein StabilityES
      • Tsunekawa N.
      • Matsumoto M.
      • Tone S.
      • Nishida T.
      • Fujimoto H.
      The Hsp70 homolog gene, Hsc70t, is expressed under translational control during mouse spermiogenesis.
      Oxct2a rg12.5Succinyl-CoA:3-ketoacid-coenzyme A transferase 2AMetabolismES
      • Koga M.
      • Tanaka H.
      • Yomogida K.
      • Nozaki M.
      • Tsuchida J.
      • Ohta H.
      • Nakamura Y.
      • Masai K.
      • Yoshimura Y.
      • Yamanaka M.
      • Iguchi N.
      • Nojima H.
      • Matsumiya K.
      • Okuyama A.
      • Nishimune Y.
      Isolation and characterization of a haploid germ cell-specific novel complementary deoxyribonucleic acid; testis-specific homologue of succinyl CoA:3-Oxo acid CoA transferase.
      Aldoart2 rg12.3Aldolase 1 A Retrogene 2GlycolysisES
      • Vemuganti S.A.
      • Bell T.A.
      • Scarlett C.O.
      • Parker C.E.
      • de Villena F.P.
      • O'Brien D.A.
      Three male germline-specific aldolase A isozymes are generated by alternative splicing and retrotransposition.
      Aldoa12.2Aldolase AGlycolysisES
      • Vemuganti S.A.
      • Bell T.A.
      • Scarlett C.O.
      • Parker C.E.
      • de Villena F.P.
      • O'Brien D.A.
      Three male germline-specific aldolase A isozymes are generated by alternative splicing and retrotransposition.
      Spata2011.8Spermatogenesis associated 20Antioxidant-LikeES
      • Shi H.J.
      • Wu A.Z.
      • Santos M.
      • Feng Z.M.
      • Huang L.
      • Chen Y.M.
      • Zhu K.
      • Chen C.L.
      Cloning and characterization of rat spermatid protein SSP411: a thioredoxin-like protein.
      Eef211.6Elongation factor 2Protein SynthesisESc.r.
      EGFP10.9Enhanced Green Fluorescent ProteinTransgeneESc.r.
      Prdx49.8peroxiredoxin-4AnitoxidantES
      • Sasagawa I.
      • Matsuki S.
      • Suzuki Y.
      • Iuchi Y.
      • Tohya K.
      • Kimura M.
      • Nakada T.
      • Fujii J.
      Possible involvement of the membrane-bound form of peroxiredoxin 4 in acrosome formation during spermiogenesis of rats.
      Cap28.6Adenylyl cyclase-associated protein 2CytoskeletonUnknown
      • Swiston J.
      • Hubberstey A.
      • Yu G.
      • Young D.
      Differential expression of CAP and CAP2 in adult rat tissues.
      Pgam26.8Phosphoglycerate mutase 2GlycolysisES
      • Fundele R.
      • Winking H.
      • Illmensee K.
      • Jagerbauer E.M.
      Developmental activation of phosphoglycerate mutase-2 in the testis of the mouse.
      Wdr16.8WD repeat domain 1Actin BindingUnknownc.r.
      Glul6.7Glutamate-ammonia ligaseMetabolismES
      • Lie-Venema H.
      • de Boer P.A.
      • Moorman A.F.
      • Lamers W.H.
      Role of the 5′ enhancer of the glutamine synthetase gene in its organ-specific expression.
      Tpi16.2Triosephosphate isomerase 1GlycolysisUnknown
      • Russell D.L.
      • Kim K.H.
      Expression of triosephosphate isomerase transcripts in rat testis: evidence for retinol regulation and a novel germ cell transcript.
      Aif16.2Allograft inflammatory factor 1InflammationES
      • Iida H.
      • Doiguchi M.
      • Yamashita H.
      • Sugimachi S.
      • Ichinose J.
      • Mori T.
      • Shibata Y.
      Spermatid-specific expression of Iba1, an ionized calcium binding adapter molecule-1, in rat testis.
      Asrgl16.1Asparaginase-like 1Protein ModificationUnknown
      • Bush L.A.
      • Herr J.C.
      • Wolkowicz M.
      • Sherman N.E.
      • Shore A.
      • Flickinger C.J.
      A novel asparaginase-like protein is a sperm autoantigen in rats.
      Gstm55.5Glutathione S-transferase Mu 5AntioxidantES
      • Nakamura N.
      • Mori C.
      • Eddy E.M.
      Molecular complex of three testis-specific isozymes associated with the mouse sperm fibrous sheath: hexokinase 1, phosphofructokinase M, and glutathione S-transferase mu class 5.
      Phyhipl4.7Phytanoyl-CoA 2-hydroxylase interacting protein-likePhytanic Acid MetabolismUnknownc.r.
      Pdilt4.5Protein disulfide isomerase-like, testis expressedProtein FoldingES
      • van Lith M.
      • Karala A.R.
      • Bown D.
      • Gatehouse J.A.
      • Ruddock L.W.
      • Saunders P.T.
      • Benham A.M.
      A developmentally regulated chaperone complex for the endoplasmic reticulum of male haploid germ cells.
      Eno1 p-rg3.9Enolase 1GlycolysisES
      • Gitlits V.M.
      • Toh B.H.
      • Loveland K.L.
      • Sentry J.W.
      The glycolytic enzyme enolase is present in sperm tail and displays nucleotide-dependent association with microtubules.
      Eef1g3.7Eukaryotic translation elongation factor 1 gammaProtein SynthesisUnknown
      • Maywood E.S.
      • Chahad-Ehlers S.
      • Garabette M.L.
      • Pritchard C.
      • Underhill P.
      • Greenfield A.
      • Ebling F.J.
      • Akhtar R.A.
      • Kyriacou C.P.
      • Hastings M.H.
      • Reddy A.B.
      Differential testicular gene expression in seasonal fertility.
      Hspa4l3.4Heat shock 70 kDa protein 4LProtein FoldingES
      • Held T.
      • Paprotta I.
      • Khulan J.
      • Hemmerlein B.
      • Binder L.
      • Wolf S.
      • Schubert S.
      • Meinhardt A.
      • Engel W.
      • Adham I.M.
      Hspa4l-deficient mice display increased incidence of male infertility and hydronephrosis development.
      Pebp13.2Phosphatidylethanolamine binding protein 1Kinase InhibitorUnknown
      • Moffit J.S.
      • Boekelheide K.
      • Sedivy J.M.
      • Klysik J.
      Mice lacking Raf kinase inhibitor protein-1 (RKIP-1) have altered sperm capacitation and reduced reproduction rates with a normal response to testicular injury.
      Thop13.2Thimet oligopeptidase 1ProteolysisUnknownc.r.
      Pcmt13.0Protein-l-isoaspartate O-methyltransferase 1Protein ModificationUnknownc.r.
      a Reported relative abundance observed histologically or by western blot during spermiogenesis.
      b c.r., current report.
      c Confidence Score <100.
      d Proteins encoded by 3 highly identical heat shock protein 70 retrogenes (HSPA1a, HSPA1L, HSPA1b) identified in same protein spot.
      Figure thumbnail gr8
      Fig. 8Mass spectrometry predicts protein localization in rat spermatids. A, Localization of anti-hnRNPK IgG binding (red) in wild-type rat testis. Note: labeling in pachytene spermatocytes (PS) and round spermatids (RS), but not in elongating spermatids (ES). Hoechst 33342 dye labels nuclei (blue) of all testis cells. Scale bar, 30 μm. B, Localization of anti-STMN1 IgG binding (red) in GESptd1 rat testis. Note: labeling in pachytene spermatocytes (PS) and round spermatids (RS), but not in elongating spermatids (ES). Hoechst 33342 dye labels nuclei (blue) of all testis cells. Asterisks marks stage IX seminiferous tubule containing step 9 spermatids starting to elongate. Scale bar, 100 μm. C, Localization of anti-GLUL IgG binding (red) in rat GESptd1 testis. Note: labeling in elongating spermatids (ES), but not in pachytene spermatocytes (PS) or round spermatids (RS). Hoechst 33342 dye labels nuclei (blue) of all testis cells. Scale bar, 100 μm.
      Figure thumbnail gr9
      Fig. 9Testicular expression of PCBP1 and EEF2 correlate with mass spectrometry signals. A, Localization of anti-PCBP1 IgG binding (red) in wild-type rat testis. Note: widespread labeling in pachytene spermatocytes (PS) and round spermatids (RS), but predominantly focal localization to granules in elongating spermatids (ES). Hoechst 33342 dye labels nuclei (blue) of all testis cells. Roman numbers estimate spermatogenic stage. Scale bars, 30 μm. B, Localization of Anit-EEF2 IgG Binding in wild-type rat Testis. Round Spermatid (RS); Elongating Spermatid (ES); Pachytene Spermatocyte (Py); Preleptotene Spermatocyte (PL); Intermediate Spermatogonia (Int Spg); Sertoli Cell (SC). Roman numbers estimate spermatogenic stage. Scale bar, 30 μm.
      Ontology analysis of all sequenced proteins revealed the round spermatid fraction to be dominated by proteins that regulate RNA binding, processing and transport (supplemental Data File S4). This cluster included hnRNPA2/b1, hnRNPA3, hnRPDL, hnRNPK, hnRNPL, hnRNPM, PABPC1, PABPC4, PCBP1, PCBP3, PTBP2, PSIP1, RGSL1, RUVBL2, SARNP2, TDRD6, and TDRD7 (Table I). Interestingly, with respect to post-transcriptional phases of spermatozoan biology, fractions containing elongating spermatids were dominated by alternative forms of glycolytic/metabolic enzymes, redox enzymes, protein synthesis enzymes, and factors that mediate protein stability (Table II).
      Retrogene products were also over-represented within the elongating spermatid proteins sequenced; five of the top seven proteins in Table II are predicted to be encoded by retrogenes.

      DISCUSSION

      Here, a proteomics approach using flow sorted fractions of round and elongating spermatids revealed global down-regulation of an abundant RNA binding proteome expressed prior to up-regulation of metabolic factors during spermatid elongation. Based on biochemical activities in other biological processes, the identified factors are predicted to drive post-transcriptional RNA processing during meiotic and post-meiotic steps in spermatozoan development. Genome-wide studies find testes to express an extraordinarily diverse repertoire of tissue-specific mRNAs (
      • Eddy E.M.
      Male germ cell gene expression.
      ,
      • Schultz N.
      • Hamra F.K.
      • Garbers D.L.
      A multitude of genes expressed solely in meiotic or postmeiotic spermatogenic cells offers a myriad of contraceptive targets.
      ,
      • Johnston D.S.
      • Wright W.W.
      • Dicandeloro P.
      • Wilson E.
      • Kopf G.S.
      • Jelinsky S.A.
      Stage-specific gene expression is a fundamental characteristic of rat spermatogenic cells and Sertoli cells.
      ), alternatively spliced mRNAs (
      • He C
      • Zuo Z.
      • Chen H.
      • Zhang L.
      • Zhou F.
      • Cheng H.
      • Zhou R.
      Genome-wide detection of testis- and testicular cancer-specific alternative splicing.
      ,
      • Harr B.
      • Turner L.M.
      Genome-wide analysis of alternative splicing evolution among Mus subspecies.
      ), translationally repressed mRNAs (
      • Schmidt E.E.
      • Hanson E.S.
      • Capecchi M.R.
      Sequence-independent assembly of spermatid mRNAs into messenger ribonucleoprotein particles.
      ,
      • Kleene K.C.
      Patterns of translational regulation in the mammalian testis.
      ,
      • Yang J.
      • Medvedev S.
      • Reddi P.P.
      • Schultz R.M.
      • Hecht N.B.
      The DNA/RNA-binding protein MSY2 marks specific transcripts for cytoplasmic storage in mouse male germ cells.
      ), and translated retrogene mRNAs (
      • Vemuganti S.A.
      • de Villena F.P.
      • O'Brien D.A.
      Frequent and recent retrotransposition of orthologous genes plays a role in the evolution of sperm glycolytic enzymes.
      ). Testes not only express an unusually wide array of total exons compared with most tissues, but also express exons encoding tissue-specific, cis-acting splicing elements; a majority of these elements bind hnRNP-family proteins that mediate exon skipping (
      • Wang X.
      • Wang K.
      • Radovich M.
      • Wang Y.
      • Wang G.
      • Feng W.
      • Sanford J.R.
      • Liu Y.
      Genome-wide prediction of cis-acting RNA elements regulating tissue-specific pre-mRNA alternative splicing.
      ). Accordingly, testes express a highly diverse collection of transcripts encoding proteins that regulate alternative splicing (
      • Grosso A.R.
      • Gomes A.Q.
      • Barbosa-Morais N.L.
      • Caldeira S.
      • Thorne N.P.
      • Grech G.
      • von Lindern M.
      • Carmo-Fonseca M.
      Tissue-specific splicing factor gene expression signatures.
      ). This conduit of post-transcriptional genetic diversity unleashed by the testis is directly related to its fundamental role in supporting sex-specific processes required for spermatozoan development and fertility (
      • Eddy E.M.
      Male germ cell gene expression.
      ,
      • Schultz N.
      • Hamra F.K.
      • Garbers D.L.
      A multitude of genes expressed solely in meiotic or postmeiotic spermatogenic cells offers a myriad of contraceptive targets.
      ,
      • Johnston D.S.
      • Wright W.W.
      • Dicandeloro P.
      • Wilson E.
      • Kopf G.S.
      • Jelinsky S.A.
      Stage-specific gene expression is a fundamental characteristic of rat spermatogenic cells and Sertoli cells.
      ). Thus, it seems logical that we identified RNA regulatory factors as a major, differentially expressed class of proteins in fractions of round spermatids.
      Heterogeneous RNA-binding proteins that mediate pre-mRNA splicing, including: hnRNPA2b1 (
      • Matsui M.
      • Horiguchi H.
      • Kamma H.
      • Fujiwara M.
      • Ohtsubo R.
      • Ogata T.
      Testis- and developmental stage-specific expression of hnRNP A2/B1 splicing isoforms, B0a/b.
      ,
      • Clower C.V.
      • Chatterjee D.
      • Wang Z.
      • Cantley L.C.
      • Vander Heiden M.G.
      • Krainer A.R.
      The alternative splicing repressors hnRNP A1/A2 and PTB influence pyruvate kinase isoform expression and cell metabolism.
      ), hnRPDL (
      • Zubovic L.
      • Baralle M.
      • Baralle F.E.
      Mutually exclusive splicing regulates the Nav 1.6 sodium channel function through a combinatorial mechanism that involves three distinct splicing regulatory elements and their ligands.
      ), hnRNPK (
      • Expert-Bezancon A.
      • Le Caer J.P.
      • Marie J.
      Heterogeneous nuclear ribonucleoprotein (hnRNP) K is a component of an intronic splicing enhancer complex that activates the splicing of the alternative exon 6A from chicken beta-tropomyosin pre-mRNA.
      ), hnRNPM (
      • Hovhannisyan R.H.
      • Carstens R.P.
      Heterogeneous ribonucleoprotein m is a splicing regulatory protein that can enhance or silence splicing of alternatively spliced exons.
      ), PCBP1 (
      • Meng Q.
      • Rayala S.K.
      • Gururaj A.E.
      • Talukder A.H.
      • O'Malley B.W.
      • Kumar R.
      Signaling-dependent and coordinated regulation of transcription, splicing, and translation resides in a single coregulator, PCBP1.
      ), PCBP3 (
      • Wang Y.
      • Gao L.
      • Tse S.W.
      • Andreadis A.
      Heterogeneous nuclear ribonucleoprotein E3 modestly activates splicing of tau exon 10 via its proximal downstream intron, a hotspot for frontotemporal dementia mutations.
      ), PSIP1 (
      • Pradeepa M.M.
      • Sutherland H.G.
      • Ule J.
      • Grimes G.R.
      • Bickmore W.A.
      Psip1/Ledgf p52 binds methylated histone H3K36 and splicing factors and contributes to the regulation of alternative splicing.
      ,
      • Ge H.
      • Si Y.
      • Wolffe A.P.
      A novel transcriptional coactivator, p52, functionally interacts with the essential splicing factor ASF/SF2.
      ), and PTBP2 (
      • Boutz P.L.
      • Stoilov P.
      • Li Q.
      • Lin C.H.
      • Chawla G.
      • Ostrow K.
      • Shiue L.
      • Ares Jr., M.
      • Black D.L.
      A post-transcriptional regulatory switch in polypyrimidine tract-binding proteins reprograms alternative splicing in developing neurons.
      ) comprised the most notable protein cluster we indentified in round spermatids. Several of these same proteins (i.e. hnRNPA2b/1, hnRNPK, hnRNPL, PCBP1, PTBP2) (
      • Hamilton B.J.
      • Nichols R.C.
      • Tsukamoto H.
      • Boado R.J.
      • Pardridge W.M.
      • Rigby W.F.
      hnRNP A2 and hnRNP L bind the 3′UTR of glucose transporter 1 mRNA and exist as a complex in vivo.
      ,
      • Majumder M.
      • Yaman I.
      • Gaccioli F.
      • Zeenko V.V.
      • Wang C.
      • Caprara M.G.
      • Venema R.C.
      • Komar A.A.
      • Snider M.D.
      • Hatzoglou M.
      The hnRNA-binding proteins hnRNP L and PTB are required for efficient translation of the Cat-1 arginine/lysine transporter mRNA during amino acid starvation.
      ,
      • Zheng S.
      • Gray E.E.
      • Chawla G.
      • Porse B.T.
      • O'Dell T.J.
      • Black D.L.
      PSD-95 is post-transcriptionally repressed during early neural development by PTBP1 and PTBP2.
      ,
      • Ostareck D.H.
      • Ostareck-Lederer A.
      • Shatsky I.N.
      • Hentze M.W.
      Lipoxygenase mRNA silencing in erythroid differentiation: The 3′UTR regulatory complex controls 60S ribosomal subunit joining.
      ), plus PABPC1 (
      • Yanagiya A.
      • Delbes G.
      • Svitkin Y.V.
      • Robaire B.
      • Sonenberg N.
      The poly(A)-binding protein partner Paip2a controls translation during late spermiogenesis in mice.
      ) and two Tudor-family proteins, TDRD6 and TDRD7 (
      • Hosokawa M.
      • Shoji M.
      • Kitamura K.
      • Tanaka T.
      • Noce T.
      • Chuma S.
      • Nakatsuji N.
      Tudor-related proteins TDRD1/MTR-1, TDRD6 and TDRD7/TRAP: domain composition, intracellular localization, and function in male germ cells in mice.
      ,
      • Tanaka T.
      • Hosokawa M.
      • Vagin V.V.
      • Reuter M.
      • Hayashi E.
      • Mochizuki A.L.
      • Kitamura K.
      • Yamanaka H.
      • Kondoh G.
      • Okawa K.
      • Kuramochi-Miyagawa S.
      • Nakano T.
      • Sachidanandam R.
      • Hannon G.J.
      • Pillai R.S.
      • Nakatsuji N.
      • Chuma S.
      Tudor domain containing 7 (Tdrd7) is essential for dynamic ribonucleoprotein (RNP) remodeling of chromatoid bodies during spermatogenesis.
      ,
      • Vasileva A.
      • Tiedau D.
      • Firooznia A.
      • Muller-Reichert T.
      • Jessberger. R
      Tdrd6 is required for spermiogenesis, chromatoid body architecture, and regulation of miRNA expression.
      ,
      • Lachke S.A.
      • Alkuraya F.S.
      • Kneeland S.C.
      • Ohn T.
      • Aboukhalil A.
      • Howell G.R.
      • Saadi I.
      • Cavallesco R.
      • Yue Y.
      • Tsai A.C.
      • Nair K.S.
      • Cosma M.I.
      • Smith R.S.
      • Hodges E.
      • Alfadhli S.M.
      • Al-Hajeri A.
      • Shamseldin H.E.
      • Behbehani A.
      • Hannon G.J.
      • Bulyk M.L.
      • Drack A.V.
      • Anderson P.J.
      • John S.W.
      • Maas R.L.
      Mutations in the RNA granule component TDRD7 cause cataract and glaucoma.
      ) are established mediators of translational regulation. Similarly, hnRNPA2b1, hnRNPI, hnRNPL, PTBP2, TDRD7, and RUVBL2 prevent premature translation, but do so by regulating mRNA stability (
      • Fahling M.
      • Mrowka R.
      • Steege A.
      • Martinka P.
      • Persson P.B.
      • Thiele B.J.
      Heterogeneous nuclear ribonucleoprotein-A2/B1 modulate collagen prolyl 4-hydroxylase, alpha (I) mRNA stability.
      ,
      • Hui J.
      • Reither G.
      • Bindereif A.
      Novel functional role of CA repeats and hnRNP L in RNA stability.
      ,
      • Soderberg M.
      • Raffalli-Mathieu F.
      • Lang M.A.
      Inflammation modulates the interaction of heterogeneous nuclear ribonucleoprotein (hnRNP) I/polypyrimidine tract binding protein and hnRNP L with the 3′untranslated region of the murine inducible nitric-oxide synthase mRNA.
      ). In fact, In vitro studies demonstrated PTBP2 to reduce translation of the testis specific PGK2 by stimulating its decay (
      • Xu M.
      • McCarrey J.R.
      • Hecht N.B.
      A cytoplasmic variant of the KH-type splicing regulatory protein serves as a decay-promoting factor for phosphoglycerate kinase 2 mRNA in murine male germ cells.
      ). In somatic tissues, hnRNPA3 (
      • Ma A.S.
      • Moran-Jones K.
      • Shan J.
      • Munro T.P.
      • Snee M.J.
      • Hoek K.S.
      • Smith R.
      Heterogeneous nuclear ribonucleoprotein A3, a novel RNA trafficking response element-binding protein.
      ), hnRNPA2b1 (
      • Raju C.S.
      • Goritz C.
      • Nord Y.
      • Hermanson O.
      • Lopez-Iglesias C.
      • Visa N.
      • Castelo-Branco G.
      • Percipalle P.
      In cultured oligodendrocytes the A/B-type hnRNP CBF-A accompanies MBP mRNA bound to mRNA trafficking sequences.
      ), and SNARP/CIP29 (
      • Dufu K.
      • Livingstone M.J.
      • Seebacher J.
      • Gygi S.P.
      • Wilson S.A.
      • Reed R.
      ATP is required for interactions between UAP56 and two conserved mRNA export proteins, Aly CIP29, to assemble the TREX complex.
      ) mediate RNA transport and packaging between subcellular compartments, particularly export from the nucleus. Thus, collective expression and then loss of this protein cluster in spermatids prior to up-regulating glycolytic and protein biosynthetic machinery during spermatid elongation sheds light on how the haploid genome ultimately produces, and then translates such a diverse transcriptome.
      With respect to nuclear localization of pre-mRNA splicing factors, a strong positive correlation exists between several nuclear RNA binding proteins identified in the flow-sorted round spermatid fraction and their histological localization in testis sections (
      • Matsui M.
      • Horiguchi H.
      • Kamma H.
      • Fujiwara M.
      • Ohtsubo R.
      • Ogata T.
      Testis- and developmental stage-specific expression of hnRNP A2/B1 splicing isoforms, B0a/b.
      ,
      • Kamma H.
      • Horiguchi H.
      • Wan L.
      • Matsui M.
      • Fujiwara M.
      • Fujimoto M.
      • Yazawa T.
      • Dreyfuss G.
      Molecular characterization of the hnRNP A2/B1 proteins: tissue-specific expression and novel isoforms.
      ,
      • Xu M.
      • Hecht N.B.
      Polypyrimidine tract binding protein 2 stabilizes phosphoglycerate kinase 2 mRNA in murine male germ cells by binding to its 3′UTR.
      ,
      • Hosokawa M.
      • Shoji M.
      • Kitamura K.
      • Tanaka T.
      • Noce T.
      • Chuma S.
      • Nakatsuji N.
      Tudor-related proteins TDRD1/MTR-1, TDRD6 and TDRD7/TRAP: domain composition, intracellular localization, and function in male germ cells in mice.
      ,
      • Tanaka T.
      • Hosokawa M.
      • Vagin V.V.
      • Reuter M.
      • Hayashi E.
      • Mochizuki A.L.
      • Kitamura K.
      • Yamanaka H.
      • Kondoh G.
      • Okawa K.
      • Kuramochi-Miyagawa S.
      • Nakano T.
      • Sachidanandam R.
      • Hannon G.J.
      • Pillai R.S.
      • Nakatsuji N.
      • Chuma S.
      Tudor domain containing 7 (Tdrd7) is essential for dynamic ribonucleoprotein (RNP) remodeling of chromatoid bodies during spermatogenesis.
      ,
      • Seidah N.G.
      • Day R.
      • Hamelin J.
      • Gaspar A.
      • Collard M.W.
      • Chretien M.
      Testicular expression of PC4 in the rat: molecular diversity of a novel germ cell-specific Kex2/subtilisin-like proprotein convertase.
      ,
      • Vasileva A.
      • Tiedau D.
      • Firooznia A.
      • Muller-Reichert T.
      • Jessberger. R
      Tdrd6 is required for spermiogenesis, chromatoid body architecture, and regulation of miRNA expression.
      ,
      • Yu Y.
      • Zhao C.
      • Lv Z.
      • Chen W.
      • Tong M.
      • Guo X.
      • Wang L.
      • Liu J.
      • Zhou Z.
      • Zhu H.
      • Zhou Q.
      • Sha J.
      Microinjection manipulation resulted in the increased apoptosis of spermatocytes in testes from intracytoplasmic sperm injection (ICSI) derived mice.
      ,
      • Xu M.
      • McCarrey J.R.
      • Hecht N.B.
      A cytoplasmic variant of the KH-type splicing regulatory protein serves as a decay-promoting factor for phosphoglycerate kinase 2 mRNA in murine male germ cells.
      ). Experimentally, we obtained 100% agreement between our mass-spectrometry results and actual localization of spermatid proteins identified in 26 independent studies previously conducted in rats and mice (Table I, Table II). Moreover, 17 of the 26 validating reports were conducted using mice. This demonstrates clear conservation of processes that control protein expression during spermiogenesis in these rodents.
      However, it is interesting that only ∼35% of spermatids in the elongating spermatid fraction retained their nucleus following the isolation procedure. Though biochemically suitable for comparing proteins sorted as different cellular and sub-cellular fractions; theoretically, this should bias our results by ∼threefold when measuring the relative abundance of nuclear proteins down-regulated during spermiogenesis. Thus, the clear absence of false-positive discoveries following histological validation of >50% of the total identified proteins in testis sections was intriguing. This fact potentially underscores dramatic remodeling of nuclear protein composition that normally occurs during late steps in spermiogenesis (
      • Yu Y.E.
      • Zhang Y.
      • Unni E.
      • Shirley C.R.
      • Deng J.M.
      • Russell L.D.
      • Weil M.M.
      • Behringer R.R.
      • Meistrich M.L.
      Abnormal spermatogenesis and reduced fertility in transition nuclear protein 1-deficient mice.
      ,
      • Oliva R.
      • Dixon G.H.
      Vertebrate protamine genes and the histone-to-protamine replacement reaction. Prog Nucleic Acids Res.
      ,
      • Meistrich M.L.
      • Trostle-Weige P.K.
      • Lin R.
      • Bhatnagar Y.M.
      • Allis C.D.
      Highly acetylated H4 is associated with histone displacement in rat spermatids.
      ,
      • Meistrich M.L.
      • Mohapatra B.
      • Shirley C.R.
      • Zhao M.
      Roles of transition nuclear proteins in spermiogenesis.
      ). Consequently, normal depletion of representative round spermatid nuclear proteins from fully compacted and elongated spermatids mirrors the relative abundance of these proteins detected in our collected round and elongating spermatid fractions. As such, this current approach would also be less sensitive for identifying up-regulated elongating spermatid proteins tightly association with more mature flagella and nuclei (
      • Eddy E.M.
      • Toshimori K.
      • O'Brien D.A.
      Fibrous sheath of mammalian spermatozoa.
      ). Optimization of methods for testis cell disaggregation and sorting could result in higher yields of intact elongating spermatids. Additional purification steps could also better resolve intact sorted elongating spermatids from cytoplasts/residual bodies (
      • Bellve A.R.
      • Millette C.F.
      • Bhatnagar Y.M.
      • O'Brien D.A.
      Dissociation of the mouse testis and characterization of isolated spermatogenic cells.
      ). At any rate, additional studies are required to validate localization of ∼48% remaining spermiogenic factors in Table I, Table II.
      We also demonstrated that post-transcriptional regulation of EGFP expression in GESptd1 transgenic rats mimics genes up-regulated during later steps of spermiogenesis. This is because, like many spermiogenic genes, transcription and translation of the EGFP reporter appears to be largely uncoupled in GESptd1 rats. This developmental paradigm is highly conserved in sexually reproducing organisms (
      • Henderson M.A.
      • Cronland E.
      • Dunkelbarger S.
      • Contreras V.
      • Strome S.
      • Keiper B.D.
      A germline-specific isoform of eIF4E (IFE-1) is required for efficient translation of stored mRNAs and maturation of both oocytes and sperm.
      ,
      • Kleene K.C.
      Patterns of translational regulation in the mammalian testis.
      ,
      • Wolniak S.M.
      • van der Weele C.M.
      • Deeb F.
      • Boothby T.
      • Klink V.P.
      Extremes in rapid cellular morphogenesis: post-transcriptional regulation of spermatogenesis in Marsilea vestita.
      ), and is well established for a large fraction of total transcripts (∼75%) that are stored in the nonpolysomal fractions of rodent spermatocytes and round spermatids prior to elongation (
      • Gold B.
      • Hecht N.B.
      Differential compartmentalization of messenger ribonucleic acid in murine testis.
      ,
      • Gold B.
      • Stern L.
      • Bradley F.M.
      • Hecht N.B.
      Gene expression during mammalian spermatogenesis. II. Evidence for stage-specific differences in mRNA populations.
      ). Requirement for this protracted translational delay for many sperm-specific genes is highlighted by early ectopic expression of protamine 1 or transition protein 2 in round spermatids (
      • Tseden K.
      • Topaloglu O.
      • Meinhardt A.
      • Dev A.
      • Adham I.
      • Muller C.
      • Wolf S.
      • Bohm D.
      • Schluter G.
      • Engel W.
      • Nayernia K.
      Premature translation of transition protein 2 mRNA causes sperm abnormalities and male infertility.
      ,
      • Lee K.
      • Haugen H.S.
      • Clegg C.H.
      • Braun R.E.
      Premature translation of protamine 1 mRNA causes precocious nuclear condensation and arrests spermatid differentiation in mice.
      ). Forced premature expression of these sperm histone proteins in round spermatids stimulates precocious chromatin condensation prior to spermatid elongation, and results in infertility due to a block in spermatid development or function (
      • Tseden K.
      • Topaloglu O.
      • Meinhardt A.
      • Dev A.
      • Adham I.
      • Muller C.
      • Wolf S.
      • Bohm D.
      • Schluter G.
      • Engel W.
      • Nayernia K.
      Premature translation of transition protein 2 mRNA causes sperm abnormalities and male infertility.
      ,
      • Lee K.
      • Haugen H.S.
      • Clegg C.H.
      • Braun R.E.
      Premature translation of protamine 1 mRNA causes precocious nuclear condensation and arrests spermatid differentiation in mice.
      ). Thus, based on functional roles in other cell types (
      • Ma A.S.
      • Moran-Jones K.
      • Shan J.
      • Munro T.P.
      • Snee M.J.
      • Hoek K.S.
      • Smith R.
      Heterogeneous nuclear ribonucleoprotein A3, a novel RNA trafficking response element-binding protein.
      ,
      • Raju C.S.
      • Goritz C.
      • Nord Y.
      • Hermanson O.
      • Lopez-Iglesias C.
      • Visa N.
      • Castelo-Branco G.
      • Percipalle P.
      In cultured oligodendrocytes the A/B-type hnRNP CBF-A accompanies MBP mRNA bound to mRNA trafficking sequences.
      ,
      • Dufu K.
      • Livingstone M.J.
      • Seebacher J.
      • Gygi S.P.
      • Wilson S.A.
      • Reed R.
      ATP is required for interactions between UAP56 and two conserved mRNA export proteins, Aly CIP29, to assemble the TREX complex.
      ,
      • Yu Y.E.
      • Zhang Y.
      • Unni E.
      • Shirley C.R.
      • Deng J.M.
      • Russell L.D.
      • Weil M.M.
      • Behringer R.R.
      • Meistrich M.L.
      Abnormal spermatogenesis and reduced fertility in transition nuclear protein 1-deficient mice.
      ), RNA binding proteins identified in this study, such as hnRNPA2b/1, hnRNPK, hnRNPL, PCBP1, and/or PTBP2 potentially represent factors that mediate translational repression of mRNAs in spermatocytes and round spermatids (Fig. 10). This theory further implicates that some unknown signal generated in the seminiferous epithelium inactivates and/or stimulates degradation of such RNA binding factors to facilitate spermatid elongation (Fig. 10).
      Figure thumbnail gr10
      Fig. 10Post-transcriptional regulation of rat tgGESptd1 expression. Open questions are illustrated on how expression of the GESptd1 rat transgene (tgGESptd1) is regulated post-transcriptionally during spermatid maturation. Here, translation of a tgGESptd1 encoded mRNA upon spermatid elongation is modeled hypothetically in association with hnRNP-family proteins (, ), reported small RNA processing pathways (
      • Korhonen H.M.
      • Meikar O.
      • Yadav R.P.
      • Papaioannou M.D.
      • Romero Y.
      • Da Ros M.
      • Herrera P.L.
      • Toppari J.
      • Nef S.
      • Kotaja N.
      Dicer is required for haploid male germ cell differentiation in mice.
      ), poly-A tail de-adenylation (
      • Kleene K.C.
      Poly(A) shortening accompanies the activation of translation of five mRNAs during spermiogenesis in the mouse.
      ,
      • Kleene K.C.
      • Distel R.J.
      • Hecht N.B.
      Translational regulation and deadenylation of a protamine mRNA during spermiogenesis in the mouse.
      ) and incorporation of processed transcripts into translational complexes (
      • Gold B.
      • Hecht N.B.
      Differential compartmentalization of messenger ribonucleic acid in murine testis.
      ,
      • Gold B.
      • Stern L.
      • Bradley F.M.
      • Hecht N.B.
      Gene expression during mammalian spermatogenesis. II. Evidence for stage-specific differences in mRNA populations.
      ). Mechanisms controlling relative abundance of hnRNPs upon spermatid elongation remain to be determined. Also illustrated, based on information from other studies, are hypothetical GW182/AGO-like protein components in RNA silencing complex (RISC) (
      • Korhonen H.M.
      • Meikar O.
      • Yadav R.P.
      • Papaioannou M.D.
      • Romero Y.
      • Da Ros M.
      • Herrera P.L.
      • Toppari J.
      • Nef S.
      • Kotaja N.
      Dicer is required for haploid male germ cell differentiation in mice.
      ), associated poly-A tail, de-adenylation complex (DAC) (
      • Kleene K.C.
      Poly(A) shortening accompanies the activation of translation of five mRNAs during spermiogenesis in the mouse.
      ,
      • Kleene K.C.
      • Distel R.J.
      • Hecht N.B.
      Translational regulation and deadenylation of a protamine mRNA during spermiogenesis in the mouse.
      ), 40S and 60S ribosomal subunits, 5′ and 3′ untranslated regions (UTR), eIF4A, eIF4E, eIF4G and eIF3 elongation initiation factors, poly-A-binding protein C1 (PABPC1), poly(A)-tail binding protein-interacting protein 2 (PAIP2) (
      • Yanagiya A.
      • Delbes G.
      • Svitkin Y.V.
      • Robaire B.
      • Sonenberg N.
      The poly(A)-binding protein partner Paip2a controls translation during late spermiogenesis in mice.
      ,
      • Delbes G.
      • Yanagiya A.
      • Sonenberg N.
      • Robaire B.
      PABP interacting protein 2A (PAIP2A) regulates specific key proteins during spermiogenesis in the mouse.
      ), and additional putative translational enhancers (TE) (
      • Yanagiya A.
      • Delbes G.
      • Svitkin Y.V.
      • Robaire B.
      • Sonenberg N.
      The poly(A)-binding protein partner Paip2a controls translation during late spermiogenesis in mice.
      ,
      • Delbes G.
      • Yanagiya A.
      • Sonenberg N.
      • Robaire B.
      PABP interacting protein 2A (PAIP2A) regulates specific key proteins during spermiogenesis in the mouse.
      ).
      Most recently, biochemical insight into mechanisms that trigger translational activation and de-repression of stored mRNAs was reported (
      • Yanagiya A.
      • Delbes G.
      • Svitkin Y.V.
      • Robaire B.
      • Sonenberg N.
      The poly(A)-binding protein partner Paip2a controls translation during late spermiogenesis in mice.
      ,
      • Delbes G.
      • Yanagiya A.
      • Sonenberg N.
      • Robaire B.
      PABP interacting protein 2A (PAIP2A) regulates specific key proteins during spermiogenesis in the mouse.
      ,
      • Chang Y.F.
      • Lee-Chang J.S.
      • Imam J.S.
      • Buddavarapu K.C.
      • Subaran S.S.
      • Sinha-Hikim A.P.
      • Gorospe M.
      • Rao M.K.
      Interaction between microRNAs and actin-associated protein Arpc5 regulates translational suppression during male germ cell differentiation.
      ). Yanagiya and colleagues demonstrated that a double knockout of the poly-A-binding protein-interacting proteins (i.e. PAIP2a & PAIP2b) resulted in pre-mature translation of stored mRNAs and a lack of spermatid elongation (
      • Yanagiya A.
      • Delbes G.
      • Svitkin Y.V.
      • Robaire B.
      • Sonenberg N.
      The poly(A)-binding protein partner Paip2a controls translation during late spermiogenesis in mice.
      ). It was demonstrated that up-regulation of these proteins in round spermatids just prior to elongation under normal conditions proved instrumental in titrating out repressive effects of PABPC1 on translation (
      • Yanagiya A.
      • Delbes G.
      • Svitkin Y.V.
      • Robaire B.
      • Sonenberg N.
      The poly(A)-binding protein partner Paip2a controls translation during late spermiogenesis in mice.
      ). And more recently, Delbes and colleagues showed PAIP2A to regulate translation of a specific subset of spermatid mRNAs by this pathway (
      • Delbes G.
      • Yanagiya A.
      • Sonenberg N.
      • Robaire B.
      PABP interacting protein 2A (PAIP2A) regulates specific key proteins during spermiogenesis in the mouse.
      ). Genetically, this study further predicts the existence of multiple Paibp2-like factors that orchestrate translational activation during spermiogenesis (Fig. 10; also see Supplementary Discussion).
      It is quite interesting that a relatively high percentage of retrogenes are abundantly expressed specifically in spermatids, many of which encode glycolytic enzymes (
      • Vemuganti S.A.
      • de Villena F.P.
      • O'Brien D.A.
      Frequent and recent retrotransposition of orthologous genes plays a role in the evolution of sperm glycolytic enzymes.
      ). Five of the top seven most-enriched 2D gel spots in elongating spermatid fractions contained proteins predicted to be expressed from retrogene-like elements (Table II). Again, this is similar to the GESptd1 rat lentiviral transgene, which was identified in the 11th most-enriched protein spot in elongating spermatid fractions (Table II). A recent study by Vemuganti and colleagues used bioinformatics to study this relationship, and found an over-representation of LINE and LTR elements flanking glycolytic retrogenes in rodent and primate genomes (
      • Vemuganti S.A.
      • de Villena F.P.
      • O'Brien D.A.
      Frequent and recent retrotransposition of orthologous genes plays a role in the evolution of sperm glycolytic enzymes.
      ). Likewise, EGFP is expressed from a LTR-based lentiviral retroelement we experimentally integrated directly into the germline of GESptd1 rats (Fig. 1A). Thus, in addition to chromatin factors regulating transcript levels, the striking abundance of proteins expressed from such retrogenes invokes questions regarding the existence of proteinacious and/or RNA translational enhancers that operate with their mRNA sequences during spermiogenesis (Fig. 10). Additionally, chromosomal DNA generated by reverse transcription, including retrogenes and other LTR-expressed factors, may lack elements that negatively regulate protein synthesis from their encoded mRNAs during spermatid elongation.
      The unusually robust expression, and sperm-specific nature of retroelements seems remarkably well tailored for driving sperm fitness by selecting activities that promote ATP production while generating relatively low levels of oxidative stress (
      • Storey B.T.
      Mammalian sperm metabolism: oxygen and sugar, friend and foe.
      ). In fact, the reproductive advantage of mammalian glycolytic genes expressed during sperm maturation is well documented genetically in mice (
      • Miki K.
      • Qu W.
      • Goulding E.H.
      • Willis W.D.
      • Bunch D.O.
      • Strader L.F.
      • Perreault S.D.
      • Eddy E.M.
      • O'Brien D.A.
      Glyceraldehyde 3-phosphate dehydrogenase-S, a sperm-specific glycolytic enzyme, is required for sperm motility and male fertility.
      ,
      • Danshina P.V.
      • Geyer C.B.
      • Dai Q.
      • Goulding E.H.
      • Willis W.D.
      • Kitto G.B.
      • McCarrey J.R.
      • Eddy E.M.
      • O'Brien D.A.
      Phosphoglycerate kinase 2 (PGK2) is essential for sperm function and male fertility in mice.
      ,
      • Odet F.
      • Duan C.
      • Willis W.D.
      • Goulding E.H.
      • Kung A.
      • Eddy E.M.
      • Goldberg E.
      Expression of the gene for mouse lactate dehydrogenase C (Ldhc) is required for male fertility.
      ), and three of the six core glycolytic enzymes listed Table II represent retrogenes. Thus, it is fitting that glycolytic and redox enzymes were also identified as enriched in the elongating spermatid fraction isolated from GESptd1 rats. Theoretically, this glycolytic power endows spermatozoa in many nonruminant mammalian species greater potential for driving motility (
      • Storey B.T.
      Mammalian sperm metabolism: oxygen and sugar, friend and foe.
      ), thus providing their haploid genomes an advantage for sexually equilibrating into a population (
      • Morrow E.H.
      How the sperm lost its tail: the evolution of aflagellate sperm.
      ). Similarly, it can now be asked how increased heterogeneity and abundance of the round spermatid RNA binding proteome would select for sperm fitness.
      Along with regulating the time, location and abundance of alternative mRNAs expressed in spermatids, additional clues into this later question may emerge from studies demonstrating that splicing elements in spermatogenic cells largely incorporate RNA binding proteins that select against exon inclusion (
      • Wang X.
      • Wang K.
      • Radovich M.
      • Wang Y.
      • Wang G.
      • Feng W.
      • Sanford J.R.
      • Liu Y.
      Genome-wide prediction of cis-acting RNA elements regulating tissue-specific pre-mRNA alternative splicing.
      ). Consequently, variants produced by exon skipping uniquely in spermatogenic cells could select for gamete fitness by eliminating RNA/protein coding domains that constrain spermatozoan development and function, but which are beneficial to reproductive success when expressed in somatic tissue. Another intriguing hypothesis alluded to above reflects additional roles of the germline RNA binding proteome in restricting replication and coordinating expression from integrated reverse-transcribed sequences that escape transcriptional repression specifically during gametogenesis (as modeled in GESptd1 rats) (
      • Tanaka T.
      • Hosokawa M.
      • Vagin V.V.
      • Reuter M.
      • Hayashi E.
      • Mochizuki A.L.
      • Kitamura K.
      • Yamanaka H.
      • Kondoh G.
      • Okawa K.
      • Kuramochi-Miyagawa S.
      • Nakano T.
      • Sachidanandam R.
      • Hannon G.J.
      • Pillai R.S.
      • Nakatsuji N.
      • Chuma S.
      Tudor domain containing 7 (Tdrd7) is essential for dynamic ribonucleoprotein (RNP) remodeling of chromatoid bodies during spermatogenesis.
      ,
      • Thomson T.
      • Lin H.
      The biogenesis and function of PIWI proteins and piRNAs: progress and prospect.
      ). As such, deleterious retroelement-derived copy number variation could further drive purifying selection so to enrich genomes with replication-restricted or deficient integrants expressed in nonmitotic cells. Thus, evolutionarily, spermiogenetic down-regulation of the RNA binding proteome, as observed here in GESptd1 rats, would further capture benefits gained from retroelements post-transcriptionally, and selectively during spermatid elongation.

      Acknowledgments

      We thank Gerardo Medrano, Bray S. Denard and Dalia Saidley-Alsaadi for their help with these experiments.

      Supplementary Material

      REFERENCES

        • David C.J.
        • Chen M.
        • Assanah M.
        • Canoll P.
        • Manley J.L.
        HnRNP proteins controlled by c-Myc deregulate pyruvate kinase mRNA splicing in cancer.
        Nature. 2010; 463: 364-368
        • Kir S.
        • Beddow S.A.
        • Samuel V.T.
        • Miller P.
        • Previs S.F.
        • Suino-Powell K.
        • Xu H.E.
        • Shulman G.I.
        • Kliewer S.A.
        • Mangelsdorf D.J.
        FGF19 as a postprandial, insulin-independent activator of hepatic protein and glycogen synthesis.
        Science. 2011; 331: 1621-1624
        • Licatalosi D.D.
        • Yano M.
        • Fak J.J.
        • Mele A.
        • Grabinski S.E.
        • Zhang C.
        • Darnell R.B.
        Ptbp2 represses adult-specific splicing to regulate the generation of neuronal precursors in the embryonic brain.
        Genes Dev. 2012; 26: 1626-1642
        • Zhu H.
        • Shyh-Chang N.
        • Segre A.V.
        • Shinoda G.
        • Shah S.P.
        • Einhorn W.S.
        • Takeuchi A.
        • Engreitz J.M.
        • Hagan J.P.
        • Kharas M.G.
        • Urbach A.
        • Thornton J.E.
        • Triboulet R.
        • Gregory R.I.
        • Altshuler D.
        • Daley G.Q.
        The Lin28/let-7 axis regulates glucose metabolism.
        Cell. 2011; 147: 81-94
        • Rouget C.
        • Papin C.
        • Boureux A.
        • Meunier A.C.
        • Franco B.
        • Robine N.
        • Lai E.C.
        • Pelisson A.
        • Simonelig M.
        Maternal mRNA deadenylation and decay by the piRNA pathway in the early Drosophila embryo.
        Nature. 2010; 467: 1128-1132
        • Ghosh S.
        • Marchand V.
        • Gaspar I.
        • Ephrussi A.
        Control of RNP motility and localization by a splicing-dependent structure in oskar mRNA.
        Nat. Struct. Mol. Biol. 2012; 19: 441-449
        • Kadyrova L.Y.
        • Habara Y.
        • Lee T.H.
        • Wharton R.P.
        Translational control of maternal Cyclin B mRNA by Nanos in the Drosophila germline.
        Development. 2007; 134: 1519-1527
        • Li Y.
        • Maines J.Z.
        • Tastan O.Y.
        • McKearin D.M.
        • Buszczak M.
        Mei-P26 regulates the maintenance of ovarian germline stem cells by promoting BMP signaling.
        Development. 2012; 139: 1547-1556
        • Henderson M.A.
        • Cronland E.
        • Dunkelbarger S.
        • Contreras V.
        • Strome S.
        • Keiper B.D.
        A germline-specific isoform of eIF4E (IFE-1) is required for efficient translation of stored mRNAs and maturation of both oocytes and sperm.
        J. Cell Sci. 2009; 122: 1529-1539
        • Flemr M.
        • Ma J.
        • Schultz R.M.
        • Svoboda P.
        P-body loss is concomitant with formation of a messenger RNA storage domain in mouse oocytes.
        Biol. Reprod. 2010; 82: 1008-1017
        • Jackson R.J.
        • Standart N.
        Do the poly(A) tail and 3′ untranslated region control mRNA translation?.
        Cell. 1990; 62: 15-24
        • Schmidt E.E.
        • Hanson E.S.
        • Capecchi M.R.
        Sequence-independent assembly of spermatid mRNAs into messenger ribonucleoprotein particles.
        Mol. Cell. Biol. 1999; 19: 3904-3915
        • Kleene K.C.
        Patterns of translational regulation in the mammalian testis.
        Mol. Reprod. Dev. 1996; 43: 268-281
        • Yang J.
        • Medvedev S.
        • Reddi P.P.
        • Schultz R.M.
        • Hecht N.B.
        The DNA/RNA-binding protein MSY2 marks specific transcripts for cytoplasmic storage in mouse male germ cells.
        Proc. Natl. Acad. Sci. U.S.A. 2005; 102: 1513-1518
        • Vemuganti S.A.
        • de Villena F.P.
        • O'Brien D.A.
        Frequent and recent retrotransposition of orthologous genes plays a role in the evolution of sperm glycolytic enzymes.
        BMC Genomics. 2010; 11: 285
        • He C
        • Zuo Z.
        • Chen H.
        • Zhang L.
        • Zhou F.
        • Cheng H.
        • Zhou R.
        Genome-wide detection of testis- and testicular cancer-specific alternative splicing.
        Carcinogenesis. 2007; 28: 2484-2490
        • Harr B.
        • Turner L.M.
        Genome-wide analysis of alternative splicing evolution among Mus subspecies.
        Mol. Ecol. 2010; 19: 228-239
        • Kleene K.C.
        Poly(A) shortening accompanies the activation of translation of five mRNAs during spermiogenesis in the mouse.
        Development. 1989; 106: 367-373
        • Yanagiya A.
        • Delbes G.
        • Svitkin Y.V.
        • Robaire B.
        • Sonenberg N.
        The poly(A)-binding protein partner Paip2a controls translation during late spermiogenesis in mice.
        J. Clin. Invest. 2010; 120: 3389-3400
        • Kleene K.C.
        • Distel R.J.
        • Hecht N.B.
        Translational regulation and deadenylation of a protamine mRNA during spermiogenesis in the mouse.
        Dev. Biol. 1984; 105: 71-79
        • Amann R.P.
        The cycle of the seminiferous epithelium in humans: a need to revisit?.
        J. Androl. 2008; 29: 469-487
        • Oatley J.M.
        • Brinster R.L.
        The germline stem cell niche unit in mammalian testes.
        Physiol. Rev. 2012; 92: 577-595
        • Dym M.
        • Fawcett D.W.
        The blood-testis barrier in the rat and the physiological compartmentation of the seminiferous epithelium.
        Biol Reprod. 1970; 3: 308-326
        • Clermont Y.
        Kinetics of spermatogenesis in mammals: seminiferous epithelium cycle and spermatogonial renewal.
        Physiol. Rev. 1972; 52: 198-236
        • Eddy E.M.
        Male germ cell gene expression.
        Recent Prog Horm Res. 2002; 57: 103-128
        • Schultz N.
        • Hamra F.K.
        • Garbers D.L.
        A multitude of genes expressed solely in meiotic or postmeiotic spermatogenic cells offers a myriad of contraceptive targets.
        Proc. Natl. Acad. Sci. U.S.A. 2003; 100: 12201-12206
        • Johnston D.S.
        • Wright W.W.
        • Dicandeloro P.
        • Wilson E.
        • Kopf G.S.
        • Jelinsky S.A.
        Stage-specific gene expression is a fundamental characteristic of rat spermatogenic cells and Sertoli cells.
        Proc. Natl. Acad. Sci. U.S.A. 2008; 105: 8315-8320
        • Miki K.
        • Qu W.
        • Goulding E.H.
        • Willis W.D.
        • Bunch D.O.
        • Strader L.F.
        • Perreault S.D.
        • Eddy E.M.
        • O'Brien D.A.
        Glyceraldehyde 3-phosphate dehydrogenase-S, a sperm-specific glycolytic enzyme, is required for sperm motility and male fertility.
        Proc. Natl. Acad. Sci. U.S.A. 2004; 101: 16501-16506
        • Danshina P.V.
        • Geyer C.B.
        • Dai Q.
        • Goulding E.H.
        • Willis W.D.
        • Kitto G.B.
        • McCarrey J.R.
        • Eddy E.M.
        • O'Brien D.A.
        Phosphoglycerate kinase 2 (PGK2) is essential for sperm function and male fertility in mice.
        Biol. Reprod. 2010; 82: 136-145
        • Odet F.
        • Duan C.
        • Willis W.D.
        • Goulding E.H.
        • Kung A.
        • Eddy E.M.
        • Goldberg E.
        Expression of the gene for mouse lactate dehydrogenase C (Ldhc) is required for male fertility.
        Biol. Reprod. 2008; 79: 26-34
        • Monesi V.
        Ribonucleic Acid Synthesis during Mitosis and Meiosis in the Mouse Testis.
        J. Cell Biol. 1964; 22: 521-532
        • Kierszenbaum A.L.
        • Tres L.L.
        Structural and transcriptional features of the mouse spermatid genome.
        J. Cell Biol. 1975; 65: 258-270
        • Hamra F.K.
        • Gatlin J.
        • Chapman K.M.
        • Grellhesl D.M.
        • Garcia J.V.
        • Hammer R.E.
        • Garbers D.L.
        Production of transgenic rats by lentiviral transduction of male germ-line stem cells.
        Proc. Natl. Acad. Sci. U.S.A. 2002; 99: 14931-14936
        • Dyck M.K.
        • Ouellet M.
        • Gagn M.
        • Petitclerc D.
        • Sirard M.A.
        • Pothier F.
        Testes-specific transgene expression in insulin-like growth factor-I transgenic mice.
        Mol. Reprod. Dev. 1999; 54: 32-42
        • Mehta A.K.
        • Majumdar S.S.
        • Alam P.
        • Gulati N.
        • Brahmachari V.
        Epigenetic regulation of cytomegalovirus major immediate-early promoter activity in transgenic mice.
        Gene. 2009; 428: 20-24
        • Delmas V.
        • van der Hoorn F.
        • Mellstrom B.
        • Jegou B.
        • Sassone-Corsi P.
        Induction of CREM activator proteins in spermatids: down-stream targets and implications for haploid germ cell differentiation.
        Mol Endocrinol. 1993; 7: 1502-1514
        • Reijo R.A.
        • Dorfman D.M.
        • Slee R.
        • Renshaw A.A.
        • Loughlin K.R.
        • Cooke H.
        • Page D.C.
        DAZ family proteins exist throughout male germ cell development and transit from nucleus to cytoplasm at meiosis in humans and mice.
        Biol Reprod. 2000; 63: 1490-1496
        • Kimura M.
        • Ishida K.
        • Kashiwabara S.
        • Baba T.
        Characterization of two cytoplasmic poly(A)-binding proteins, PABPC1 and PABPC2, in mouse spermatogenic cells.
        Biol. Reprod. 2009; 80: 545-554
        • Matsui M.
        • Horiguchi H.
        • Kamma H.
        • Fujiwara M.
        • Ohtsubo R.
        • Ogata T.
        Testis- and developmental stage-specific expression of hnRNP A2/B1 splicing isoforms, B0a/b.
        Biochim. Biophys. Acta. 2000; 1493: 33-40
        • Kamma H.
        • Horiguchi H.
        • Wan L.
        • Matsui M.
        • Fujiwara M.
        • Fujimoto M.
        • Yazawa T.
        • Dreyfuss G.
        Molecular characterization of the hnRNP A2/B1 proteins: tissue-specific expression and novel isoforms.
        Exp. Cell Res. 1999; 246: 399-411
        • Guillaume E.
        • Evrard B.
        • Com E.
        • Moertz E.
        • Jegou B.
        • Pineau C.
        Proteome analysis of rat spermatogonia: reinvestigation of stathmin spatio-temporal expression within the testis.
        Mol. Reprod. Dev. 2001; 60: 439-445
        • Xu M.
        • Hecht N.B.
        Polypyrimidine tract binding protein 2 stabilizes phosphoglycerate kinase 2 mRNA in murine male germ cells by binding to its 3′UTR.
        Biol. Reprod. 2007; 76: 1025-1033
        • Hosokawa M.
        • Shoji M.
        • Kitamura K.
        • Tanaka T.
        • Noce T.
        • Chuma S.
        • Nakatsuji N.
        Tudor-related proteins TDRD1/MTR-1, TDRD6 and TDRD7/TRAP: domain composition, intracellular localization, and function in male germ cells in mice.
        Dev. Biol. 2007; 301: 38-52
        • Tanaka T.
        • Hosokawa M.
        • Vagin V.V.
        • Reuter M.
        • Hayashi E.
        • Mochizuki A.L.
        • Kitamura K.
        • Yamanaka H.
        • Kondoh G.
        • Okawa K.
        • Kuramochi-Miyagawa S.
        • Nakano T.
        • Sachidanandam R.
        • Hannon G.J.
        • Pillai R.S.
        • Nakatsuji N.
        • Chuma S.
        Tudor domain containing 7 (Tdrd7) is essential for dynamic ribonucleoprotein (RNP) remodeling of chromatoid bodies during spermatogenesis.
        Proc. Natl. Acad. Sci. U.S.A. 2011; 108: 10579-10584
        • Seidah N.G.
        • Day R.
        • Hamelin J.
        • Gaspar A.
        • Collard M.W.
        • Chretien M.
        Testicular expression of PC4 in the rat: molecular diversity of a novel germ cell-specific Kex2/subtilisin-like proprotein convertase.
        Mol. Endocrinol. 1992; 6: 1559-1570
        • Vasileva A.
        • Tiedau D.
        • Firooznia A.
        • Muller-Reichert T.
        • Jessberger. R
        Tdrd6 is required for spermiogenesis, chromatoid body architecture, and regulation of miRNA expression.
        Curr. Biol. 2009; 19: 630-639
        • Yu Y.
        • Zhao C.
        • Lv Z.
        • Chen W.
        • Tong M.
        • Guo X.
        • Wang L.
        • Liu J.
        • Zhou Z.
        • Zhu H.
        • Zhou Q.
        • Sha J.
        Microinjection manipulation resulted in the increased apoptosis of spermatocytes in testes from intracytoplasmic sperm injection (ICSI) derived mice.
        PLoS One. 2011; 6: e22172
        • Weber P.
        • Cammas F.
        • Gerard C.
        • Metzger D.
        • Chambon P.
        • Losson R.
        • Mark M.
        Germ cell expression of the transcriptional co-repressor TIF1beta is required for the maintenance of spermatogenesis in the mouse.
        Development. 2002; 129: 2329-2337
        • Tanaka H.
        • Iguchi N.
        • Egydio de Carvalho C.
        • Tadokoro Y.
        • Yomogida K.
        • Nishimune Y.
        Novel actin-like proteins T-ACTIN 1 and T-ACTIN 2 are differentially expressed in the cytoplasm and nucleus of mouse haploid germ cells.
        Biol Reprod. 2003; 69: 475-482
        • McCarrey J.R.
        • Berg W.M.
        • Paragioudakis S.J.
        • Zhang P.L.
        • Dilworth D.D.
        • Arnold B.L.
        • Rossi J.J.
        Differential transcription of Pgk genes during spermatogenesis in the mouse.
        Dev. Biol. 1992; 154: 160-168
        • Tsunekawa N.
        • Matsumoto M.
        • Tone S.
        • Nishida T.
        • Fujimoto H.
        The Hsp70 homolog gene, Hsc70t, is expressed under translational control during mouse spermiogenesis.
        Mol. Reprod. Dev. 1999; 52: 383-391
        • Koga M.
        • Tanaka H.
        • Yomogida K.
        • Nozaki M.
        • Tsuchida J.
        • Ohta H.
        • Nakamura Y.
        • Masai K.
        • Yoshimura Y.
        • Yamanaka M.
        • Iguchi N.
        • Nojima H.
        • Matsumiya K.
        • Okuyama A.
        • Nishimune Y.
        Isolation and characterization of a haploid germ cell-specific novel complementary deoxyribonucleic acid; testis-specific homologue of succinyl CoA:3-Oxo acid CoA transferase.
        Biol. Reprod. 2000; 63: 1601-1609
        • Vemuganti S.A.
        • Bell T.A.
        • Scarlett C.O.
        • Parker C.E.
        • de Villena F.P.
        • O'Brien D.A.
        Three male germline-specific aldolase A isozymes are generated by alternative splicing and retrotransposition.
        Dev. Biol. 2007; 309: 18-31
        • Shi H.J.
        • Wu A.Z.
        • Santos M.
        • Feng Z.M.
        • Huang L.
        • Chen Y.M.
        • Zhu K.
        • Chen C.L.
        Cloning and characterization of rat spermatid protein SSP411: a thioredoxin-like protein.
        J. Androl. 2004; 25: 479-493
        • Sasagawa I.
        • Matsuki S.
        • Suzuki Y.
        • Iuchi Y.
        • Tohya K.
        • Kimura M.
        • Nakada T.
        • Fujii J.
        Possible involvement of the membrane-bound form of peroxiredoxin 4 in acrosome formation during spermiogenesis of rats.
        Eur. J. Biochem. 2001; 268: 3053-3061
        • Swiston J.
        • Hubberstey A.
        • Yu G.
        • Young D.
        Differential expression of CAP and CAP2 in adult rat tissues.
        Gene. 1995; 165: 273-277
        • Fundele R.
        • Winking H.
        • Illmensee K.
        • Jagerbauer E.M.
        Developmental activation of phosphoglycerate mutase-2 in the testis of the mouse.
        Dev. Biol. 1987; 124: 562-566
        • Lie-Venema H.
        • de Boer P.A.
        • Moorman A.F.
        • Lamers W.H.
        Role of the 5′ enhancer of the glutamine synthetase gene in its organ-specific expression.
        Biochem. J. 1997; 323: 611-619
        • Russell D.L.
        • Kim K.H.
        Expression of triosephosphate isomerase transcripts in rat testis: evidence for retinol regulation and a novel germ cell transcript.
        Biol. Reprod. 1996; 55: 11-18
        • Iida H.
        • Doiguchi M.
        • Yamashita H.
        • Sugimachi S.
        • Ichinose J.
        • Mori T.
        • Shibata Y.
        Spermatid-specific expression of Iba1, an ionized calcium binding adapter molecule-1, in rat testis.
        Biol. Reprod. 2001; 64: 1138-1146
        • Bush L.A.
        • Herr J.C.
        • Wolkowicz M.
        • Sherman N.E.
        • Shore A.
        • Flickinger C.J.
        A novel asparaginase-like protein is a sperm autoantigen in rats.
        Mol. Reprod. Dev. 2002; 62: 233-247
        • Nakamura N.
        • Mori C.
        • Eddy E.M.
        Molecular complex of three testis-specific isozymes associated with the mouse sperm fibrous sheath: hexokinase 1, phosphofructokinase M, and glutathione S-transferase mu class 5.
        Biol. Reprod. 2010; 82: 504-515
        • van Lith M.
        • Karala A.R.
        • Bown D.
        • Gatehouse J.A.
        • Ruddock L.W.
        • Saunders P.T.
        • Benham A.M.
        A developmentally regulated chaperone complex for the endoplasmic reticulum of male haploid germ cells.
        Mol. Biol. Cell. 2007; 18: 2795-2804
        • Gitlits V.M.
        • Toh B.H.
        • Loveland K.L.
        • Sentry J.W.
        The glycolytic enzyme enolase is present in sperm tail and displays nucleotide-dependent association with microtubules.
        Eur. J. Cell Biol. 2000; 79: 104-111
        • Maywood E.S.
        • Chahad-Ehlers S.
        • Garabette M.L.
        • Pritchard C.
        • Underhill P.
        • Greenfield A.
        • Ebling F.J.
        • Akhtar R.A.
        • Kyriacou C.P.
        • Hastings M.H.
        • Reddy A.B.
        Differential testicular gene expression in seasonal fertility.
        J. Biol. Rhythms. 2009; 24: 114-125
        • Held T.
        • Paprotta I.
        • Khulan J.
        • Hemmerlein B.
        • Binder L.
        • Wolf S.
        • Schubert S.
        • Meinhardt A.
        • Engel W.
        • Adham I.M.
        Hspa4l-deficient mice display increased incidence of male infertility and hydronephrosis development.
        Mol. Cell. Biol. 2006; 26: 8099-8108
        • Moffit J.S.
        • Boekelheide K.
        • Sedivy J.M.
        • Klysik J.
        Mice lacking Raf kinase inhibitor protein-1 (RKIP-1) have altered sperm capacitation and reduced reproduction rates with a normal response to testicular injury.
        J. Androl. 2007; 28: 883-890
        • Wang X.
        • Wang K.
        • Radovich M.
        • Wang Y.
        • Wang G.
        • Feng W.
        • Sanford J.R.
        • Liu Y.
        Genome-wide prediction of cis-acting RNA elements regulating tissue-specific pre-mRNA alternative splicing.
        BMC Genomics. 2009; 10: S4
        • Grosso A.R.
        • Gomes A.Q.
        • Barbosa-Morais N.L.
        • Caldeira S.
        • Thorne N.P.
        • Grech G.
        • von Lindern M.
        • Carmo-Fonseca M.
        Tissue-specific splicing factor gene expression signatures.
        Nucleic Acids Res. 2008; 36: 4823-4832
        • Clower C.V.
        • Chatterjee D.
        • Wang Z.
        • Cantley L.C.
        • Vander Heiden M.G.
        • Krainer A.R.
        The alternative splicing repressors hnRNP A1/A2 and PTB influence pyruvate kinase isoform expression and cell metabolism.
        Proc. Natl. Acad. Sci. U.S.A. 2010; 107: 1894-1899
        • Zubovic L.
        • Baralle M.
        • Baralle F.E.
        Mutually exclusive splicing regulates the Nav 1.6 sodium channel function through a combinatorial mechanism that involves three distinct splicing regulatory elements and their ligands.
        Nucleic Acids Res. 2012; 40: 6255-6269
        • Expert-Bezancon A.
        • Le Caer J.P.
        • Marie J.
        Heterogeneous nuclear ribonucleoprotein (hnRNP) K is a component of an intronic splicing enhancer complex that activates the splicing of the alternative exon 6A from chicken beta-tropomyosin pre-mRNA.
        J. Biol. Chem. 2002; 277: 16614-16623
        • Hovhannisyan R.H.
        • Carstens R.P.
        Heterogeneous ribonucleoprotein m is a splicing regulatory protein that can enhance or silence splicing of alternatively spliced exons.
        J. Biol. Chem. 2007; 282: 36265-36274
        • Meng Q.
        • Rayala S.K.
        • Gururaj A.E.
        • Talukder A.H.
        • O'Malley B.W.
        • Kumar R.
        Signaling-dependent and coordinated regulation of transcription, splicing, and translation resides in a single coregulator, PCBP1.
        Proc. Natl. Acad. Sci. U.S.A. 2007; 104: 5866-5871
        • Wang Y.
        • Gao L.
        • Tse S.W.
        • Andreadis A.
        Heterogeneous nuclear ribonucleoprotein E3 modestly activates splicing of tau exon 10 via its proximal downstream intron, a hotspot for frontotemporal dementia mutations.
        Gene. 2010; 451: 23-31
        • Pradeepa M.M.
        • Sutherland H.G.
        • Ule J.
        • Grimes G.R.
        • Bickmore W.A.
        Psip1/Ledgf p52 binds methylated histone H3K36 and splicing factors and contributes to the regulation of alternative splicing.
        PLoS Genet. 2012; 8: e1002717
        • Ge H.
        • Si Y.
        • Wolffe A.P.
        A novel transcriptional coactivator, p52, functionally interacts with the essential splicing factor ASF/SF2.
        Mol. Cell. 1998; 2: 751-759
        • Boutz P.L.
        • Stoilov P.
        • Li Q.
        • Lin C.H.
        • Chawla G.
        • Ostrow K.
        • Shiue L.
        • Ares Jr., M.
        • Black D.L.
        A post-transcriptional regulatory switch in polypyrimidine tract-binding proteins reprograms alternative splicing in developing neurons.
        Genes Dev. 2007; 21: 1636-1652
        • Hamilton B.J.
        • Nichols R.C.
        • Tsukamoto H.
        • Boado R.J.
        • Pardridge W.M.
        • Rigby W.F.
        hnRNP A2 and hnRNP L bind the 3′UTR of glucose transporter 1 mRNA and exist as a complex in vivo.
        Biochem. Biophys. Res. Commun. 1999; 261: 646-651
        • Majumder M.
        • Yaman I.
        • Gaccioli F.
        • Zeenko V.V.
        • Wang C.
        • Caprara M.G.
        • Venema R.C.
        • Komar A.A.
        • Snider M.D.
        • Hatzoglou M.
        The hnRNA-binding proteins hnRNP L and PTB are required for efficient translation of the Cat-1 arginine/lysine transporter mRNA during amino acid starvation.
        Mol. Cell. Biol. 2009; 29: 2899-2912
        • Zheng S.
        • Gray E.E.
        • Chawla G.
        • Porse B.T.
        • O'Dell T.J.
        • Black D.L.
        PSD-95 is post-transcriptionally repressed during early neural development by PTBP1 and PTBP2.
        Nat. Neurosci. 2012; 15 (S1): 381-388
        • Ostareck D.H.
        • Ostareck-Lederer A.
        • Shatsky I.N.
        • Hentze M.W.
        Lipoxygenase mRNA silencing in erythroid differentiation: The 3′UTR regulatory complex controls 60S ribosomal subunit joining.
        Cell. 2001; 104: 281-290
        • Lachke S.A.
        • Alkuraya F.S.
        • Kneeland S.C.
        • Ohn T.
        • Aboukhalil A.
        • Howell G.R.
        • Saadi I.
        • Cavallesco R.
        • Yue Y.
        • Tsai A.C.
        • Nair K.S.
        • Cosma M.I.
        • Smith R.S.
        • Hodges E.
        • Alfadhli S.M.
        • Al-Hajeri A.
        • Shamseldin H.E.
        • Behbehani A.
        • Hannon G.J.
        • Bulyk M.L.
        • Drack A.V.
        • Anderson P.J.
        • John S.W.
        • Maas R.L.
        Mutations in the RNA granule component TDRD7 cause cataract and glaucoma.
        Science. 2011; 331: 1571-1576
        • Fahling M.
        • Mrowka R.
        • Steege A.
        • Martinka P.
        • Persson P.B.
        • Thiele B.J.
        Heterogeneous nuclear ribonucleoprotein-A2/B1 modulate collagen prolyl 4-hydroxylase, alpha (I) mRNA stability.
        J. Biol. Chem. 2006; 281: 9279-9286
        • Hui J.
        • Reither G.
        • Bindereif A.
        Novel functional role of CA repeats and hnRNP L in RNA stability.
        RNA. 2003; 9: 931-936
        • Soderberg M.
        • Raffalli-Mathieu F.
        • Lang M.A.
        Inflammation modulates the interaction of heterogeneous nuclear ribonucleoprotein (hnRNP) I/polypyrimidine tract binding protein and hnRNP L with the 3′untranslated region of the murine inducible nitric-oxide synthase mRNA.
        Mol. Pharmacol. 2002; 62: 423-431
        • Xu M.
        • McCarrey J.R.
        • Hecht N.B.
        A cytoplasmic variant of the KH-type splicing regulatory protein serves as a decay-promoting factor for phosphoglycerate kinase 2 mRNA in murine male germ cells.
        Nucleic Acids Res. 2008; 36: 7157-7167
        • Ma A.S.
        • Moran-Jones K.
        • Shan J.
        • Munro T.P.
        • Snee M.J.
        • Hoek K.S.
        • Smith R.
        Heterogeneous nuclear ribonucleoprotein A3, a novel RNA trafficking response element-binding protein.
        J. Biol. Chem. 2002; 277: 18010-18020
        • Raju C.S.
        • Goritz C.
        • Nord Y.
        • Hermanson O.
        • Lopez-Iglesias C.
        • Visa N.
        • Castelo-Branco G.
        • Percipalle P.
        In cultured oligodendrocytes the A/B-type hnRNP CBF-A accompanies MBP mRNA bound to mRNA trafficking sequences.
        Mol. Biol. Cell. 2008; 19: 3008-3019
        • Dufu K.
        • Livingstone M.J.
        • Seebacher J.
        • Gygi S.P.
        • Wilson S.A.
        • Reed R.
        ATP is required for interactions between UAP56 and two conserved mRNA export proteins, Aly CIP29, to assemble the TREX complex.
        Genes Dev. 2010; 24: 2043-2053
        • Yu Y.E.
        • Zhang Y.
        • Unni E.
        • Shirley C.R.
        • Deng J.M.
        • Russell L.D.
        • Weil M.M.
        • Behringer R.R.
        • Meistrich M.L.
        Abnormal spermatogenesis and reduced fertility in transition nuclear protein 1-deficient mice.
        Proc. Natl. Acad. Sci. U.S.A. 2000; 97: 4683-4688
        • Oliva R.
        • Dixon G.H.
        Vertebrate protamine genes and the histone-to-protamine replacement reaction. Prog Nucleic Acids Res.
        Mol. Biol. 1991; 40: 25-94
        • Meistrich M.L.
        • Trostle-Weige P.K.
        • Lin R.
        • Bhatnagar Y.M.
        • Allis C.D.
        Highly acetylated H4 is associated with histone displacement in rat spermatids.
        Mol. Reprod. Dev. 1992; 31: 170-181
        • Meistrich M.L.
        • Mohapatra B.
        • Shirley C.R.
        • Zhao M.
        Roles of transition nuclear proteins in spermiogenesis.
        Chromosoma. 2003; 111: 483-488
        • Eddy E.M.
        • Toshimori K.
        • O'Brien D.A.
        Fibrous sheath of mammalian spermatozoa.
        Microsc. Res. Tech. 2003; 61: 103-115
        • Bellve A.R.
        • Millette C.F.
        • Bhatnagar Y.M.
        • O'Brien D.A.
        Dissociation of the mouse testis and characterization of isolated spermatogenic cells.
        J. Histochem. Cytochem. 1977; 25: 480-494
        • Wolniak S.M.
        • van der Weele C.M.
        • Deeb F.
        • Boothby T.
        • Klink V.P.
        Extremes in rapid cellular morphogenesis: post-transcriptional regulation of spermatogenesis in Marsilea vestita.
        Protoplasma. 2011; 248: 457-473
        • Gold B.
        • Hecht N.B.
        Differential compartmentalization of messenger ribonucleic acid in murine testis.
        Biochemistry. 1981; 20: 4871-4877
        • Gold B.
        • Stern L.
        • Bradley F.M.
        • Hecht N.B.
        Gene expression during mammalian spermatogenesis. II. Evidence for stage-specific differences in mRNA populations.
        J. Exp. Zool. 1983; 225: 123-134
        • Tseden K.
        • Topaloglu O.
        • Meinhardt A.
        • Dev A.
        • Adham I.
        • Muller C.
        • Wolf S.
        • Bohm D.
        • Schluter G.
        • Engel W.
        • Nayernia K.
        Premature translation of transition protein 2 mRNA causes sperm abnormalities and male infertility.
        Mol. Reprod. Dev. 2007; 74: 273-279
        • Lee K.
        • Haugen H.S.
        • Clegg C.H.
        • Braun R.E.
        Premature translation of protamine 1 mRNA causes precocious nuclear condensation and arrests spermatid differentiation in mice.
        Proc. Natl. Acad. Sci. U.S.A. 1995; 92: 12451-12455
        • Delbes G.
        • Yanagiya A.
        • Sonenberg N.
        • Robaire B.
        PABP interacting protein 2A (PAIP2A) regulates specific key proteins during spermiogenesis in the mouse.
        Biol. Reprod. 2012; 86: 95
        • Chang Y.F.
        • Lee-Chang J.S.
        • Imam J.S.
        • Buddavarapu K.C.
        • Subaran S.S.
        • Sinha-Hikim A.P.
        • Gorospe M.
        • Rao M.K.
        Interaction between microRNAs and actin-associated protein Arpc5 regulates translational suppression during male germ cell differentiation.
        Proc. Natl. Acad. Sci. U.S.A. 2012; 109: 5750-5755
        • Storey B.T.
        Mammalian sperm metabolism: oxygen and sugar, friend and foe.
        Int. J. Dev. Biol. 2008; 52: 427-437
        • Morrow E.H.
        How the sperm lost its tail: the evolution of aflagellate sperm.
        Biol. Rev. Camb. Philos. Soc. 2004; 79: 795-814
        • Thomson T.
        • Lin H.
        The biogenesis and function of PIWI proteins and piRNAs: progress and prospect.
        Annu. Rev. Cell Dev. Biol. 2009; 25: 355-376
        • Zufferey R.
        • Dull T.
        • Mandel R.J.
        • Bukovsky A.
        • Quiroz D.
        • Naldini L.
        • Trono D.
        Self-inactivating lentivirus vector for safe and efficient in vivo gene delivery.
        J. Virol. 1998; 72: 9873-9880
        • Gatlin J.
        • Padgett A.
        • Melkus M.W.
        • Kelly P.F.
        • Garcia J.V.
        Long-term engraftment of nonobese diabetic/severe combined immunodeficient mice with human CD34+ cells transduced by a self-inactivating human immunodeficiency virus type 1 vector.
        Hum. Gene Ther. 2001; 12: 1079-1089
        • Cronkhite J.T.
        • Norlander C.
        • Furth J.K.
        • Levan G.
        • Garbers D.L.
        • Hammer R.E.
        Male and female germline specific expression of an EGFP reporter gene in a unique strain of transgenic rats.
        Dev. Biol. 2005; 284: 171-183
        • Korhonen H.M.
        • Meikar O.
        • Yadav R.P.
        • Papaioannou M.D.
        • Romero Y.
        • Da Ros M.
        • Herrera P.L.
        • Toppari J.
        • Nef S.
        • Kotaja N.
        Dicer is required for haploid male germ cell differentiation in mice.
        PLoS One. 2011; 6: e24821