BioOne.org will be down briefly for maintenance on 14 May 2025 between 18:00-22:00 Pacific Time US. We apologize for any inconvenience.
Registered users receive a variety of benefits including the ability to customize email alerts, create favorite journals list, and save searches.
Please note that a BioOne web account does not automatically grant access to full-text content. An institutional or society member subscription is required to view non-Open Access content.
Contact helpdesk@bioone.org with any questions.
The LH/hCG receptor, a member of the G protein coupled receptor family mediates the cellular actions of LH in the ovary. A considerable amount of information regarding its structure, mechanism of activation, and regulation of expression has emerged in recent years. Here we provide a brief overview of the current information on the structural organization of the receptor and the mechanism of receptor mediated signaling as well as an in-depth discussion on recent developments pertaining to the regulation of receptor expression. Specifically, we describe studies from our laboratory showing that the posttranscriptional regulation of the receptor involves an LH/hCG receptor mRNA-binding protein. We also propose a model to explain the loss of steady-state LH/hCG receptor mRNA levels during receptor down-regulation.
Bolette Bjerregaard, Christine Wrenzycki, Frantisek Strejcek, Jozef Laurincik, Peter Holm, Robert L. Ochs, Christine Rosenkranz, Henrik Callesen, Detlef Rath, Heiner Niemann, Poul Maddox-Hyttel
The expression of nucleolar-related proteins was studied as an indirect marker of the ribosomal RNA (rRNA) gene activation in porcine embryos up to the blastocyst stage produced in vivo and in vitro. A group of the in vivo-developed embryos were cultured with α-amanitin to block the de novo embryonic mRNA transcription. Localization of proteins involved in the rRNA transcription (upstream binding factor [UBF], topoisomerase I, RNA polymerase I [RNA Pol I], and the RNA Pol I-associated factor PAF53) and processing (fibrillarin, nucleophosmin, and nucleolin) was assessed by immunocytochemistry and confocal laser-scanning microscopy, and mRNA expression was determined by semiquantitative reverse transcription-polymerase chain reaction (RT-PCR). These findings were correlated with ultrastructural data and autoradiography following 20-min [3H]uridine incubation. Additionally, expression of the pocket proteins pRb and p130, which are involved in cell-cycle regulation, was assessed by semiquantitative RT-PCR up to the blastocyst stage. Toward the end of third cell cycle, the nuclei in non-α-amanitin-treated, in vivo-produced embryos displayed different stages of transformation of the nuclear precursor bodies (NPBs) into fibrillogranular nucleoli associated with autoradiographic labeling. However, on culture with α-amanitin, NPBs were not transformed into a fibrillogranular nucleolus during this cell cycle, demonstrating that embryonic nucleogenesis requires de novo mRNA transcription. Moreover, immunolocalization of RNA Pol I, but not of UBF, and the mRNA expression of PAF53 and UBF were significantly reduced or absent after culture with α-amanitin, indicating that RNA Pol I, PAF53, and presumably, UBF are derived from de novo embryonic transcription. Embryonic genomic activation was delayed in porcine embryos produced in vitro compared to the in vivo-derived counterparts with respect to mRNAs encoding PAF53 and UBF. Moreover, differences existed in the mRNA expression patterns of pRb between in vivo- and in vitro-developed embryos. These findings show, to our knowledge for the first time, a nucleolus-related gene expression in the preimplantation porcine embryo, and they highlight the differences in quality between in vivo and in vitro-produced embryos.
Vladimir Baran, Antonin Pavlok, Bolette Bjerregaard, Christine Wrenzycki, Doris Hermann, Vlada V. Philimonenko, Georgios Lapathitis, Pavel Hozak, Heiner Niemann, Jan Motlik
The aim of this study was to describe the dynamic changes in the localization of the key nucleolar protein markers, fibrillarin, B23/nucleophosmin, C23/nucleolin, protein Nopp140, during the final stages of bovine oocyte growth. All these proteins were present in the large reticulated nucleoli of oocytes from the small-size category follicles (<1 mm). The entire nucleolus exhibited strong positivity for UBF (upstream binding factor, RNA polymerase I-specific transcription initiation factor), which displayed a dotted staining pattern. In contrast, protein p130 was diffusely distributed throughout the nucleus and excluded from nucleoli. In oocytes approaching the late period of growth (2–3-mm follicles), UBF localization shifted to the nucleolar periphery. Double staining of UBF-p130 revealed a gradual accumulation of p130 at the periphery shell around the nucleolus. In fully grown oocytes (>3-mm follicles), all studied nucleolar proteins were detected in the small compact nucleoli. The cap structure, attached to the compact nucleolus surface, was positive for UBF and PAF53 (subunit of RNA polymerase I). The UBF-positive cap showed a close structural association with p130. It is concluded that, during the process of oocyte nucleolus compaction, UBF and PAF53, proteins involved in the rDNA transcription, are segregated from fibrillarin and Nopp140, proteins essential for early steps of pre-rRNA processing. The observed changes may reflect the transition from pre-rRNA synthesis to pre-rRNA processing as an analysis of the relative abundance of the developmentally important gene transcripts confirmed. In addition, discovered structural association between UBF and p130 suggests a role for pocket proteins in ribosomal gene silencing in mammalian oocytes.
Capacitation is a process that confers fertilizing ability to spermatozoa and this critical event occurs in the development of mammalian spermatozoa during their transit through the female reproductive tract and precedes fertilization. Because spermatozoa are relatively silent in transcription and translation, posttranslational modifications perform the regulatory functions in these cells during capacitation. In this report, we identify a candidate protein, dihydrolipoamide dehydrogenase, which is a post-pyruvate metabolic enzyme, exhibiting tyrosine phosphorylation during hamster spermatozoal capacitation. This is the first report showing dihydrolipoamide dehydrogenase as a phosphoprotein. The cDNA sequence of hamster testes dihydrolipoamide dehydrogenase does not show any variation from the already reported mammalian dihydrolipoamide dehydrogenases. Downregulation of the activity of the hamster spermatozoal enzyme by its specific inhibitor, 5-methoxyindole-2-carboxylic acid, blocks acrosome reaction completely and hyperactivation partially, confirming the role of dihydrolipoamide dehydrogenase in hamster spermatozoal capacitation. We also delineate the temporal involvement of glucose and pyruvate-lactate, showing that the former is required in the earlier stages and the latter for the later stages of hamster spermatozoal capacitation. The essentiality of pyruvate-lactate during hyperactivation and acrosome reaction necessitates the involvement of the post-pyruvate-lactate enzyme, dihydrolipoamide dehydrogenase.
Belclare and Cambridge are prolific sheep breeds, the origins of which involved selecting ewes with exceptionally high litter size records from commercial flocks. The variation in ovulation rate in both breeds is consistent with segregation of a gene (or genes) with a large effect on this trait. Sterile ewes, due to a failure of normal ovarian follicle development, occur in both breeds. New naturally occurring mutations in genes for the oocyte-derived growth factors growth differentiation factor 9 (GDF9) and bone morphogenetic protein 15 (BMP15) are described. These mutations are associated with increased ovulation rate in heterozygous carriers and sterility in homozygous carriers in both breeds. This is the first time that a mutation in the gene for GDF9 has been found that causes increased ovulation rate and infertility in a manner similar to inactivating mutations in BMP15, and shows that GDF9 is essential for normal folliculogenesis in sheep. Furthermore, it is shown, for the first time in any species, that individuals with mutations in both GDF9 and BMP15 have a greater ovulation rate than sheep with either of the mutations separately.
In the course of mammalian spermiogenesis, a unique chromatin remodeling process takes place within elongating and condensing spermatid nuclei. The histone-to-protamine exchange results in efficient packaging and increased stability of the paternal genome. Although not fully understood, this change in chromatin architecture must require a global but transient appearance of endogenous DNA strand breaks because most of the DNA supercoiling is eliminated in the mature sperm. To establish the extent of DNA strand breakage and the stage specificity at which these breaks are created and repaired, we performed a sensitive terminal deoxynucleotidyl transferase (TdT)-mediated dUTP nick-end labeling (TUNEL) assay to detect in situ DNA strand breaks on both mice and human testis cross sections. In the mouse, we established that DNA strand breaks are indeed detected in the whole population of elongating spermatids between stages IX and XI of the seminiferous epithelium cycle perfectly coincident with the chromatin remodeling as revealed by histone H4 hyperacetylation. Similarly, TUNEL analyses performed on human testis sections revealed an elevated and global increase in the levels of DNA strand breaks present in nuclei of round-shaped spermatids also coincident with chromatin remodeling. The demonstration of the global character of the transient DNA strand breaks in mammalian spermiogenesis suggests that deleterious consequences on genetic integrity of the male gamete may arise from any disturbance in the process. In addition, this investigation may shed some light on the origin of the low success rate that has been encountered so far with intracytoplasmic injection procedures making use of round spermatids in humans.
External Ca2 entry into myometrial smooth-muscle cells is important to uterine contraction and hence to labor progression and parturition. Proteins of the transient receptor potential (Trp) channel family are putative capacitative Ca2 entry channels that respond to contractant-generated signals and intracellular Ca2 store depletion. Quantitative reverse transcription-polymerase chain reaction was used to examine the relative expression of TrpC mRNAs in rat myometrium and determine their expression pattern during pregnancy and labor. rTrpC1, rTrpC2, rTrpC4, rTrpC5, rTrpC6, and rTrpC7 mRNAs, but not rTrpC3 mRNA, were expressed in nonpregnant rat myometrium. With the exception of rTrpC7, the resulting products were sequenced and found to be identical with published sequences; new rTrpC7 sequence exhibited >88% homology to mouse and human TrpC7 coding regions. Relative to β-actin mRNA, rTrpC4 mRNA was expressed in the greatest abundance. rTrpC1, 5, and 6 mRNAs were expressed at lower levels, whereas rTrpC2 and 7 mRNAs were barely detectable. This relative expression pattern was also observed throughout the course of gestation. There were no major differences in expression of rTrpC1, 2, 4, or 7 mRNAs between Day 13 and Day 21 of gestation or labor. Rat TrpC5 and TrpC6 mRNA expression decreased in pregnancy but was not altered between Day 13 and Day 21 or in labor. Western blot analysis generally confirmed these observations with respect to protein expression. These data suggest that rTrpC4 may play a major role in regulated Ca2 entry in myometrial cells and throughout pregnancy but do not rule out contributions from other Trp proteins.
In porcine oocytes, acquisition of meiotic competence coincides with a decrease of general transcriptional activity at the end of the oocyte growth phase and, specifically, of ribosomal RNA (rRNA) synthesis in the nucleolus. The present study investigated the regulation of rRNA synthesis during porcine oocyte growth. Localization and expression of components involved in regulation of the rRNA synthesis (the RNA polymerase I-associated factor PAF53, upstream binding factor [UBF], and the pocket proteins p130 and pRb) were assessed by immunocytochemistry and semiquantitative reverse transcription-polymerase chain reaction and correlated with ultrastructural analysis and autoradiography following [3H]uridine incubation in growing and fully grown porcine oocytes. In addition, meiotic resumption, ultrastructure, and expression of p130, UBF, and PAF53 were analyzed in growing and fully grown porcine oocytes cultured with 100 μM butyrolactone I (BL-I), a potent inhibitor of cyclin-dependent kinases, to gain insight concerning the regulation of rRNA transcription during meiotic arrest. Immunocytochemical analysis demonstrated that p130 became colocalized with UBF and PAF53 and that the intensity of the PAF53 labeling decreased toward the end of the oocyte growth phase. These data suggest that the decrease in rRNA synthesis is regulated through inhibition of UBF by p130 as well as by decreased availability of PAF53. Moreover, expression of mRNA encoding PAF53 was decreased at the end of the oocyte growth phase. At the morphological level, these events coincided with inactivation of the nucleolus, as visualized by the transformation of the fibrillogranular nucleolus to an electron-dense fibrillar sphere with remnants of the fibrillar centers at the surface. Meiotic inhibition with 100 μM BL-I had a detrimental effect on the ability of porcine oocytes to resume meiosis and on nucleolus morphology, resulting in a lack of RNA synthetic capability as the fibrillar components, where rRNA transcription and initial processing occur, condensed or even disintegrated.
In all species studied, fertilization induces intracellular Ca2 ([Ca2 ]i) oscillations required for oocyte activation and embryonic development. This species-specific pattern has not been studied in the equine, partly due to the difficulties linked to in vitro fertilization in this species. Therefore, the objective of this study was to use intracytoplasmic sperm injection (ICSI) to investigate fertilization-induced [Ca2 ]i signaling and, possibly, ascertain problems linked to the success of this technology in the horse. In vivo- and in vitro-matured mare oocytes were injected with a single motile stallion sperm. Few oocytes displayed [Ca2 ]i responses regardless of oocyte source and we hypothesized that this may result from insufficient release of the sperm-borne active molecule (sperm factor) into the oocyte. However, permeabilization of sperm membranes with Triton-X or by sonication did not alleviate the deficient [Ca2 ]i responses in mare oocytes. Thus, we hypothesized that a step downstream of release, possibly required for sperm factor function, is not appropriately accomplished in horse oocytes. To test this, ICSI-fertilized horse oocytes were fused to unfertilized mouse oocytes, which are known to respond with [Ca2 ]i oscillations to injection of stallion sperm, and [Ca2 ]i monitoring was performed. Such pairs consistently displayed [Ca2 ]i responses demonstrating that the sperm factor is appropriately released into the ooplasm of horse oocytes, but that these are unable to activate and/or provide the appropriate substrate that is required for the sperm factor delivered by ICSI to initiate oscillations. These findings may have implications to improve the success of ICSI in the equine and other livestock species.
During spermatogenesis, developing germ cells migrate progressively across the seminiferous epithelium. This event requires extensive restructuring of cell-cell actin-based adherens junctions (AJs), such as the ectoplasmic specialization (ES, a testis-specific AJ type), between Sertoli cells and elongating/elongate spermatids. It was postulated that proteases and protease inhibitors worked in a yin-yang relationship to regulate these events. If this is true, then it is anticipated that both proteases and protease inhibitors are found at the ES. Indeed, matrix metalloprotease (MMP)-2, membrane-type 1 (MT1)-MMP and their inhibitor, tissue-inhibitor of metalloproteases (TIMP)-2, were shown to localize at the apical ES. In order to identify the putative MMP substrate as well as the unknown binding ligand for α6β1 integrin in the ES, immunofluorescent microscopy coupled with immunoprecipitation techniques were used to demonstrate that laminin γ3, largely a germ cell product, was present at the apical ES and could form a bona fide complex with β1-integrin. Furthermore, the structural interactions of MMP-2 and MT1-MMP with laminin γ3 and β1-integrin, but not with N-cadherin or nectin-3, have implicated the crucial role of MMP-2/MT1-MMP in the regulation of integrin/laminin-based ES dynamics. Using an in vivo model to study AJ dynamics where adult rats were treated with 1-(2,4-dichlorobenzyl)-indazole-3-carbohydrazide (AF-2364) to disrupt Sertoli-germ cell adhesive function, an induction of active MMP-2, active MT1-MMP and TIMP-2 but not active MMP-9 was detected between 0.5 and 8 h after AF-2364 treatment. This time frame coincided with the depletion of elongating/elongate spermatids from the epithelium, illustrating the synergistic relationships between MMP-2, MT1-MMP, and TIMP-2 in AJ disassembly. Perhaps the most important of all, the use of a specific MMP-2 and MMP-9 inhibitor, (2R)-2-[(4-biphenylylsulfonyl)amino]-3-phenylpropionic acid, could effectively delay the AF-2364-induced elongating/elongate spermatid loss from the epithelium, demonstrating the pivotal role of MMP-2 activation in ES disassembly. Collectively, these studies illustrate that the β1-integrin/laminin γ3 complex is a putative ES-structural protein complex, which is regulated, at least in part, by the activation of MMP-2 involving MT1-MMP and TIMP-2 at the apical ES. The net result of this interaction likely regulates germ cell movement in the seminiferous epithelium.
Six variants of P12, a Kazal-type trypsin inhibitor in the secretion of male mouse accessory sexual glands, were made using single-site mutations including R19L, Y21V, D22G, R43G, K44S, and R45T, based on one-letter-code mutation of amino acids. The other two variants, Nd10 and Cd8, were made using the deletion of 10 and 8 residues from the N- and C-terminals, respectively. Their CD profiles revealed maintenance of the P12 conformation in the seven variants, excluding Cd8, which became unfolded. Only R19L entirely lost the ability while the other variants were as strong as P12 in inhibiting the trypsin digestion of N-benzoyl-Phe-Val-Arg 7-amido-4-methylcoumarin. The immunocytochemical results demonstrated that D22G and Cd8 failed to bind to sperm, Y21V very weakly did so, and the other variants retained their sperm-binding abilities. Concomitantly, the immunocytochemical stainability of each ligand was parallel to its inhibitory effect on 125I-P12-sperm binding, and a synthetic oligopeptide corresponding to residues 18–24 of P12 was able to inhibit P12-sperm binding. The data together concluded that R19 was essential for protease inhibition and D22 and/or Y21 mainly being responsible for the binding of P12 to sperm. The steric arrangement of R19, Y21, and D22 on the tertiary structure of P12 is discussed.
As a result of searching recently available chicken (ch) expressed sequence tag databases, a new Tumor Necrosis Factor Receptor Super Family (TNFRSF) member with similarity to the murine (m) TNFRSF23 decoy receptor (DcR) has been identified. However, by comparison with the mTNFRSF23, there exist at least two splice variants of chTNFRSF23, one of which includes an intracellular death domain (TNFRSF23.v1) characteristic of death receptors, and the other with a truncated cytoplasmic domain of a DcR (named TNFRSF23.v2). These two splice variants of chTNFRSF23 display differential patterns of mRNA expression across various hen tissues, with the highest levels observed within reproductive tissues. More specifically, TNFRSF23.v1 is most highly expressed in preovulatory follicle granulosa cells in the ovary, whereas TNFRSF23.v2 mRNA is found at highest levels in ovarian stromal tissue. Primary culture experiments with granulosa cells determined that expression of TNFRSF23.v1 mRNA was decreased by protein kinase A signaling and enhanced by transforming growth factor (TGF) α treatment. Interestingly, TGFβ1 and signaling via protein kinase C also enhanced levels of TNFRSF23.v1 expression but only in undifferentiated granulosa cells from prehierarchal follicles. Based on patterns of mRNA expression and its endocrine/paracrine regulation, we predict that ovarian chTNFRSF23 represents a modulator of granulosa cell survival and/or differentiation. Finally, the characterization of these receptor variants is of considerable interest from an evolutionary perspective in that they provide additional evidence to support a continuing divergence of TNFRSF members throughout vertebrate evolution.
Inorganic arsenic, an environmental contaminant, produces a variety of stress responses in mammalian cells, including metabolic abnormalities accompanied by growth inhibition and carcinogenesis. Much of the toxicity of arsenic is known to stem from its uncoupling effects on mitochondria. Because previously we had shown that mitochondrial dysfunction can disrupt oocyte and embryo development, we investigated effects of arsenite on meiotic progression and early embryo development in mice. Six-week-old CD-1 mice were treated with 0 (solvent as control), 8 mg/kg (a dose previously established in mice as the maternal no-observed-adverse-effect level), and 16 mg/kg doses of sodium arsenite every 2 days for a total of seven i.p. injections ver a period of 14 days. The incidence of meiotic anomalies, characterized by spindle disruption and/or chromosomal misalignment, was significantly increased in arsenite-treated groups (25% after 8 mg/kg and 62.5% after 16 mg/kg), compared to normal metaphase II in control oocytes. Further, arsenite treatment significantly decreased cleavage rates of zygotes at 24 h, morula formation at 72 h, and development to blastocysts at 96 h in a dose-dependent manner. The total cell number in developed blastocysts did not differ significantly between the 8 mg/kg arsenite treatment and control groups, but was significantly reduced in the 16 mg/kg arsenite treatment group. Moreover, the percentage of apoptotic nuclei was significantly increased in blastocysts following 16 mg/kg arsenite treatment. These data suggest that arsenite causes meiotic aberrations, which may contribute to decreased cleavage and preimplantation development, as well as increased apoptosis.
Myometrial growth and remodeling during pregnancy depends on increased synthesis of interstitial matrix proteins. We hypothesize that the presence of mechanical tension in a specific hormonal environment regulates the expression of extracellular matrix (ECM) components in the uterus. Myometrial tissue was collected from pregnant rats on Gestational Days 0, 12, 15, 17, 19, 21, 22, 23 (labor), and 1 day postpartum and ECM expression was analyzed by Northern blotting. Expression of fibronectin, laminin β2, and collagen IV mRNA was low during early gestation but increased dramatically on Day 23 during labor. Expression of fibrillar collagens (type I and III) peaked Day 19 and decreased near term. In contrast, elastin mRNA remained elevated from midgestation onward. Injection of progesterone (P4) on Days 20–23 (to maintain elevated plasma P4 levels) delayed the onset of labor, caused dramatic reductions in the levels of fibronectin and laminin mRNA, and prevented the fall of collagen III mRNA levels on Day 23. Treatment of pregnant rats with the progesterone receptor antagonist RU486 on Day 19 induced preterm labor on Day 20 and a premature increase in mRNA levels of collagen IV, fibronectin, and laminin. Analysis of the uterine tissue from unilaterally pregnant rats revealed that most of the changes in ECM gene expression occurred specifically in the gravid horn. Our results show a decrease in expression of fibrillar collagens and a coordinated temporal increase in expression of components of the basement membrane near term associated with decreased P4 and increased mechanical tension. These ECM changes contribute to myometrial growth and remodeling during late pregnancy and the preparation for the synchronized contractions of labor.
We examined the modification of the MC31 molecule during capacitation, the acrosome reaction, and studied its role in fertilization. These studies revealed that the molecular mass of MC31 in cauda spermatozoa was approximately 28 000–26 000 Dalton (28–26 kDa). A limited change in molecular mass was seen in capacitated spermatozoa. Treatment of sperm extracts with peptide-N-glycosidase (PN glycosidase) reduced the molecular mass of MC31 in both cauda and capacitated spermatozoa from 28–26 kDa to 23–20 kDa, suggesting that MC31 from both cauda and capacitated spermatozoa is glycosylated, and indicating that capacitation induces minor posttranslational modifications in the structure of the MC31 antigen. The MC31 antigen was redistributed from the midpiece of cauda epididymal spermatozoa to the head and equatorial segment after capacitation and acrosome reaction, respectively, when traced by indirect immunofluorescence under in vitro fertilization (IVF) conditions. Some spermatozoa did not stain for the MC31 antigen and might represent spermatozoa that have shed the antigen. IVF experiments conducted to assess the effect of an anti-MC31 monoclonal antibody (mMC31) revealed that this antibody significantly (P < 0.01) inhibited fertilization of cumulus-invested zona pellucida-intact and the zona pellucida-free oocytes in a dose-dependent manner. However, sperm-oolemma binding was not affected. These findings suggest the MC31 antigen facilitates sperm-oocyte interactions.
The reproductive cycle of wild rabbits (Oryctolagus cuniculus) living in Zembra Island (North Tunisia) is dependent on an external factor, the photoperiod: the gonads are inhibited by long days and stimulated by short days or melatonin implants. Here we studied the role of an internal factor, thyroid hormones and the possible thyroid-gonadal interrelationships, in animals captured on Zembra Island and maintained in natural conditions of photoperiod and temperature. We determined the seasonal profile of the thyroid and testis cycles and investigated the effects of castration and thyroidectomy on the seasonal testosterone and thyroxine cycles. Plasma thyroxine and testosterone levels followed similar, parallel seasonal patterns, with a peak in autumn (October) and low values from January to August. In thyroidectomized animals, plasma testosterone levels, although significantly higher than those in controls (P < 0.001), remained low throughout the 13 mo of the experiment, and no testicular reactivation was observed in the fall. In castrated animals, despite the increase in thyroxine concentration in the 3 mo following castration (P < 0.01), plasma thyroxine levels remained low during the 2 yr of the study. We then investigated the combined effects of long days (16L:8D) and moderately high temperature (25°C) on these two endocrine axes. In constant gonado-inhibiting conditions (16L:8D), whether the temperature was kept constantly high or allowed to fluctuate naturally, no reactivation of the thyroid and testicular axes was observed in the fall. In control animals, the peaks of testosterone and thyroxine concentrations observed in September were larger (P < 0.001) than those in animals subjected to the same natural photoperiod conditions but with constantly high temperature. The lower level of autumnal testis stimulation (P < 0.001) in animals maintained in conditions of constant high temperature (25°C) may be attributed to the low thyroxine levels induced by high temperature. These results clearly confirm that the thyroid and testicular cycles display similar seasonal variations and show that the thyroid and gonadal axes are strictly interdependent. This study provides the first demonstration, for a given species, that the seasonal reactivation of gonad activity is controlled by the thyroid, and thyroid activity is controlled by the gonads.
Spermatogenesis is a complicated process of germ cell differentiation, involving programmatic expression of diverse cell type- and developmental stage-specific genes. To date, the vitamin-A-deficiency (VAD) rats and postnatal rats are two models commonly used to study spermatogenesis. In the present study, we studied the expression of 1185 known genes in the vitamin-A-deficient and retinol-reinitiated spermatogenesis of rat testis using Clontech Atlas rat cDNA expression arrays. The mRNA expression patterns of post-vitamin-A (PVA) testis on Days 15 and 35 were compared with those of the spermatogenic arrested rat testis on Day 0. About 9% (110/1185) of the genes studied were highly expressed. When compared with VAD rat testis on Day 0, 20 and 31 genes were differentially expressed by a factor of twofold or greater on Days 15 and 35, respectively. Four genes (cytochrome P450 17, sulfated glycoprotein 2, protein kinase inhibitor, and cathepsin L) that play important roles in spermatogenesis were selected and their gene expression patterns were confirmed by semiquantitative reverse transcription-polymerase chain reaction. Comparison of the expression patterns of these genes between PVA-VAD and postnatal rat testis in developmentally matched stages revealed substantial differences during the early stages of spermatogenesis. This discrepancy could be caused by either the presence of arrested but mature somatic cells in the PVA-VAD testis that may contribute to a unique gene expression pattern in this model or the direct effect of retinol on spermatogenesis. Therefore, caution is needed in interpreting the gene expression data using the PVA-VAD and postnatal rat models in studying spermatogenesis in rat testes.
Dendritic cells (DCs) are known to play a major role in the induction, maintenance, and regulation of immune responses. Recently, DCs have been described to be present at the feto-maternal interface in human decidua. However, only limited information is available about DC presence, phenotype, and—more importantly—function throughout gestation. Thus, we analyzed local (uterine) and systemic (blood) DCs in a murine model. DBA/2J mated CBA/J females with vaginal plugs were separated and killed on Gestation Days (GDs) 1.5, 3.5, 5.5, 6.5, 7.5, 8.5, 10.5, 13.5, 15.5, or 17.5. Frequency of uterine and blood CD11c DC, phenotype (coexpression of CD8α and major histocompatibility complex class II [MHC II] antigens), and presence of intracellular cytokines (interleukins 12 and 10) were determined by flow cytometry. The morphology of DC in the pregnant uterus was evaluated by immunohistochemistry. In uterus, the relative number of CD11c cells increased from GD 5.5, reaching a plateau on GD 9.5 until GD 17.5, while a transient peak of systemic CD11c cells was found on GD 8.5 and 10.5. The vast majority of uterine DCs were CD8α− and thus, belonged to the myeloid lineage. Interestingly, a significant peak of lymphoid DC was present on GD 1.5 and 5.5. Further, significantly more intracellular interleukin 10 than interleukin 12 was present in CD11c cells. Interestingly, mature DCs (MHC II) were diminished from GD 5.5 to 8.5. Characterization of CD11c cell kinetics in uterus and blood reveals variation of phenotype during pregnancy, pointing toward an eminent immunoregulatory role of DCs throughout gestation at the feto-maternal interface.
Gonadotropins stimulate ovarian proteolytic enzyme activity that is believed to be important for the remodeling of the follicular extracellular matrix. Membrane type 1-matrix metalloproteinase (MT1-MMP) has been identified in vitro as an activator of pro-MMP-2 by forming a complex with tissue inhibitors of metalloproteinase-2 (TIMP-2). In the present study, the expression pattern of MT1-MMP mRNA and the role of MT1-MMP were examined in the ovary using the gonadotropin-treated immature rat model. Ovaries were collected at selected times after eCG or hCG. RNase protection assays revealed a transient increase in MT1-MMP mRNA beginning 4 h after hCG. High expression of MT1-MMP mRNA was localized to the theca-interstitial layer of developing and preovulatory follicles, while low expression was observed in the granulosa cell layer of developing follicles by in situ hybridization. The localization pattern of MT1-MMP mRNA was compared with TIMP-2 mRNA. Both MMP-2 and TIMP-2 mRNA were expressed in the theca layer of preovulatory follicles, showing a similarity to MT1-MMP mRNA expression. To further determine whether MT1-MMP activates pro-MMP-2 in the ovary, crude plasma membrane fractions from preovulatory ovaries were analyzed by gelatin zymography. In plasma membrane fractions, pro-MMP-2 increased around the time of ovulation. Upon incubation, pro-MMP-2 was activated with the highest levels of activation at 12 h post-hCG. The addition of MT1-MMP antibody or excess TIMP-2 to membrane fractions inhibited pro-MMP-2 activation. The increase in MT1-MMP mRNA may be an important part of the mechanism necessary for the efficient generation of active MMP-2 during the ovulatory process.
Male Siberian hamsters (Phodopus sungorus) housed in long days (LD), but not short days (SD) release luteinizing hormone (LH) when exposed to females. This study examined whether this response is specific to a female and identifies the source of a stimulus that induces LH release. Serum concentrations of LH, testosterone (T), follicle stimulating hormone (FSH), and cortisol were examined in all experiments. T concentrations mirrored the LH response; FSH and cortisol were unchanged in response to all stimuli. Exposure to an LD female, irrespective of her reproductive status, but not an SD female, elicited LH release. Exposure to another male did not trigger LH release. Males released LH when allowed physical contact with an anesthetized female, but not when separated from a normally active female, suggesting that tactile or nonvolatile chemosensory stimuli elicit LH release. Urine and secretions collected from the vagina as well as oral, midventral, perineal, and rectal glands, elicited marked behavioral responses in male P. sungorus. Despite these behavioral responses, only feces from females elicited LH release in males. Males released LH in response to feces extracted from the rectum and to cotton swabs that had been rubbed against the rectal mucosa, suggesting that a component of rectal secretions may trigger LH release in male Siberian hamsters. Taken together, these data and previous data from our laboratory indicate that both the production of and the response to a pheromone that triggers the selective release of LH is regulated by day length.
Chilling injury is one of the major limiting factors for achieving optimal cryopreservation of gametes. This study aimed to determine potential chilling-induced damage on several structural aspects of early preantral mouse follicles. Mechanically isolated intact early preantral follicles (type 3b-4) were exposed to 0°C for 1, 5, 10, or 30 min. Control and chilled follicles were analyzed by confocal microscopy after staining for tubulin, F-actin, and chromatin, and by electron microscopy. Chilling for only 1 min was sufficient to cause depolymerization of microtubules in the oocyte and the surrounding granulosa cell layer as evidenced by a substantial decrease in fluorescence intensity after antitubulin labeling. Cooling for longer periods caused alterations in microtubule organization in the follicle-enclosed oocyte. These alterations included the loss of interphase microtubules, concomitant with the formation of perinuclear or cortical microtubule asters and sometimes a complete disappearance of microtubules. The extent of microtubule modification was related to the time of chilling, but was fully reversible after rewarming follicles at 37°C for 1 h. Chilling had only minor effects on the actin-containing elements located predominantly in the oocyte cortex and the transzonal projections. Ultrastructural analysis confirmed that oocyte-somatic cell interactions were present. There was no influence on the chromatin configuration within the follicle-enclosed oocyte. These results indicate that mouse follicles are relatively tolerant to direct chilling injury and, as a consequence, are able to withstand the cooling-warming steps during conventional cryopreservation procedures.
Daily rhythms in the timing of the preovulatory surge and the display of reproductive behavior are reversed in diurnal and nocturnal rodents, but little is known about the neural mechanisms underlying these differences. We examined this issue by comparing a diurnal murid rodent, Arvicanthis niloticus (the grass rat), to a nocturnal one, Rattus norvegicus (the lab rat). In the first study, we established that sequential estradiol and progesterone treatment induces a proestrous-like rise in LH secretion and in the percentage of GnRH neurons that express Fos in grass rats, as is the case in lab rats. Next, we tested the hypothesis that differences in the timing of estrus-related events in diurnal and nocturnal species are caused by differences in rhythms in responsiveness to steroid hormones. We found rhythms in GnRH neuron activity, as indicated by Fos, that were 12 hours out of phase in grass rats and lab rats. These patterns persisted in both species when animals were housed in constant darkness for 5 days, suggesting that they are driven by an endogenous circadian mechanism. These results indicate that steroid-primed grass rats and lab rats are similar with respect to the temporal relationship among estrus-related events, but that the timing of these events relative to the light-dark cycle is dramatically different and that this difference is caused by endogenous circadian mechanisms.
Calcitonin gene-related peptide (CGRP) is a potent vasodilator neuropeptide known to be involved in the regulation of vascular tone. Results of previous studies from our laboratory and others suggest that vascular sensitivity to CGRP is enhanced during pregnancy and that the female sex steroid hormones estradiol-17β (E2) and progesterone (P4) may be involved in this process. We hypothesized that CGRP receptors in the mesenteric artery are increased during pregnancy and with sex steroid hormone treatments. In the present study, we investigated whether pregnancy and female sex steroid hormones modulate the CGRP-receptors CGRP-A and CGRP-B in the mesenteric artery in the rat. The CGRP-A receptor consists of calcitonin receptor-like receptor (CRLR) and receptor activity-modifying protein 1 (RAMP1); however, the CGRP-B receptor needs to be further characterized. Messenger RNA levels for CRLR and RAMP1 were assessed by reverse transcription-polymerase chain reaction, and CGRP-B receptor proteins levels were determined by Western blot analysis. In addition, [125I]CGRP binding was measured by Scatchard analysis. Both mRNA for CGRP-A (CRLR and RAMP1) and the protein for CGRP-B receptors in mesenteric arteries were increased with pregnancy compared to nonpregnant, diestrous animals. A P4 antagonist, RU-486, downregulated and P4 upregulated these receptors in mesenteric arteries (P < 0.05) in pregnant rats. In adult ovariectomized rats, P4 upregulated CRLR and RAMP1 mRNA levels as well as [125I]CGRP-binding sites. The CGRP-B-receptor protein levels were significantly (P < 0.05) elevated by P4 and by combined E2 and P4 treatment. Together with earlier findings, these data suggest that increases in the expression of CGRP-A (CRLR and RAMP1) and CGRP-B receptors in mesenteric arteries may be important in reducing vascular resistance and in vascular adaptations that occur during pregnancy; in addition, P4 may be involved in this process.
The effect of insulin-like growth factor-I (IGF-I) on the concentrations of follicular fluid factors during follicle deviation and the development of dominance was studied in mares in two experiments. Transvaginal ultrasound guidance was used for intrafollicular injection and subsequent sequential sampling of follicular fluid. Treatment involved a single injection of IGF-I into the second-largest follicle (F2) at the expected beginning of deviation (Hour 0) based on diameter (≥20 mm) of the largest follicle (F1). Mares in IGF-I groups were given a dose of 500 μg (experiment 1) or 250, 25, or 2.5 μg (experiment 2). Ablation of F1 at Hour 24 was done in experiment 1, but not in experiment 2. The 500- and 250-μg doses stimulated growth, leading to ovulation of F2 in 10 of 10 and 4 of 5 mares in the two experiments, respectively, compared to 4 of 12 and 0 of 5 in saline-injected controls. These doses prevented (P < 0.05) the increase in IGF binding protein-2 and androstenedione that occurred in F2 of controls and increased (P < 0.05) the concentrations of activin-A, inhibin-A, and vascular endothelial growth factor (VEGF). The 500-μg dose stimulated higher (P < 0.05) concentrations of estradiol, but not until Hour 48, whereas the lower doses were ineffective. In experiment 2, free IGF-I concentrations in F2 at Hour 24 decreased progressively as the dose decreased so that concentrations for the 2.5-μg dose were higher (P < 0.05) than in F2 of controls and similar (not significantly different) to endogenous concentrations in F1. Correspondingly, concentrations of androstenedione in F2 at Hour 24 were lower (P < 0.05) and concentrations of activin-A, inhibin-A, and VEGF were higher (P < 0.05) after treatment of F2 with the 2.5-μg dose than in F2 of controls and were similar to concentrations in F1. Hence, a physiologic intrafollicular dose of IGF-I did not stimulate estradiol production but reduced the production of androstenedione and stimulated the production of activin-A, inhibin-A, and VEGF during follicle selection in mares.
Potassium channels play important roles in many cellular processes, including cell-cycle progression and cell differentiation. In the present study, we investigated the pattern of expression of the mouse ether-à-go-go-related (KCNH2; MERG1A) potassium channel during mouse embryogenic development. Analysis by reverse transcription-polymerase chain reaction revealed maternal MERG1A transcripts until the late 2-cell stage of development, after which MERG1A expression from the zygotic genome was low until the 8-cell stage, then rose in the morula, but was low in trophoblast compared to inner cell mass cells. A trophoblast stem cell line also was shown to express MERG1A mRNA. Immunoblotting of oocytes, blastocysts, and the trophoblast stem cell line revealed different posttranslationally processed forms of MERG1A. Immunofluorescence analysis showed that the subcellular localization of MERG1A varied at different stages of the embryogenic cell cycle. In addition, MERG1A protein levels increased following compaction at the 8-cell stage, and its distribution became polarized. This relocalization of MERG1A was affected by treatment with specific inhibitors of ether-à-go-go-related gene (ERG)-channel function and of actin polymerization. Puromycin treatment of morulae indicated that membrane-associated MERG1A had a half-life of greater than 24 h. The ERG-specific inhibitor E-4031 reduced the incidence of blastocyst formation and the number of cells per blastocyst. These results show that MERG1A is developmentally regulated and suggest that it might play a role in early mouse embryogenic development.
Messenger RNA polyadenylation is one of the processes that control gene expression in all eukaryotic cells and tissues. In mice, two forms of the regulatory polyadenylation protein CstF-64 are found. The gene Cstf2 on the X chromosome encodes this form, and it is expressed in all somatic tissues. The second form, τCstF-64 (encoded by the autosomal gene Cstf2t), is expressed in a more limited set of tissues and cell types, largely in meiotic and postmeiotic male germ cells and, to a smaller extent, in brain. We report here that whereas CstF-64 and τCstF-64 expression in rat tissues resembles their expression in mouse tissues, significant differences also are found. First, unlike in mice, in which CstF-64 was expressed in postmeiotic round and elongating spermatids, rat CstF-64 was absent in those cell types. Second, unlike in mice, τCstF-64 was expressed at significant levels in rat liver. These differences in expression suggest interesting differences in X-chromosomal gene expression between these two rodent species.
Lipocalin-type prostaglandin D synthase (L-PGDS), a bifunctional protein, is expressed in the male reproductive organs of many species. However, the expression and regulation of L-PGDS in rat are still uncertain. The present study investigated the regionalization and regulation of L-PGDS expression in rat testis and epididymis by in situ hybridization and immunohistochemistry under the conditions of sexual maturation, castration, and ethylene dimethane sulfonate (EDS) treatments. In sexually mature rats, L-PGDS mRNA was weakly expressed only in the testicular peritubular cells, whereas L-PGDS immunostaining was highly detected in the Leydig cells by Day 70 postpartum. During sexual maturation, L-PGDS mRNA expression was highly detected in the caput, corpus, and cauda of the epididymis 70 days after birth. Compared with normal L-PGDS expression in adult epididymis, both L-PGDS mRNA expression and protein immunostaining were significantly reduced in the caput, corpus, and cauda epididymis after castration. Testosterone propionate treatment induced a significant increase of L-PGDS expression in the epididymis of castrated rats. Compared with adult rat epididymis, L-PGDS mRNA and protein expression was down-regulated after EDS treatment. Testosterone propionate treatment could induce an increase of L-PGDS mRNA and protein expression in the epididymis of EDS-treated rats. In conclusion, both castration and EDS treatments caused a significant decrease of L-PGDS expression in the epididymis, whereas testosterone propionate treatment could induce an increase of L-PGDS expression in the epididymis of both castrated and EDS-treated rats, indicating that L-PGDS expression in the rat epididymis can be up-regulated by testosterone.
Successful ovulation and implantation processes play a crucial role in female fertility. Adamts-1, a matrix metalloproteinase with disintegrin and thrombospondin motifs, has been suggested to be regulated by the progesterone receptor in the hormonal pathway leading to ovulation. With the primary aim of investigating the role of Adamts-1 in female fertility, we generated Adamts-1 null mice. Forty-five percent of the newborn Adamts-1 null mice die, with death most likely caused by a kidney malformation that becomes apparent at birth. Surviving female null mice were subfertile, whereas males reproduced normally. Ovulation in null females was impaired because of mature oocytes remaining trapped in ovarian follicles. No uterine phenotype was apparent in Adamts-1 null animals. Embryo implantation occurred normally, the uteri were capable of undergoing decidualization, and no morphological changes were observed. These results demonstrate that a functional Adamts-1 is required for normal ovulation to occur, and hence the Adamts-1 gene plays an important role in female fertility, primarily during the tissue remodeling process of ovulation.
Although studies in transgenic mice suggest that estrogen is important for development of the testis, very little is known about the potential role of estrogen in maturation of the primate fetal testis. Therefore, as a first step to determine whether estrogen regulates maturation of the fetal primate testis, we used immunocytochemistry to determine estrogen receptor (ER) α and β expression in the fetal baboon testis. Second, we established methods to quantify ERβ mRNA levels by competitive reverse transcription-polymerase chain reaction in Sertoli cells isolated by laser capture microdissection (LCM) from the fetal baboon testis. ERβ protein expression was abundant in the nuclei of Sertoli, peritubular, and interstitial cells in baboon fetuses at mid (Day 100) and late (Day 165) gestation (term is 184 days). ERβ mRNA level was 0.03 attomole/femtomole 18S rRNA in Sertoli cell nuclei and associated cytoplasm isolated by LCM. ERα was expressed in low level in seminiferous tubules and in moderate level in peritubular cells on Day 165. Germ cells expressed very little ERα or ERβ protein, whereas the baboon fetal epididymis exhibited extensive ERα and ERβ immunostaining at mid- and late gestation. In contrast to the robust expression of ERβ, androgen receptor protein was not demonstrable within the cells of the seminiferous cords but was abundantly expressed in epididymal epithelial cells of the fetal baboon. In summary, the results of this study show that the fetal baboon testis and epididymis expressed the ERα and ERβ, and we suggest that our nonhuman primate baboon model can be used to study the potential role of estrogen on maturation of the fetal testis.
We compared the expression of bovine prolactin-related protein-1 (bPRP-1), placental lactogen (bPL), and pregnancy-associated glycoproteins-1 (bPAG-1) and -9 (bPAG-9) genes in artificially inseminated (AI) and nuclear transferred (NT) cows during the first trimester of gestation using real-time reverse transcription-polymerase chain reaction and in situ hybridization. Placentomal (cotyledonary, caruncular) and interplacentomal (intercotyledonary, intercaruncular) tissues of AI and NT cows carrying either motile (M) or immotile (IM) fetuses were examined. Transcripts for bPL and bPAG-9 were lower (P < 0.01) in the fetal membranes of NT (n = 4) cows at Day 30 of gestation, compared with AI (n = 4) cows. There was no difference in the mean (± SEM) levels of expressions of bPRP-1, bPL, and PAG-1 in the placentomal and interplacentomal tissues of AI (n = 5) and NT (M, n = 4) cows at Day 60 of gestation. The mRNAs for bPRP-1, bPL, bPAG-1, and bPAG-9 genes were higher (P < 0.01) in the caruncular tissue of AI cows, compared with NT (IM, n = 4) cows at Day 60 of gestation. Expression of bPRP-1, bPL, bPAG-1, and bPAG-9 in the placentomal and interplacentomal tissues of the NT (n = 3) group varied considerably more, compared with the AI (n = 4) group at Day 100 of gestation. These findings suggest defective binucleate cell-specific gene transcriptional commands in NT cows.
The epididymal epithelium contributes to formation of a luminal fluid that is essential for the protection of spermatozoa from a variety of insults including changes in oxygen tension. A key regulator of the response to oxygen debt in many cells is hypoxia-inducible factor-1 (HIF-1). A transcription factor composed of α and β subunits, HIF-1 activates genes that mediate oxygen homeostasis and cell survival pathways or trigger cell death responses. Previously we have shown that HIF-1α mRNA is expressed in the adult rat epididymis. Goals of this study were to determine whether HIF-1α protein is activated by ischemia in the rat epididymis, to determine whether epididymal HIF-1α mRNA expression is androgen dependent, and to identify epididymal cell types expressing HIF-1α and β. Immunoblot analysis revealed that HIF-1α protein is primarily present in corpus and cauda of the normoxic epididymis and unaffected by ischemia, whereas HIF-1β was detected equally in all regions and also unaffected by ischemia. HIF-1α mRNA expression in all regions was not affected by 15 days bilateral orchiectomy. Principal cells stained positive for HIF-1α by immunocytochemistry, with the epithelium of initial segment and caput epididymidis staining less intensely than corpus and cauda. HIF-1β immunoreactivity was equally present in principal cells in all regions. Clear, narrow, and basal cells were unreactive for HIF-1α and β. The presence of HIF-1 in normoxic epididymis and the regional distribution of HIF-1α suggests fundamental differences in how proximal and distal regions of the epididymis maintain oxygen homeostasis to protect the epithelium and spermatozoa from hypoxia.
While the expression and epigenetic differences of imprinted genes have been extensively characterized in the mouse and human, little is known about imprinted genes in livestock species. In the current study, eight genes that are imprinted in the human or mouse were investigated in preimplantation bovine embryos. Amplified cDNA was created from three single metaphase II (MII) oocytes or embryos throughout preimplantation development. The imprinted genes Dlk1 and Mest (isoform 1) had no detectable transcripts during preimplantation development. Gnas and Grb10 were expressed in most embryos from the 2-cell to blastocyst stages of development. Mest (isoform 2) was expressed in all oocytes and embryos, except for one blastocyst sample. Ndn and Xist were expressed from the 8–16-cell stage (maternal-to-zygotic transition, MZT) onwards. Sgce was expressed until the MZT, and Nnat in both early (α form) and late (β form) stage embryos. The paternally imprinted genes Gnas, Grb10, and Xist were expressed in both in vitro-fertilized (IVF) and parthenogenetically activated (PA) blastocysts as expected. Of the four maternally imprinted genes expressed in the blastocyst (Mest, Ndn, Nnat, and Sgce), Nnat alone showed differential mRNA expression between IVF and PA blastocysts, suggesting imprinting by this stage of development. In conclusion, seven of the eight genes investigated showed mRNA expression during preimplantation development, indicating a potential role during early development. Also significant is the observation that Nnat is imprinted by the blastocyst stage of development although the other genes are not, indicating a temporal imprinting program.
Cells containing different GnRH peptides currently are thought to have distinct locations and functions in the brain. Lake whitefish is the first salmonid species to have three forms of GnRH peptide in contrast to later-evolving salmonids (salmon and trout) in which only two forms have been identified. Our objective was to isolate the cDNAs that code for these transcripts and to localize the transcripts for the three forms of GnRH in adult lake whitefish brain. Also, we provide phylogenetic analysis of these three whitefish genes based on their preprohormone sequence. From whitefish we isolated cDNAs encoding chicken (c)GnRH-II, salmon (s)GnRH, and the novel whitefish (wf)GnRH. The three cDNAs each encode only one GnRH and are placed in separate groups with phylogenetic analysis. A combination of in situ hybridization and immunocytochemistry with two antisera revealed neurons that expressed protein and/or mRNA for cGnRH-II in the midbrain and hindbrain; sGnRH in the olfactory nerve and bulb, ventral telencephalon, and preoptic area; and wfGnRH in the same latter two brain regions and the hypothalamus. Thus, in the anterior brain, cells containing sGnRH and wfGnRH were in the same brain areas but not at identical locations in the ventral telencephalon and preoptic area. Based on our results, we speculate that both sGnRH and wfGnRH have gonadotropin-releasing roles in the lake whitefish brain.
In all systems examined so far, the G2/M phase transition is controlled by the M-phase promoting factor (MPF), a complex of cdc2 (CDK1) and cyclin B1. Histone H1 kinase activity and MPF components are present in pachytene spermatocytes (PS). However, it has not been demonstrated yet that direct inhibition of MPF activity prevents the G2/M transition in these cells. When roscovitine, a potent inhibitor of CDK1, CDK2, and CDK5 activities, was added to cocultures of PS with Sertoli cells, the number of both secondary spermatocytes and round spermatids formed were lower than in control cultures, despite similar cell viability. This effect of roscovitine was reversible, did not involve the Sertoli cells, and was dependent on the concentration of the inhibitor. Roscovitine did not modify the amount of MPF in these germ cells but inhibited the CDK1- or CDK2-associated histone H1 kinase activity of PS. Hence a functional relationship between cyclin-dependent kinase activity and the spontaneous processing of the first meiotic division and, for the first time, of the second meiotic division of male germ cells is shown.
The Bcl2 modifying factor (Bmf) is a pro-apoptotic member of the Bcl2 family of apoptosis-related proteins that has been shown to initiate apoptosis in response to the loss of attachment of cells from their basal lamina (anoikis). Experimental reduction in intratesticular testosterone concentration brings about the death of spermatids as a consequence of their sloughing from Sertoli cells. Given the role of Bmf in anoikis in other systems, we hypothesized that Bmf would be expressed in germ cells and that its expression and normal distribution might be altered under conditions that induce widespread germ cell loss. To test these hypotheses, we demonstrated that Bmf indeed is expressed in the testis and cloned the full-length rat Bmf cDNA. Immunohistochemistry revealed that Bmf is present in the subacrosomal space of postmeiotic spermatids from step 4 to 16 of spermiogenesis. To test the hypothesis that Bmf expression and distribution are altered by conditions that elicit anoikis, intratesticular testosterone was reduced by implanting Silastic capsules containing testosterone and estradiol into adult rats for 8 weeks. As hypothesized, this resulted in a significant change in Bmf distribution relative to untreated animals. In particular, Bmf exhibited a loss of its normal subacrosomal distribution, becoming redistributed throughout the cytoplasm and nucleus, and appeared in cells in which it is not normally expressed (e.g., pachytene spermatocytes). Additionally, Bmf mRNA expression increased in response to lowered testosterone. These results suggest that Bmf may well be involved in germ cell apoptosis and/or anoikis in response to decreased intratesticular testosterone concentration.
Previous studies revealed that oocytes of different genetic strains (e.g, C57BL/6 and DBA/2) modify maternal and paternal pronuclei differently, affecting early preimplantation development. To determine whether these strain-dependent effects would also apply to oocyte modifications of somatic cell nuclei introduced during cloning procedures, we compared the efficiency of development of parthenogenetic and cloned embryos made with DBA/2, C57BL/6, and (B6D2)F1 oocytes. Our results reveal significant differences in the ability of oocytes of different genetic backgrounds to support parthenogenetic development in different culture media. Additionally, our results reveal oocyte strain-dependent differences in the ability to support cloned embryo development beyond what can be accounted for on the basis of differences in parthenogenesis. Thus, the previously documented differences in oocyte-directed parental genome modification are accompanied in the same strains by differences in the ability of oocytes to modify somatic cell nuclei and support clonal development, raising the possibility that these oocyte functions may be mediated by related mechanisms. These results provide a genetic basis for further studies seeking to identify specific genes that determine oocyte phenotype, as well as genes that determine the success of nuclear reprogramming and clonal development.
The present study tested the hypothesis that chronic hypoxia alters pregnancy-mediated adaptation of Ca2 homeostasis and contractility in the uterine artery. Uterine arteries were isolated from nonpregnant and near-term pregnant ewes of normoxic control or high-altitude (3820 m) hypoxic (oxygen pressure in the blood [PaO2], 60 mm Hg) treatment for 110 days. Contractions and intracellular-free Ca2 concentration ([Ca2 ]i) were measured simultaneously in the same tissue. In normoxic animals, pregnancy increased norepinephrine (NE), but not 5-hydroxy-thymide (5-HT) or KCl, contractile sensitivity in the uterine artery. Chronic hypoxia significantly attenuated NE-induced contractions in the pregnant, but not nonpregnant, uterine arteries. Similarly, 5-HT-mediated contractions of nonpregnant arteries were not changed. In the pregnant uterine artery, chronic hypoxia significantly increased NE-mediated Ca2 mobilization, but decreased the Ca2 sensitivity. In addition, hypoxia increased the calcium ionophore A23187-induced relaxation in pregnant, but not nonpregnant, uterine arteries. However, the A23187-mediated reduction of [Ca2 ]i was significantly impaired in hypoxic arteries. In contrast, hypoxia significantly increased the slope of the [Ca2 ]i-tension relationship of A23187-induced reductions in [Ca2 ]i and tension in the pregnant uterine artery. The results suggest that the contractility of nonpregnant uterine artery is insensitive to moderate chronic hypoxia, but the adaptation of sympathetic tone that normally occurs in the uterine artery during pregnancy is inhibited by chronic hypoxia. In addition, changes in Ca2 sensitivity of myofilaments play a predominant role in the adaptation of uterine artery contractility to pregnancy and chronic hypoxia.
Protein kinase C (PKC) and mitogen-activated protein kinase (MAPK) in cumulus cells are involved in FSH-induced meiotic resumption of cumulus-enclosed oocytes (CEOs), but their regulation and cross talk are unknown. The present experiments were designed to investigate 1) the possible involvement of MAPK cascade in PKC-induced meiotic resumption; 2) the regulation of PKC on MAPK activity in FSH-induced oocyte maturation; and 3) the pattern of PKC and MAPK function in induced meiotic resumption of mouse oocytes. PKC activators, phorbol 12-myristate 13-acetate (PMA) and 1-oleoyl-2-acetyl-sn-glycerol (OAG), induced the meiotic resumption of CEOs and activation of MAPK in cumulus cells, whereas this effect could be abolished by PKC inhibitors, calphostin C and chelerythrine, or MEK inhibitor U0126. These results suggest that PKC might induce the meiotic reinitiation of CEOs by activating MAPK in cumulus cells. Both PKC inhibitors and U0126 inhibited the FSH-induced germinal vesicle breakdown (GVBD) of oocytes and MAPK activation in cumulus cells, suggesting that PKC and MAPK are involved in FSH-induced GVBD of mouse CEOs. Protein synthesis inhibitor cycloheximide (CHX) inhibited FSH- or PMA-induced oocyte meiotic resumption, but not the MAPK activation in cumulus cells. FSH and PKC activators induced the GVBD in denuded oocytes cocultured with cumulus cells in hypoxanthine (HX)-supplemented medium, and this effect could be reversed by U0126. Thus, when activated by FSH and PKC, MAPK may stimulate the synthesis of specific proteins in cumulus cells followed by secretion of an unknown positive factor that is capable of inducing GVBD in oocytes.
Equol (7-hydroxy-3[4′hydroxyphenyl]-chroman) is the major metabolite of the phytoestrogen daidzein, one of the main isoflavones found abundantly in soybeans and soy foods. Equol may be an important biologically active molecule based on recent studies demonstrating that equol can modulate reproductive function. In this study, we examined the effects of equol on prostate growth and LH secretion and determined some of the mechanisms by which it might act. Administration of equol to intact male rats for 4–7 days reduced ventral prostate and epididymal weight and increased circulating LH levels. Using binding assays, we determined that equol specifically binds 5α-dihydrotestosterone (DHT), but not testosterone, dehydroepiandrosterone, or estrogen with high affinity. Equol does not bind the prostatic androgen receptor, and has a modest affinity for recombinant estrogen receptor (ER) β, and no affinity for ERα. In castrated male rats treated with DHT, concomitant treatment with equol blocked DHT's trophic effects on the ventral prostate gland growth and inhibitory feedback effects on plasma LH levels without changes in circulating DHT. Therefore, equol can bind circulating DHT and sequester it from the androgen receptor, thus altering growth and physiological hormone responses that are regulated by androgens. These data suggest a novel model to explain equol's biological properties. The significance of equol's ability to specifically bind and sequester DHT from the androgen receptor have important ramifications in health and disease and may indicate a broad and important usage for equol in the treatment of androgen-mediated pathologies.
We generated a clone of bovine somatic cell nuclear transfer embryos using oocyte pools from defined maternal sources to study nuclear-cytoplasmic interactions. Nucleocytoplasmic hybrids were reconstructed with Bos taurus (Brown Swiss) granulosa cells and oocytes that contained B. taurus A (Simmental), B. taurus B (Simmental), or Bos indicus (Dwarf Zebu) cytoplasm. Another set of embryos was reconstructed with randomly selected Brown Swiss (B. taurus R) oocytes. Embryo transfer resulted in nine (12.5%), nine (13.8%), three (50%), and 11 (16.7%) Day 80 fetuses, of which eight (11.1%), three (4.6%), three (50%), and 10 (15.2%) were viable, respectively. The proportion of viable fetuses was affected by cytoplasm (likelihood ratio test, P < 0.02) and was higher for embryos with B. indicus cytoplasm than for the B. taurus A (P < 0.05) and B (P < 0.01) groups. Furthermore, the proportion of surviving Day 80 fetuses was reduced for B. taurus B as compared with B. taurus A and B. taurus R cytoplasm (P < 0.05 and P < 0.02). Body weight of nucleocytoplasmic hybrid fetuses was not significantly different from Brown Swiss control fetuses produced by artificial insemination (AI), but fetuses reconstructed with random cytoplasts of the same breed as the nuclear donor exhibited overgrowth (P < 0.01) and a higher coefficient of variation in weight. Furthermore, body weight, crown rump length, thorax circumference (P < 0.05), and femur length (P < 0.01) of fetuses with B. taurus A cytoplasm differed from fetuses with B. taurus R cytoplasms. Fetal skin, heart, and liver cells with B. indicus cytoplasm showed a greater increase in number per time period (P < 0.001) and oxygen consumption rate per cell (skin and liver, P < 0.001; heart, P < 0.08) in comparison with their counterparts with B. taurus A cytoplasm. These data point to complex oocyte cytoplasm-dependent epigenetic modifications and/or nuclear DNA-mitochondrial DNA interactions with relevance to nuclear transfer and other reproductive technologies such as ooplasmic transfer in human assisted reproduction.
Ionizing radiation (IR) and consequent induction of DNA double-strand breaks (DSBs) causes activation of the protein ataxia telangiectasia mutated (ATM). Normally, ATM is present as inactive dimers; however, in response to DSBs, the ATM dimer partners cross-phosphorylate each other on serine 1981, and kinase active ATM monomers are subsequently released. We have studied the presence of both nonphosphorylated as well as active serine 1981 phosphorylated ATM (pS1981-ATM) in the mouse testis. In the nonirradiated testis, ATM was present in spermatogonia and spermatocytes until stage VII of the cycle of the seminiferous epithelium, whereas pS1981-ATM was found only to be present in the sex body of pachytene spermatocytes. In response to IR, ATM became activated by pS1981 cross-phosphorylation in spermatogonia and Sertoli cells. Despite the occurrence of endogenous programmed DSBs during the first meiotic prophase and the presence of ATM in both spermatogonia and spermatocytes, pS1981 phosphorylated ATM did not appear in spermatocytes after treatment with IR. These results show that spermatogonial ATM and ATM in the spermatocytes are differentially regulated. In the mitotically dividing spermatogonia, ATM is activated by cross-phosphorylation, whereas during meiosis nonphosphorylated ATM or differently phosphorylated ATM is already active. ATM has been shown to be present at the synapsed axes of the meiotic chromosomes, and in the ATM knock-out mice spermatogenesis stops at pachytene stage IV of the seminiferous epithelium, indicating that indeed nonphosphorylated ATM is functional during meiosis. Additionally, ATM is constitutively phosphorylated in the sex body where its continued presence remains an enigma.
Spermiogenesis is a complex process that is regulated by a plethora of genes and interactions between germ and somatic cells. Here we report a novel mutant mouse strain that carries a transgene insertional/translocational mutation and exhibits dominant male sterility. We named the mutation dominant spermiogenesis defect (Dspd). In the testes of Dspd mutant mice, spermatids detached from the seminiferous epithelium at different steps of the differentiation process before the completion of spermiogenesis. Microinsemination using spermatids collected from the mutant testes resulted in the birth of normal offspring. These observations indicate that the major cause of Dspd infertility is (are) a defect(s) in the Sertoli cell-spermatid interaction or communication in the seminiferous tubules. Fluorescent in situ hybridization (FISH) analysis revealed a translocation between chromosomes 7F and 14C at the transgene insertion site. The deletion of a genomic region of chromosome 7F greater than 1 megabase and containing at least six genes (Cttn, Fadd, Fgf3, Fgf4, Fgf15, and Ccnd1) was associated with the translocation. Cttn encodes the actin-binding protein cortactin. Immunohistochemical analysis revealed localization of cortactin beside elongated spermatids in wild-type testes; abnormality of cortactin localization was found in mutant testes. These data suggest an important role of cortactin in Sertoli cell-spermatid interactions and in the Dspd phenotype.
This article is only available to subscribers. It is not available for individual sale.
Access to the requested content is limited to institutions that have
purchased or subscribe to this BioOne eBook Collection. You are receiving
this notice because your organization may not have this eBook access.*
*Shibboleth/Open Athens users-please
sign in
to access your institution's subscriptions.
Additional information about institution subscriptions can be foundhere