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 processes of making transgenic animals by microinjecting DNA into the pronucleus of a fertilized oocyte or after the transfection of embryonic stem cells are now well established. However, attempts have also been made, with varying degrees of success, to use spermatozoa as a vector for transgenesis in mammals and other vertebrates during the last decade. A number of different approaches for making transgenic spermatozoa have been developed. These include directly incubating mature, isolated spermatozoa with DNA or pretreating mature, isolated spermatozoa before assisted fertilization. Microinjection procedures have also been established to transfect male germ cells directly in vivo within the seminiferous tubules or to reimplant previously isolated male germ cells submitted to in vitro transfection into a recipient testis. The latter two techniques present the advantage of being able to create transgenic progeny simply by mating with wild-type females, which avoids the possibility of interference or damage as a result of assisted fertilization or the manipulation of embryos. The different aspects of sperm-mediated transgenesis are presented.
Blastocyst attachment to the lining of the mammalian uterus during early implantation involves the initial apposition of the trophoblast to the uterine epithelial surface. Osteopontin (OPN) is a glycoprotein component of the extracellular matrix that is secreted by the glandular epithelium of mammalian uteri at the time of implantation. This protein is recognized by several members of the integrin family and promotes cell-cell attachment and adhesion. In the present study, rabbit uteri were examined using Northern and in situ hybridization to evaluate the temporal and spatial distribution of OPN mRNA during early pregnancy. Northern blot analysis demonstrated a dramatic increase in OPN expression on Days 4–7 of pregnancy, corresponding to the rise in circulating progesterone and the time of initial embryo attachment in this species. In situ hybridization analysis revealed OPN mRNA expression on Day 6.75 of pregnancy, which was most prominent on endometrial epithelium. Using immunofluorescence, OPN protein was present on the glandular epithelium on Day 6.75 of pregnancy, but was absent on blastocysts. Further, no expression of OPN mRNA or protein was found in the nonpregnant endometrium. Induction of endometrial OPN expression was observed in unmated rabbits treated with progesterone alone and was prevented by cotreatment with the antiprogestin ZK137.316. Estradiol-17β had no effect on OPN expression by itself, and estrogen priming was not necessary to demonstrate the stimulatory effect of progesterone. In The rabbit uterus, as in other mammalian species studied, OPN is expressed in a stage-specific manner by the endometrial glands during the peri-implantation period and is regulated by progesterone.
Chemokines play a role in endometrial physiology and pathology and may affect endometrial receptivity and menstrual shedding. Chemokines exert their effect by binding to their relevant receptors, the expression levels of which may modulate their action. In the present study, we examined the expression of chemokine receptors CXCR1 and CXCR2 (receptors for interleukin-8) and CCR5 (receptor for RANTES [regulated-on-activation, normal-T-cell-expressed and -secreted], macrophage inflammatory protein [MIP]-1α, and MIP-1β) in human endometrium. Human endometria (n = 35) were grouped according to the menstrual cycle phase and examined by immunohistochemistry for CXCR1, CXCR2, and CCR5. In both epithelial and stromal cells, CXCR1 and CXCR2 immunoreactivity was detected. Staining was most prominent at the apical and basal aspects of epithelial cells. Intense CCR5 immunostaining was observed in epithelial and stromal compartments throughout the menstrual cycle. Epithelial and stromal staining for CXCR1 reached a peak at the midsecretory phase, during which it was significantly higher than the level of staining during the proliferative phase (P < 0.05). Immunostaining for CXCR2 and CCR5 showed no significant variation across the menstrual cycle. Expression of interleukin-8 and RANTES in endometrium, together with the presence of their receptors, suggests that autocrine and paracrine interactions involving these chemokines may participate in endometrial physiology.
The human placenta utilizes both active and passive mechanisms to evade rejection by the maternal immune system. We investigated the pattern of expression of the B7 family of immunomodulatory molecules B7-H1 (PD-L1), B7-2 (CD86), and B7-1 (CD80) at the term maternal-fetal interface. Northern blot and reverse transcription-polymerase chain reaction (RT-PCR) analyses showed that B7-H1 mRNA is abundant in term placenta and that cytotrophoblasts are sources of this message. Immunohistochemistry demonstrated that B7-H1 is constitutively expressed by the syncytiotrophoblast and by extravillous cytotrophoblasts, both of which are juxtaposed to maternal blood and tissue. By contrast, placental stromal cells, including macrophages, lacked the protein. Expression of B7-H1 protein was low in first-trimester placenta compared to second- and third-trimester tissue (P < 0.05) and was enhanced in cultured cytotrophoblasts by treatment with either interferon-γ or epidermal growth factor (P < 0.05), suggesting that one or both of these mediators regulates B7-H1 expression in the placenta. RT-PCR and immunofluorescence analysis of term placental tissue revealed different patterns of expression of the immunostimulatory protein, B7-2. In contrast to B7-H1, B7-2 mRNA and protein were absent in cytotrophoblast cells but present in maternal macrophages and some fetal macrophages. The B7-1 mRNA and protein were absent at the maternal-fetal interface. These studies document expression of the B7 family proteins at the maternal-fetal interface and demonstrate that B7-H1 is positioned such that it could facilitate protection of fetal cells against activated maternal leukocytes. Conversely, B7-2 was absent on trophoblasts and was appropriately localized to fetal and maternal macrophages, which may participate in antigen presentation.
The cystic fibrosis transmembrane conductance regulator (CFTR) and aquaporin-9 (AQP-9) are present in the luminal membrane of the epididymis, where they play an important role in formation of the epididymal fluid. Evidence is accumulating that CFTR regulates other membrane transport proteins besides functioning as a cAMP-activated chloride channel. We have explored the possible interaction between epididymal CFTR and AQP-9 by cloning them from the rat epididymis and expressing them in Xenopus oocytes. The effects of the expressed proteins on oocyte water permeability were studied by immersing oocytes in a hypo-osmotic solution, and the ensuing water flow was measured using a gravimetric method. The results show that AQP-9 alone caused an increase in oocyte water permeability, which could be further potentiated by CFTR. This potentiation was markedly reduced by phloretin and lonidamine (inhibitors of AQP-9 and CFTR, respectively). The regulation of water permeability by CFTR was also demonstrated in intact rat epididymis luminally perfused with a hypo-osmotic solution. Osmotic water reabsorption across the epididymal tubule was reduced by phloretin and lonidamine. Elevation of intracellular cAMP with 3-isobutyl-1-methylxanthine increased osmotic water permeability, whereas inhibiting protein kinase A with H-89 (N-(2-[p-bromocinnamylamino]ethyl)-5-isoquinoline sulfonamide hydrochloride) reduced it. These results are consistent with a role for CFTR in controlling water permeability in the epididymis in vivo. We conclude that this additional role of CFTR in controlling water permeability may have an impact on the genetic disease cystic fibrosis, in which men with a mutated CFTR gene have abnormal epididymis and infertility.
The aryl hydrocarbon receptor (AhR) is a ligand-activated transcription factor that binds various environmental contaminants. Despite our knowledge regarding the role of the AhR in mediating toxicity, little is known about the physiological role of the AhR. Previous studies indicate that the AhR may regulate folliculogenesis, because AhR-deficient (AhRKO) mice have fewer preantral and antral follicles than wild-type (WT) mice during postnatal life. Thus, the first objective of the present study was to test the hypothesis that AhR deficiency reduces the numbers of preantral and antral follicles by slowing growth and/or increasing atresia of follicles. Because alterations in follicular growth or atresia can affect the ability to ovulate, the second objective was to test whether AhR deficiency reduces the number of ovulated eggs. To test these hypotheses, follicular growth was compared in WT and AhRKO ovaries using morphometric techniques and by measuring the ability of the ovary and follicles to grow in response to eCG. Atresia was compared in WT and AhRKO ovaries using morphometric techniques, TUNEL assays, and 3′-end labeling of fragmented DNA. Ovulation was compared in WT and AhRKO mice by assessing the number of corpora lutea per ovary. The results indicate that follicular growth and ovulation were reduced in AhRKO ovaries compared to WT ovaries. The WT ovaries had a 1.5-fold increase in the number of preantral and antral follicles between Postnatal Days 32 and 45, were more responsive to eCG, and contained more corpora lutea than AhRKO ovaries. In contrast, no significant difference was observed in the incidence of atresia in WT and AhRKO ovaries. Taken together, these results suggest that the AhR may regulate growth, but not atresia, of preantral and antral follicles in the mouse ovary.
Zona pellucida C (ZPC) is a major glycoprotein of the zona pellucida that possesses the sperm receptor function. ZPC induces autoantibody that can block sperm/oocyte interaction. We selected the partial sequence of rabbit ZPC (amino acid 263–415, rZPC′) as the target and constructed the pCMV4-rZPC′ gene vaccine by using DNA recombinant techniques. The total RNA was extracted from the ovaries of the sexually healthy female rabbit, and the rZPC′ cDNA, which was amplified by reverse transcription-polymerase chain reaction, was directly inserted into the cloning vector PCR2.1 to construct the PCR2.1-rZPC′. This insertion fragment was subcloned into the pCMV4 vector to form the pCMV4-rZPC′ prototype DNA vaccine. All experimental BALB/C mice and New Zealand rabbits received i.m. injection of pCMV4-rZPC′ vaccine three times. The results show that 1) the pCMV4-rZPC′ construct expresses rZPC′ cDNA in mice muscle cells, 2) 60% of the immunized female mice were infertile at 6 wk after the immunization, 3) the mice immunized with pCMV4-rZPC′ DNA vaccine developed anti-rZPC antibodies that bound to the ovarian ZP in situ, and 4) antibodies against rZPC′ were also bound to normal animal ovarian ZP in vitro. The results indicate that anti-rZPC antibodies developed from pCMV4-rZPC′ DNA vaccine can prevent the fertility course and do not interfere with normal follicular development. The pCMV4-rZPC′ DNA vaccine may be possible to develop as a contraceptive vaccine.
Arabinda Mandal, Kenneth L. Klotz, Jagathpala Shetty, Friederike L. Jayes, Michael J. Wolkowicz, Laura C. Bolling, Scott A. Coonrod, Michael B. Black, Alan B. Diekman, Timothy A. J. Haystead, Charles J. Flickinger, John C. Herr
We report the presence of a unique, non-bacteriolytic, c (chicken or conventional type) lysozyme-like protein, SLLP1, in the acrosome of human sperm. C lysozymes are bacteriolytic and can also bind to N-acetylglucosamines linked by β-1,4 glycosidic bonds. Most of the invariant residues (17 out of 20), including all the cysteines, were conserved in SLLP1, but the two catalytic residues E35 and D52 of c lysozymes were replaced with T and N, respectively. The full-length cDNA encodes a protein of 215 aa with a predicted protease cleavage site between A87 and K88. The processed form of SLLP1, which showed an exon-intron organization similar to human c lysozyme, was the major isoform in the acrosome of ejaculated sperm. As expected, based on its sequence, the mature protein secreted from yeast showed no bacteriolytic activity. A significant decrease (54%, P ≤ 0.001) in the number of sperm bound to zona-free hamster eggs was observed in the presence of antisera to recombinant SLLP1. SLLP1 mRNA (size, ∼1 kb) appeared to be expressed only in the testis and in the Burkitt lymphoma Raji cell line. The gene SPACA3 encodes SLLP1 and contains five exons at locus 17q11.2. Because of its typical c lysozyme-like sequence, genomic organization, conservation of putative substrate-binding sites even in the absence of catalytic residues, and localization in the acrosomal matrix, we hypothesize that, after acrosome reaction, SLLP1 could be a potential receptor for the egg oligosaccharide residue N-acetylglucosamine, which is present in the extracellular matrix over the egg plasma membrane, within the perivitelline space, pores of zona pellucida, and cumulus layers.
Nitric oxide (NO) production plays an important role in regulating preimplantation embryo development. NO is produced from l-arginine by the enzyme nitric oxide synthase (NOS), which has three isoforms: endothelial (eNOS), neuronal (nNOS), and inducible (iNOS). It has been previously shown that inhibition of NO production by NG-nitro-l-arginine (l-NA) inhibits the development of two-cell embryos to the four-cell stage. However, excess NO also halts embryo development, possibly through the production of free radicals. We hypothesize that multiple NOS isoforms are expressed in order to ensure normal preimplantation embryo development and that, in this process, NO acts through the cGMP pathway. Using reverse transcription-polymerase chain reaction, mRNA for all three NOS isoforms was amplified from two-cell, four-cell, morula, and blastocyst embryos. However, blastocyst-stage embryos isolated midmorning on Day 4 of pregnancy expressed only nNOS and eNOS, whereas those isolated midafternoon again expressed all three NOS isoforms. Culture of one-cell embryos in various concentrations of Whitten (positive control), S-nitroso-N-acetylpenicillamine (SNP, a NO donor), l-NA, and/or 8-Br-cGMP demonstrated that NO is acting, at least in part, through cGMP in preimplantation embryo development. In addition, we determined that a critical concentration of NO and cGMP is required for normal embryo development and deviations from this concentration lead to developmental arrest and/or apoptosis of the embryo. This data provides support for a requirement of NO in preimplantation embryo development and one mechanism through which it regulates mitotic division in these embryos.
Involvement of Ad4BP/SF-1 in the ovarian cytochrome P-450 aromatase (oP450arom) gene expression was investigated using ovarian follicles of the Nile tilapia, possessing an average 14-day spawning cycle. The promoter region (5′ flanking region) of oP450arom gene cloned from tilapia contains two Ad4 binding sites. Subsequently, a cDNA encoding Ad4BP/SF-1 was cloned from the ovarian follicles. It is expressed in gonadal tissues, brain, and kidney. Oligonucleotide probes containing putative orphan nuclear receptor binding motifs (derived from promoter region of the aromatase gene) formed complexes with in vitro-translated Ad4BP/SF-1 and nuclear extracts of tilapia ovarian (midvitellogenic) follicles, indicating that Ad4BP/SF-1 is one of the transcriptional regulators for aromatase gene expression. Northern blot analysis revealed that the expression of both oP450arom and Ad4BP/SF-1 increased in parallel with ovarian growth from Day 0 to Day 5 after spawning and declined sharply from Day 8 to Day 11. On the day of spawning (Day 14), the expression of both correlates became undetectable. In vitro incubation of post vitellogenic full-grown immature follicles (corresponding to Day 11 after spawning) with hCG purged both oP450arom and Ad4BP/SF-1 messenger RNA transcripts at 18 h. Conversely, in vitro incubation of late vitellogenic follicles (corresponding to Day 8 after spawning) with hCG retained Ad4BP/SF-1 messenger RNA transcripts more or less steadily and up-regulated oP450arom. Ad4BP/SF-1 probably acts as a transcriptional modulator to implement the paradoxical actions of gonadotropins on oP450arom gene.
There is very little information regarding the testis structure and function in domestic cats, mainly data related to the cycle of seminiferous epithelium and sperm production. The testis weight in cats investigated in the present study was 1.2 g. Compared with most mammalian species investigated, the value of 0.08% found for testes mass related to the body mass (gonadosomatic index) in cats is very low. The tunica albuginea volume density (%) in these animals was relatively high and comprised about 19% of the testis. Seminiferous tubule and Leydig cell volume density (%) in cats were approximately 90% and 6%, respectively. The mean tubular diameter was 220 μm, and 23 m of seminiferous tubule were found per testis and per gram of testis. The frequencies of the eight stages of the cycle, characterized according to the tubular morphology system, were as follows: stage 1, 24.9%; stage 2, 12.9%; stage 3, 7.7%; stage 4, 17.6%; stage 5, 7.2%; stage 6, 11.9%; stage 7, 6.8%; and stage 8, 11 %. The premeiotic and postmeiotic stage frequency was 46% and 37%, respectively. The duration of each cycle of seminiferous epithelium was 10.4 days and the total duration of spermatogenesis based on 4.5 cycles was 46.8 days. The number of round spermatids for each pachytene primary spermatocytes (meiotic index) was 2.8, meaning that significant cell loss (30%) occurred during the two meiotic divisions. The total number of germ cells and the number of round spermatids per each Sertoli cell nucleolus at stage 1 of the cycle were 9.8 and 5.1, respectively. The Leydig cell volume was approximately 2000 μm3 and the nucleus volume 260 μm3. Both Leydig and Sertoli cell numbers per gram of testis in cats were approximately 30 million. The daily sperm production per gram of testis in cats (efficiency of spermatogenesis) was approximately 16 million. To our knowledge, this is the first investigation to perform a more detailed and comprehensive study of the testis structure and function in domestic cats. Also, this is the first report in the literature showing Sertoli and Leydig cell number per gram of testis and the daily sperm production in any kind of feline species. In this regard, besides providing a background for comparative studies with other felids, the data obtained in the present work might be useful in future studies in which the domestic cat could be utilized as an appropriate receptor model for preservation of genetic stock from rare or endangered wild felines using the germ cell transplantation technique.
To clarify the mechanism of estradiol-17β production in the ovarian follicle of red seabream, in vitro effects of luteinizing hormone (LH), follicle-stimulating hormone (FSH), and insulin-like growth factor (IGF-I) on aromatase activity (conversion of testosterone to estradiol-17β) and cytochrome P450 aromatase (P450arom) mRNA expression in ovarian fragments of red seabream were investigated. Of the growth factors used in the present study, only IGF-I stimulated both aromatase activity and P450arom gene expression in the ovarian fragments of red seabream. LH from red seabream pituitary, but not FSH, stimulated both aromatase activity and P450arom gene expression. IGF-I slightly enhanced the LH-induced aromatase activity and P450arom gene expression. These data and our previous results indicate that LH, but not FSH, stimulates estradiol-17β production in the ovarian follicle of red seabream through stimulation of aromatase activity and P450arom gene expression and IGF-I enhances the LH-stimulated P450arom gene expression.
Oocytes and embryos of many species, including mammals, contain a unique linker (H1) histone, termed H1oo in mammals. It is uncertain, however, whether other H1 histones also contribute to the linker histone complement of these cells. Using immunofluorescence and radiolabeling, we have examined whether histone H10, which frequently accumulates in the chromatin of nondividing cells, and the somatic subtypes of H1 are present in mouse oocytes and early embryos. We report that oocytes and embryos contain mRNA encoding H10. A polymerase chain reaction-based test indicated that the poly(A) tail did not lengthen during meiotic maturation, although it did so beginning at the four-cell stage. Antibodies raised against histone H10 stained the nucleus of wild-type prophase-arrested oocytes but not of mice lacking the H10 gene. Following fertilization, H10 was detected in the nuclei of two-cell embryos and less strongly at the four-cell stage. No signal was detected in H10 −/− embryos. Radiolabeling revealed that species comigrating with the somatic H1 subtypes H1a and H1c were synthesized in maturing oocytes and in one- and two-cell embryos. Beginning at the four-cell stage in both wild-type and H10 −/− embryos, species comigrating with subtypes H1b, H1d, and H1e were additionally synthesized. These results establish that histone H10 constitutes a portion of the linker histone complement in oocytes and early embryos and that changes in the pattern of somatic H1 synthesis occur during early embryonic development. Taken together with previous results, these findings suggest that multiple H1 subtypes are present on oocyte chromatin and that following fertilization changes in the histone H1 complement accompany the establishment of regulated embryonic gene expression.
The aim of the present study was to examine the acute and chronic effects of the gonadotropin-releasing hormone agonist (GnRH-a) leuprolide acetate (LA) on the expression of the steroidogenic acute regulatory protein (StAR), the cytochrome P450 side-chain cleavage enzyme (P450scc), and steroid production in antral ovarian follicles obtained from prepubertal equine choriogonadotropin (eCG)-treated rats. Follicular contents of StAR and P450scc proteins were measured by Western blotting following in vivo injection of eCG (control) and eCG LA (LA) to prepubertal rats. Treatment with eCG for 2 h resulted in no change in StAR protein content, but it was markedly increased at 4 and 8 h after hormone treatment. However, coadministration of eCG LA produced a significant increase (P < 0.05) in StAR protein levels at 2, 4, and 8 h when compared with eCG treatment. Acute and chronic treatment with either eCG or eCG LA did not alter the P450scc protein levels in freshly isolated follicles. The increase in StAR protein expression following LA treatment was qualitatively similar to StAR mRNA expression, as determined by quantitative reverse transcription-polymerase chain reaction (RT-PCR) analysis. Furthermore, administration of eCG demonstrated a time-dependent increase (2–8 h) in the levels of StAR mRNA, and these levels were markedly increased by eCG LA. However, the temporal response pattern of StAR mRNA was much greater at 2 h following LA administration when compared with controls. In addition, 48 h of LA treatment in eCG-treated rats resulted in a significant increase (P < 0.05) in follicular progesterone levels, whereas significant decreases in androgen (testosterone and androsterone) and estradiol levels were observed. Similar results were obtained when serum androgens and estradiol were measured, but serum progesterone levels were unchanged. Collectively, these findings demonstrate that the inhibitory effect of LA on ovarian androgen and estradiol levels is related to changes in the follicular levels of StAR protein and steroid production.
Pituitary LH and FSH are known to be the major regulators of ovarian function. In the last few years, however, there has been evidence that growth hormone (GH) is also involved in ovarian regulation. Therefore, the aim of our study was to elucidate the mechanisms of GH action during in vitro maturation (IVM) of bovine cumulus-oocyte complexes (COCs). As shown by detection of the nuclear cell proliferation-associated antigen Ki-67, COCs matured in vitro in the presence of GH revealed a significantly (P < 0.05) higher proportion of proliferating cumulus cells (12.6%) compared with the COCs matured in the control medium TCM 199 (9.9%). In contrast, the percentage of proliferating cells was not increased by supplementation of the medium with a combination of GH and insulin-like-growth factor I (IGF-I). Apoptosis as determined by TUNEL (terminal doxynucleotidyl transferase mediated dUTP nick-end labeling) was significantly (P < 0.05) reduced in the cumulus cells by GH treatment. COCs matured with a combination of GH and IGF-I revealed the lowest percentage of apoptotic cells (11%). The localization and quantification of the gap junction protein connexin 43 (Cx 43) demonstrated that GH induced a significant decrease in the synthesis of the Cx 43 protein in the cumulus cells. Our results imply that GH increases cumulus expansion by promotion of cell proliferation and inhibition of apoptosis. Whereas the increase in cell proliferation is a direct effect of GH, the antiapoptotic effects of GH during in vitro maturation are modulated by IGF-I. Stimulatory effects of GH on oocyte maturation are correlated with changes in the synthesis of gap junction proteins.
Hyperactivated sperm motility is usually characterized by high-amplitude flagellar bends and asymmetrical flagellar beating. There is evidence that an inositol 1,4,5-trisphosphate (IP3) receptor-gated Ca2 store in the base of the flagellum provides Ca2 to initiate hyperactivation; however, the identity of the store was not known. Ca2 stores are membrane-bounded organelles, and the only two membrane-bounded organelles found in this region of sperm are the redundant nuclear envelope (RNE) and mitochondria. Transmission electron micrographs revealed two different compartments of RNE, one enriched with nuclear pores and the other containing few pores but extensive membranous structures with enlarged cisternae. Immunolabeling showed that IP3 receptors and calreticulin are located in the region containing enlarged cisternae. In other cell types, mitochondria adjacent to Ca2 stores are actively involved in modulating Ca2 signals by taking up Ca2 released from stores and also may respond by increasing production of NADH and ATP to support increased energy demand. Nevertheless, bull sperm did not show an increase in NADH when Ca2 was released from intracellular stores by thapsigargin to induce hyperactivation. Consistently, no net increase in ATP production was detected when sperm were hyperactivated, although ATP was hydrolyzed at a greater rate. Furthermore, blocking Ca2 efflux from mitochondria by CGP-37157, a specific inhibitor of the mitochondrial Na/Ca2 exchanger, did not inhibit the development of hyperactivated motility. We concluded that the intracellular Ca2 store is the part of RNE that contains enlarged cisternae and that Ca2 is released directly to the axoneme to trigger hyperactivated motility without the active participation of mitochondria.
Earlier studies have implicated the significance of transforming growth factor-β3 (TGFβ3) in the regulation of Sertoli cell tight junction (TJ) dynamics, possibly via its inhibitory effects on the expression of occludin, claudin-11, and zonula occludens-1 (ZO-1). Yet the mechanism by which TGFβ3 regulates the Sertoli cell TJ-permeability barrier is not known. Using techniques of semiquantitative reverse transcription-PCR (RT-PCR), immunoblotting, immunohistochemistry, and inhibitors against different kinases coupled with physiological techniques to assess the Sertoli cell TJ barrier function, it was shown that this TGFβ3-induced effect on Sertoli cell TJ dynamics is mediated via the p38 mitogen-activated protein (MAP) kinase pathway. First, the assembly of the Sertoli cell-TJ barrier was shown to be associated with a transient but significant decline in both the TGFβ3 production and expression by Sertoli cells. Furthermore, addition of TGFβ3 to Sertoli cell cultures during TJ assembly indeed perturbed the TJ barrier with an IC50 at ∼9 pM. Second, the TGFβ3-induced disruption of the TJ barrier was associated with a transient induction in MEKK2 but not the other upstream signaling molecules that mediate TGFβ3 action, such as Smad2, Cdc42, Rac2, and N-Ras, suggesting this effect might be mediated via the p38 MAP kinase pathway. This postulate was confirmed by the observation that TGFβ3 also induced the protein level of the activated and phosphorylated form of p38 MAP kinase at the time the TJ barrier was perturbed. Third, and perhaps the most important of all, this TGFβ3-mediated inhibitory effect on the TJ barrier and the TGFβ3-induced p-p38 MAP kinase production could be blocked by SB202190, a specific p38 MAP kinase inhibitor, but not U0126, a specific MEK1/2 kinase inhibitor. These results thus unequivocally demonstrate that TGFβ3 utilizes the p38 MAP kinase pathway to regulate Sertoli cell TJ dynamics.
In avian species, a glycoprotein homologous to mammalian ZPC is synthesized in the granulosa cells of developing follicles. We have previously reported that the newly synthesized ZPC (proZPC) in granulosa cells is cleaved at a consensus furin cleavage site to generate mature ZPC prior to secretion. In the present study, we examined the effect of the proteolytic cleavage of proZPC on ZPC secretion by using a specific inhibitor of furin endoprotease and site-directed mutagenesis of the furin cleavage site. Western blot analysis demonstrated that the furin inhibitor efficiently blocked both the proteolytic cleavage of proZPC and the subsequent ZPC secretion. A site-directed mutant that possessed a mutated sequence for furin cleavage was not secreted from the cells. The immunocytochemical observations indicated that proZPC produced in the presence of a furin inhibitor or those produced by the site-directed mutant of the furin cleavage site had accumulated in the endoplasmic reticulum. These results indicate that proZPC is proteolytically cleaved at the consensus furin cleavage site with furin-like protease, and the failure of this cleavage results in its accumulation in the endoplasmic reticulum. Therefore, the C-terminal proteolytic processing of proZPC at the consensus furin cleavage site is a prerequisite event for quail ZPC secretion.
Germ cell nuclear factor (GCNF), an orphan receptor in the nuclear receptor superfamily, is expressed predominantly in developing germ cells in the adult mouse. Two Gcnf transcripts (7.4 and 2.1 kilobase [kb]) encoded by a single copy gene are expressed in the testis of several mammalian species. To identify features that regulate Gcnf expression, we characterized the structure and sequence of the mouse gene and its two transcripts and determined the expression profile of the GCNF protein during spermatogenesis. Genomic fragments spanning part of the 5′-untranslated region (UTR), the coding sequence, and the complete 3′-UTR (∼80 kb) were isolated and sequenced. The 3′-UTRs of the two transcripts are quite distinct. The 7.4 kb transcript, which appears earlier in spermatogenesis, has a very long 3′-UTR of 4451 nucleotides. In contrast, the 2.1 kb transcript, which is expressed predominantly during the haploid phase of spermatogenesis, has a 3′-UTR that is only 202 nucleotides in length. Additional analyses indicate that both transcripts share the same coding region and are associated with polysomes. A single GCNF protein band was detected in testis extracts by Western blotting with a specific antiserum. Immunohistochemical analysis showed that GCNF is localized in the nuclei of pachytene spermatocytes and round spermatids. GCNF is first detectable in early pachytene spermatocytes (stage II) and is continuously expressed until spermatids begin to elongate in stage IX. Although GCNF is generally distributed throughout the nucleus, it is particularly prominent in heterochromatic regions at some stages and in condensed chromosomes undergoing the meiotic divisions. This expression profile suggests that GCNF plays a role in transcriptional regulation during meiosis and the early haploid phase of spermatogenesis.
Recent evidence from our research suggested the direct role of ghrelin in the control of testicular function. However, the pattern of expression and hormonal regulation of the gene encoding its cognate receptor (i.e., the growth hormone-secretagogue receptor [GHS-R]) in the male gonad remains to be fully elucidated. In this paper, overall expression of GHS-R mRNA in rat testis was compared with that of the functional receptor form, namely GHS-R type 1a, in different developmental and experimental settings. In addition, cellular distribution of GHS-R within adult testis tissue was assessed. Our analyses demonstrated persistent expression of the GHS-R gene in rat testis throughout postnatal development. In contrast, testicular expression of GHS-R type 1a mRNA remained undetectable before puberty and sharply increased thereafter. In adult testis, GHS-R1a mRNA expression presented a scattered pattern of cellular distribution, including Sertoli and Leydig cells that also showed specific GHS-R1a immunoreactivity. Expression of total GHS-R and specific GHS-R1a mRNAs was detected in isolated seminiferous tubule preparations, with varying levels throughout the defined stages of the spermatogenic cycle. In addition, testicular expression of total GHS-R and GHS-R1a mRNAs was up-regulated by exposure to ghrelin in vitro and after stimulation with FSH in vivo. In conclusion, our data demonstrate that expression of the GHS-R gene in rat testis takes place in a developmental, stage-specific, and hormonally regulated manner. Divergent expression of total GHS-R and type 1a specific mRNAs was detected at certain stages of postnatal development and spermatogenic cycle, thus raising the possibility that, in addition to net changes in GHS-R gene expression, the balance between receptor subtypes may represent a novel mechanism for the tuning of ghrelin sensitivity in rat testis.
During mammalian spermatogenesis, the transcription of several genes in Sertoli cells is turned on and off as the adjacent male germ cells progress through the stages of the cycle of the seminiferous epithelium. A requirement for defining how germ cells regulate this process is the identification of a promoter that confers, in vivo, accurate, stage-specific gene expression in Sertoli cells. To date, such a promoter has not been identified. Using transgenic mice, we show that the 3-kilobase genomic fragment immediately upstream of the rat cathepsin L translation start site directs expression of the reporter gene, β-galactosidase, only in Sertoli cells. The expression pattern of the reporter gene recapitulated that of the endogenous gene in Sertoli cells as 75% of the seminiferous tubules that contained X-gal positive Sertoli cells were at stages VI–VIII and β-galactosidase enzymatic activity was 4-fold higher in mature testes compared with immature testes. This is, to our knowledge, the first identification of a promoter region that contains all of the regulatory elements required for accurate, stage-specific gene expression in Sertoli cells.
It has been proposed that stage-specific gene expression in Sertoli cells results from sequential activation and repression of transcription. However, the exact molecular mechanisms are unknown. As a first step in addressing this fundamental issue, we recently demonstrated that a 3-kilobase (kb) genomic fragment immediately upstream of the rat cathepsin L translation start site directed stage-specific expression of a reporter gene only in Sertoli cells of transgenic mice in a manner comparable to that of the endogenous gene (predominantly in stages VI–VIII tubules). Supporting the activation/repression model of regulation, an upstream domain that mediated an inhibitory effect by male germ cells was identified within this 3-kb promoter region. In the present study, we localized and characterized the regulatory elements that activate transcription. Analyses of a series of 5′ deletion constructs demonstrated that a 120-base pair (bp) region that spans the transcription start site of the rat cathepsin L gene was sufficient to activate transcription in Sertoli cells isolated from sexually mature rats. Within this region, electrophoretic mobility shift assays showed that one member of the Sp/XKLF family of factors, Sp3, specifically bound to a GC-box. Furthermore, Sp1-binding activity was not detected in nuclear extracts from Sertoli cells of sexually mature rats. Finally, the GC-box was shown to be essential for promoter activity since mutating this binding motif abolished promoter activity. Collectively, these results suggest that the GC-box is a critical regulatory element for the cathepsin L promoter in mature Sertoli cells.
In the avian species, germline chimera production could be possible by transfer of donor germ cells into the blood vessel of recipient embryos. This study was conducted to establish an efficient transfer system of chicken gonadal primordial germ cells (gPGCs) for producing the chimeras having a high capacity of germline transmission. Gonadal PGCs retrieved from 5.5-day-old embryos (stage 28) of Korean Ogol chicken (KOC with i/i gene) were transferred into the dorsal aorta of 2.5-day-old embryos (stage 17) of White Leghorn chicken (WL with I/I gene). Prospective evaluations of whether culture duration (0, 5, or 10 days) and subsequent Ficoll separation of gPGCs before transfer affected chimera production and germline transmission in the chimeras were made while retrospective analysis was conducted for examining the effect of chimera sexuality. A testcross analysis by artificial insemination of presumptive chimeras with adult KOC was performed for evaluating each treatment effect. First, comparison was made for evaluating whether experimental treatments could improve chimera production, but none of the treatments were significantly (P = 0.6831) influenced (5.1%–14.4%). Second, it was determined whether each treatment could enhance germline transmission in produced chimeras. More (P < 0.0001) progenies with black feathers (i/i) were produced in the germline chimeras derived from the transfer of 10-day-cultured gPGCs than from the transfer of 0- or 5-day-cultured gPGCs (0.6%–7.8% vs. 10.7%–49.7%). Ficoll separation was negatively affected (P < 0.0001), whereas there was no effect in chimera sexuality (P = 0.6011). In conclusion, improved germline transmission of more than a 45% transmission rate was found in chicken chimeras produced by transfer of 10-day-cultured gPGCs being separated without Ficoll treatment.
We previously reported the presence of vascular endothelial growth factor (VEGF) in testicular cells, and high concentrations of VEGF have been measured in semen, although its role in male reproduction remains obscure. In the present study we focus on understanding the mechanism of VEGF production by mouse Leydig cells cultured in vitro. Production of VEGF protein in medium by testicular cells was markedly increased by the addition of hCG in a time- and dose-dependent manner. Gonadotropin-stimulated VEGF production was mediated by cAMP-dependent protein kinase A (PKA), as evidenced by the effect of hCG being mimicked by 8Br-cAMP and being abolished in the presence of a PKA-specific inhibitor, H-89. Protein kinase C was not involved, as evidenced by phorbol 12-myristate 13-acetate having no influence on VEGF production by Leydig cells. In addition to hCG, atrial natriuretic peptide was also able to stimulate VEGF production, suggesting that cGMP is able to cross-activate PKA. A specific Src kinase inhibitor, PP2, could completely block the stimulatory effects of both gonadotropin and 8Br-cAMP on VEGF production by Leydig cells, implying an involvement of the Src kinase pathway. Furthermore, addition of U0126, an inhibitor of MEK 1/2, abolished the increase in VEGF production stimulated by both hCG and 8Br-cAMP. A similar inhibitory effect was observed by the addition of SB203580, a p38 mitogen-activated protein kinase inhibitor. Thus, in conclusion, Leydig cells are able to produce VEGF by a process under gonadotropic control, and PKA plays a key role in this process. Downstream of PKA, it appears that both MEK 1/2 and Src kinase-dependent pathways are involved, although further research will be necessary to determine the precise link between PKA and other kinases involved.
The objective of this study was to determine whether nitric oxide (NO) is produced locally in the bovine corpus luteum (CL) and whether NO mediates prostaglandin F2α (PGF2α)-induced regression of the bovine CL in vivo. The local production of NO was determined in early I, early II, mid, late, and regressed stages of CL by determining NADPH-d activity and the presence of inducible and endothelial NO synthase immunolabeling. To determine whether inhibition of NO production counteracts the PGF2α-induced regression of the CL, saline (10 ml/h; n = 10) or a nonselective NOS inhibitor (Nω-nitro-l-arginine methyl ester dihydrochloride [L-NAME]; 400 mg/h; n = 9) was infused for 2 h on Day 15 of the estrous cycle into the aorta abdominalis of Holstein/Polish Black and White heifers. After 30 min of infusion, saline or cloprostenol, an analogue of PGF2α (aPGF2α; 100 μg) was injected into the aorta abdominalis of animals infused with saline or L-NAME. NADPH-diaphorase activity was present in bovine CL, with the highest activity at mid and late luteal stages (P < 0.05). Inducible and endothelial NO synthases were observed with the strongest immunolabeling in the late CL (P < 0.05). Injection of aPGF2α increased nitrite/nitrate concentrations (P < 0.01) and inhibited P4 secretion (P < 0.05) in heifers that were infused with saline. Infusion of L-NAME stimulated P4 secretion (P < 0.05) and concomitantly inhibited plasma concentrations of nitrite/nitrate (P < 0.05). Concentrations of P4 in heifers infused with L-NAME and injected with aPGF2α were higher (P < 0.05) than in animals injected only with aPGF2α. The PGF2α analogue shortened the cycle length compared with that of saline (17.5 ± 0.22 days vs. 21.5 ± 0.65 days P < 0.05). L-NAME blocked the luteolytic action of the aPGF2α (22.6 ± 1.07 days vs. 17.5 ± 0.22 days, P < 0.05). These results suggest that NO is produced in the bovine CL. NO inhibits luteal steroidogenesis and it may be one of the components of an autocrine/paracrine luteolytic cascade induced by PGF2α.
The objective of this study was to improve the conditions for oocyte development in vitro beginning with the primordial follicles of newborn mice. Previous studies showed that oocytes competent of meiotic maturation, fertilization, and preimplantation could develop in vitro from primordial follicles. However, the success rates were low and only one live offspring was produced (0.5% of embryos transferred). A revised protocol was compared with the original protocol using oocyte maturation and preimplantation development as end points. The percentage of oocytes maturing to metaphase II and developing to the blastocyst stage was significantly improved using the revised protocol. In addition, we compared the production of offspring from two-cell stage embryos derived from in vitro-grown and in vivo-grown oocytes. Of 1160 transferred two-cell stage embryos derived from in vitro-grown oocytes, 66 (5.7%) developed to term and 7 pups (10.6%) died at birth. The remaining 59 pups (27 females, 32 males) survived to adulthood. By comparison, of 437 transferred two-cell stage embryos derived from in vivo-grown oocytes, 76 (17.4%) developed to term and 4 (5.3%) died at birth. The remaining 72 pups (35 females, 37 males) survived to adulthood. These studies provide proof of the principle that fully competent mammalian oocytes can develop in vitro from primordial follicles and present a significant advance in oocyte culture technology.
Christian Tessier, Anne Prigent-Tessier, Lei Bao, Carlos M. Telleria, Susan Ferguson-Gottschall, Gil B. Gibori, Yan Gu, Jennifer M. Bowen-Shauver, Nelson D. Horseman, Geula Gibori
Successful pregnancy requires profound differentiation and reorganization of the uterine tissues including, as pregnancy progresses, extensive apoptosis of decidual tissue to accommodate the developing conceptus. We have previously shown a positive correlation between expression of activin A and apoptosis in the decidua and have also shown that expression of activin A occurs at the time when prolactin (PRL) receptors disappear from decidual cells. The goals of this study were to examine whether activin A plays a role in decidual apoptosis and whether expression of activin A in the decidua is regulated by PRL and placental lactogens. Studies were carried out using primary rat decidual cells, a decidual cell line (GG-AD), and PRL null mice. Treatment of decidual cells with activin A significantly increased DNA degradation, caspase 3 activity, and caspase 3 mRNA expression. However, this effect was observed only in the absence of endogenous activin production by these cells. Addition of follistatin to decidual cells that were producing activin A decreased both caspase 3 activity and mRNA expression. Similarly, addition of activin-blocking antibodies to cultures of GG-AD cells, which also produce activin A, caused a reduction in both DNA degradation and caspase 3 activity. PRL and placental lactogens caused an inhibition of activin A mRNA expression in primary decidual cells. Even more convincingly, decidua of PRL null mice expressed abundant activin A at a time when no expression of this hormone is detected in wild-type mice and treatment of PRL null mice with PRL caused a profound inhibition of activin A mRNA expression. In summary, our investigations into the role and regulation of decidual activin have revealed that activin A can induce cell death in the decidua and that its expression is under tight regulation by PRL and placental lactogens.
A link between intrauterine infection and premature labor is widely accepted, yet the fetal inflammatory responses to such infections are not well understood. Our aim was to use a sheep model in which an inflammatory state was induced by lipopolysaccharide (LPS) administration during pregnancy to the maternal systemic, intra-amniotic or extra-amniotic compartments. Fetal and maternal blood gases and uterine electromyographic activity along with fetal and maternal circulating concentrations of prostaglandins PGE2 and PGFM, cortisol, and interleukin-6 were determined. Maternal systemic LPS treatment resulted in mild maternal hypoxemia, a rise in temperature, greater fetal hypoxemia, and a marked rise in fetal cortisol and PGE2 concentrations that persisted for 48 h. Intra-amniotic administration of LPS at doses higher than those used systemically caused an increase in fetal cortisol and PGE2 concentrations as well as a rise in uterine activity, but these were lesser in magnitude. Extra-amniotic LPS administration caused no overt fetal or maternal inflammatory responses. We conclude that maternal LPS treatment markedly elevated fetal cortisol and PGE2 concentrations. This may be a potential protective mechanism that aids the fetus in the event of premature delivery. The attenuated fetal response to intra-amniotic LPS treatment, despite the much higher dose used, may support a role for the amniotic fluid in protecting the fetus from endotoxin exposure during pregnancy.
Major histocompatibility complex (MHC) class I molecules, consisting of an α chain and β2-microglobulin (β2MG), play an important role in immune rejection responses by discriminating self and nonself and are increased by type I interferons during antiviral responses. Interferon tau (IFNτ), the pregnancy-recognition signal in ruminants, is a type I interferon produced by the ovine conceptus between Days 11 and 21 of gestation. In study 1, expression of MHC class I α chain and β2MG mRNA and protein was detected primarily in endometrial luminal epithelium (LE) and glandular epithelium (GE) on Days 10 and 12 of the estrous cycle and pregnancy. On Days 14–20 of pregnancy, MHC class I and β2MG expression increased only in endometrial stroma and GE and, concurrently, was absent in LE and superficial ductal GE (sGE). Although neither MHC class I nor β2MG proteins were detected in Day 20 trophectoderm, β2MG mRNA was detected in conceptus trophectoderm. In study 2, cyclic ewes were ovariectomized on Day 5, treated daily with progesterone to Day 16, received intrauterine infusions between Days 11 and 16 of either control serum proteins or recombinant ovine IFNτ, and were hysterectomized on Day 17. The IFNτ increased MHC class I and β2MG expression only in endometrial stroma and GE. During pregnancy, MHC class I and β2MG gene expression is inhibited in endometrial LE and sGE but, paradoxically, is stimulated by IFNτ in the stroma and GE. The silencing of MHC class I α chain and β2MG genes in the endometrial LE and sGE during pregnancy recognition and establishment may be a critical mechanism preventing immune rejection of the conceptus allograft.
The pubertal initiation of spermatogenesis is reliant on androgens, and during this time, 5α-reduced androgens such as dihydrotestosterone (DHT) are the predominant androgens in the testis. Two 5α-reductase (5αR) isoenzymes (5αR1 and 5αR2) have been identified, which catalyze the conversion of testosterone to the more potent androgen DHT. The present study aimed to investigate the developmental pattern of 5αR isoenzymes and their relationship to the production of 5α-reduced androgens in the postnatal rat testis. Both 5αR1 and 5αR2 isoenzyme mRNAs were measured by real-time polymerase chain reaction, isoenzyme activity levels by specific assays, and testicular androgens by radioimmunoassay after high-performance liquid chromatographic separation. Both 5αR1 and 5αR2 mRNAs and activity levels were low in the 10-day-old (prepubertal) testis, peaked between Days 20 and 40 during puberty, and then declined to low levels at 60–160 days of age. The developmental pattern of both 5αR isoenzyme activity levels was mirrored by the testicular production of 5α-reduced metabolites. Although 5αR1 was greater than 5αR2 at all ages, it is likely, given the substrate preferences of the two, that both isoenzymes contribute to the pubertal peak of 5α-reduced androgen biosynthesis. The peak in 5αR isoenzymes and 5α-reduced metabolite production coincided with the first wave of spermatogenesis in the rat, suggesting a role for 5α-reduced metabolites in the initiation of spermatogenesis. This was explored by acute administration of a 5αR inhibitor (L685,273) to immature rats. The L685,273 markedly suppressed testicular 5αR activity during puberty by 75%–86%. However, a marked increase was observed in testicular testosterone levels (in the absence of changes in LH), and no decrease was observed in the absolute levels of 5α-reduced metabolites. Therefore, whether the formation of DHT in the presence of low testosterone levels in the pubertal testis is required for the initiation of spermatogenesis cannot be tested using 5αR inhibitors. We conclude that both 5αR1 and 5αR2 isoenzymes are involved in the peak of 5α-reduced androgen biosynthesis in the testis during the pubertal initiation of spermatogenesis.
Members of the transforming growth factor (TGF) β family play critical roles in regulating placental functions. Using polymerase chain reaction (PCR)-based strategies, we have cloned four transcripts encoding full-length activin receptor-like kinase 7 (ALK7) and three novel ALK7 isoforms from the human placenta. The full-length ALK7 has 493 amino acids and exhibits all characteristics of TGFβ type I receptors, including an activin receptor-binding domain, a transmembrane domain, a GS domain, and a serine/threonine kinase domain. The three ALK7 isoforms identified include a truncated ALK7 (tALK7) and two soluble proteins designated as soluble ALK7a (sALK7a) and soluble ALK7b (sALK7b). The tALK7 lacks the first 50 amino acids of the full-length ALK7, resulting in a truncated receptor-binding domain. Both sALK7a and sALK7b lack transmembrane and GS domains. The ALK7 gene, located on chromosome 2q24.1, is composed of at least nine exons and eight introns. The isoforms of ALK7 are generated by alternative splicing. Transcripts encoding the sALK7 isoforms differ from the full-length transcript by lacking exon III or both exons III and IV in sALK7a and sALK7b, respectively. The transcript for tALK7 uses an alternative exon located within the first intron of the full-length transcript. These results indicate that four distinct proteins are encoded by the human ALK7 gene. Both reverse transcription-PCR and Western blot analysis showed that ALK7 and its isoforms are expressed in human placentae of different stages of pregnancy and that their expression is developmentally regulated. In addition, mRNA expression of Nodal, a ligand for ALK7, was also detected in placentae of different gestational age. The role of Nodal and ALK7 in human placenta is currently under investigation.
Embryonic stem (ES) cells are self-renewing, pluripotent, and capable of differentiating into all of the cell types found in the adult body. Therefore, they have the potential to replace degenerated or damaged cells, including those in the central nervous system. For ES cell-based therapy to become a clinical reality, translational research involving nonhuman primates is essential. Here, we report monkey ES cell differentiation into embryoid bodies (EBs), neural progenitor cells (NPCs), and committed neural phenotypes. The ES cells were aggregated in hanging drops to form EBs. The EBs were then plated onto adhesive surfaces in a serum-free medium to form NPCs and expanded in serum-free medium containing fibroblast growth factor (FGF)-2 before neural differentiation was induced. Cells were characterized at each step by immunocytochemistry for the presence of specific markers. The majority of cells in complex/cystic EBs expressed antigens (α-fetal protein, cardiac troponin I, and vimentin) representative of all three embryonic germ layers. Greater than 70% of the expanded cell populations expressed antigenic markers (nestin and musashi1) for NPCs. After removal of FGF-2, approximately 70% of the NPCs differentiated into neuronal phenotypes expressing either microtubule-associated protein-2C (MAP2C) or neuronal nuclear antigen (NeuN), and approximately 28% differentiated into glial cell types expressing glial fibrillary acidic protein. Small populations of MAP2C/NeuN-positive cells also expressed tyrosine hydroxylase (∼4%) or choline acetyltransferase (∼13%). These results suggest that monkey ES cells spontaneously differentiate into cells of all three germ layers, can be induced and maintained as NPCs, and can be further differentiated into committed neural lineages, including putative neurons and glial cells.
The regulation of preantral follicle growth in mammals is poorly understood. The availability of an adequate vascular supply to provide endocrine and paracrine signals may be important during the early states of follicle growth as well as the later states of follicle selection and dominance. The objective of the present study was to investigate whether vascular endothelial growth factor (VEGF) plays a role in preantral follicular development in the rat ovary. Immature (age, 21 days) Sprague-Dawley rats were injected with 500 ng of VEGF in saline or 50 μg of diethylstilbestrol (DES) in oil under the bursa of one ovary. The contralateral ovary was injected with a corresponding volume of vehicle. Rats were killed 48 h later, and the ovaries were removed and analyzed histologically. Intrabursal administration of VEGF significantly increased the number of primary and small secondary, but not of large secondary, preantral follicles in the ovary, similar to the effect of DES (P < 0.05). The VEGF stimulated preantral follicle growth in a time- and dose-dependent manner. Subcutaneous DES administration increased the number of primary and secondary follicles, and both s.c. and intrabursal estrogen administration stimulated VEGF protein expression in the rat ovary. These data indicate that VEGF stimulates preantral follicular development in the rat ovary, is regulated by estrogen, and may be one of the factors that participate in the regulation of early follicle growth in the rat.
Mutations in the transcription factor SOX9 give rise to campomelic dysplasia, a syndrome characterized by skeletal abnormalities and XY sex reversal. Sox9 is expressed at sites of chondrogenesis and in the developing testis, and, thus, it plays a role in two overtly different pathways of differentiation. Previous studies have identified the gene for type II collagen, Col2a1, as a target of Sox9 in mouse chondrocytes and implicated Col9a3 as a Sox9 target in testis. Using differential expression analysis combined with reverse transcription-polymerase chain reaction and whole-mount in situ hybridization, we have identified nonchondrocytic collagen transcript isoforms that are expressed in the early male mouse gonad. Male-specific, gonadal expression of nonchondrocytic Col2a1 was first seen at 11.5 days postcoitum (dpc) and was undetectable by 13.5 dpc. This was accompanied by increasing expression of nonchondrocytic Col9a1, Col9a2, and Col9a3, first detected at 11.5 dpc. Expression was analyzed in testes that had been depleted of germ cells by the cytotoxic drug busulfan. These studies showed Col9a3 and Col2a1 to be expressed in Sertoli cells within the developing testis cords. Nonchondrocytic type II collagen contains a cysteine-rich domain that has been shown to bind members of the transforming growth factor β superfamily of signaling molecules. Thus, this interaction may play a role in the morphogenesis and differentiation of the testis.
Paracrine factors have significant effects during folliculogenesis. Because of various morphological features, the mare is a convenient model to study in vivo the effects of factors involved in periovulatory events. In the present work, epidermal growth factor (EGF; experiment 1, n = 49 mares) and interleukin-1β and interleukin-1RA (IL-1β and IL-1RA, respectively; experiment 2, n = 80 mares) were injected intrafollicularly to evaluate the influence of these factors on in vivo maturation of equine preovulatory follicles. A transvaginal ultrasound-guided injection was performed when the diameter of the dominant follicle reached 30–34 mm. In experiment 1, the four experimental groups were 1) EGF group, intrafollicular (i.f.) injection of EGF (2 ml; 0.5 μg/ml) plus i.v. injection of physiological serum; 2) control group, no injection; 3) PBS group, i.f. injection of 2 ml of PBS plus i.v. injection of physiological serum; 4) crude equine gonadotropins (CEG) group, i.f. injection of PBS plus i.v. injection of CEG (20 mg). In experiment 2, groups 3 and 4 were the same as in experiment 1, but groups 1 and 2 were changed as follows: 1) IL-1β group, i.f. injection of IL-1β (2 ml; 0.5 μg/ml) plus i.v. injection of physiological serum; 2) IL-1RA group, i.f. injection of IL-1RA (2 ml; 0.5 μg/ml) plus i.v. injection of physiological serum. In each experiment, cumulus-oocyte complexes from dominant/injected follicles were collected by transvaginal ultrasound-guided aspiration 38 h after intrafollicular injection. Cumulus morphology and oocyte nuclear stage were assessed. Additionally, in experiment 2, 40 mares were used to determine the time of ovulation after treatments. Our results indicate that intrafollicular injection of EGF or PBS induced lower cumulus expansion and oocyte maturation rates compared with the CEG group (P < 0.05). In experiment 2, the IL-1β and CEG groups showed the same expansion rate, the same oocyte maturation rate, and the same ovulation distribution. On the other hand, the intrafollicular injection of IL-1RA, as PBS, did not induce follicle and cumulus-oocyte complex (COC) maturation. In conclusion, we confirmed that the technique of intrafollicular injection can be used in the mare to study the role of specific molecules. We demonstrated for the first time in mares that the injection of EGF did not influence in vivo COC maturation. In contrast, IL-1β injection into the dominant follicle induced in vivo oocyte maturation and the ovulation process whereas IL-1RA seemed to block these mechanisms.
Spermatogenic immunoglobulin superfamily (SgIGSF) is a mouse protein belonging to the immunoglobulin superfamily expressed in the spermatogenic cells of seminiferous tubules. We produced a specific polyclonal antibody against SgIGSF. Western blot analysis of the testes from postnatal developing mice using this antibody demonstrated multiple immunopositive bands of 80–130 kDa, which increased in number and size with the postnatal age. Enzymatic N-glycolysis caused reduction in the size of these bands to 70 kDa, indicating that SgIGSF is a glycoprotein and its glycosylation pattern and extent are developmentally regulated. Immunohistochemical analysis of the adult testis demonstrated that SgIGSF was present in the spermatogenic cells in the earlier steps of spermatogenesis and increased in amount from intermediate spermatogonia through zygotene spermatocytes but was diminished in the steps from early pachytene spermatocytes through round spermatids. After meiosis, SgIGSF reappeared in step 7 spermatids and was present in the elongating spermatids until spermiation. The immunoreactivity was localized primarily on the cell membrane. Consistent with the findings in adult testes, the analysis of the developing testes revealed that SgIGSF was expressed separately in the spermatogenic cells in earlier and later phases. Sertoli cells had no expression of SgIGSF, whereas both SgIGSF immunoprecipitated from the testis lysate and produced in COS-7 cells was shown to bind to the surface of Sertoli cells in primary culture. These results suggested that SgIGSF on the surface of spermatogenic cells binds to some membrane molecules on Sertoli cells in a heterophilic manner and thereby may play diverse roles in the spermatogenesis.
Appropriate expression of the GnRH receptor (GnRH-R) in gonadotropes is critical for GnRH signaling and hence for gonadotropin secretion and sexual development. In the present work, we have studied the ontogeny of the steady-state GnRH-R mRNA levels in pituitaries of male rats from Day 5 to Day 55, when sexual maturity is attained. Developmental changes of gonadotropin subunit (α, FSHβ, and LHβ) mRNA levels were also assessed. In addition, the role of the endogenous GnRH on the maturational changes of GnRH-R and gonadotropin subunit gene expression was investigated. Messenger RNA levels were determined by Northern blot analysis of total RNA from anterior pituitaries. Amounts of the most abundant (5.0 kb) GnRH-R mRNA increased slowly from Day 5 through the infantile and the juvenile periods, to peak at Day 35 (12-fold increase vs. Day 5). Thereafter, the levels of the GnRH-R mRNA decline slightly until Day 55 (33% decrease vs. Day 35). Parallel changes were observed on the 4.5-kb transcript of the GnRH-R gene. Alpha subunit mRNA was easily detected at Day 5, and its levels increased progressively through the infantile period (2.5-fold increase) and peaked at Day 25 (3.3-fold increase vs. Day 5) with a smooth nonstatistically significant increment until Day 35; then it decreased by 41.5% at Day 55. FSHβ and LHβ mRNA levels rose slowly until Day 25. A sharp rise occurred thereafter to reach maximum levels at Day 35 (5.8-fold for FSHβ and 3.8-fold for LHβ vs. Day 25). Thereafter, the levels of both mRNAs fell until Day 55 (44.1% decrease for FSHβ and 37.1% decrease for LHβ vs. Day 35). To ascertain whether developmental activation of the GnRH-R and gonadotropin subunit gene expression is GnRH dependent, we have studied the effect of blocking the endogenous GnRH action by treating developing male rats with the specific GnRH antagonist cetrorelix (1.5 mg/kg body weight/week, s.c.) through the infantile (Days 5–20) and the juvenile periods (Days 20–35). Cetrorelix completely blocked the rise of levels of the two most abundant species, 5.0 kb and 4.5 kb, of the GnRH-R mRNA, during both the infantile and the juvenile periods. Cetrorelix also abolished the developmental rise of the gonadotropin β subunit mRNAs during the two periods of the study. In contrast, the α subunit gene expression was not altered by cetrorelix treatment during any of the two periods. These data demonstrate that sexual maturation of male rats is accompanied by a progressive and concerted induction of GnRH-R and gonadotropin subunit gene expression. Developmental activation of GnRH-R and gonadotropin β subunit genes is GnRH dependent. The apparent GnRH-independent regulation of the α-glycoprotein subunit mRNA levels may be due to the contribution of thyrotropes and perhaps to the presence of exclusive regulatory signals for this gene.
Differentiation of extravillous trophoblast cells (EVT) to an invasive phenotype plays an essential role in establishing and maintaining feto-placental organization during human pregnancy. A switch in integrin expression occurs during this differentiation and is accompanied by changes in the extracellular matrix (ECM). Alteration of EVT behavior is also modulated by cytokines. To investigate the molecular interactions involved in the EVT differentiation, we examined the effects of cytokines and ECM on the human EVT cell line, TCL1 cells. We found that tumor necrosis factor alpha (TNFα) induced apoptosis in TCL1 cells but not in JEG3 cells derived from choriocarcinoma while the addition of interleukin-1β, leukemia inhibitory factor, or transforming growth factor had no effect on TCL1 cells. This apoptosis was suppressed when TCL1 cells were seeded on fibronectin (Fn), collagen type I (C1), collagen type IV (C4), or laminin (Ln). Wortmannin, a specific PI3 kinase inhibitor, inhibited this suppression. Spreading assays and adhesion blocking assays indicated that TCL1 cells express integrin-α5 and -α6 and β1 and β4 subunits. Adhesion on Fn is mediated by α5β1, and adhesion on C1, C4, or Ln is mediated by α6β1 integrins. TNFα suppressed α6 integrin expression and enhanced α1 integrin expression in a dose-dependent manner. In addition, aggregation of β1 subunits on C4 was detected after addition of TNFα. Taken together, these results suggest that TNFα and ECM, through activation of PI3 kinase mediated by β1 integrin signaling, might collaboratively regulate differentiation of trophoblast cells through integrin signaling in establishing and maintaining successful pregnancy.
Long-term preservation of mouse sperm by desiccation is economically and logistically attractive. The current investigation is a feasibility study of the preservation of mouse sperm by convective drying in an inert gas (nitrogen). Mouse sperm from the B6D2F1 strain isolated in an EGTA-supplemented Tris-HCl buffer were dried using three different drying rates and were stored for 18–24 h at 4°C. The mean final moisture content was <5% for all the protocols. After intracytoplasmic sperm injection (ICSI), the mean blastocyst formation rates were 64%, 58%, and 35% using the rapid-, moderate-, and slow-drying protocols, respectively. The slow-drying protocol resulted in a rate of development significantly lower than that observed using rapid- and moderate-drying protocols and indicated that a slower drying rate may be detrimental to the DNA integrity of mouse sperm. The transfer of 85 two- or four-cell embryos that were produced using rapidly desiccated sperm resulted in 11 fetuses (13%) on Day 15 compared with the production of 34 fetuses (40%) produced using the transfer of 86 two- or four-cell embryos that were produced using fresh sperm (P < 0.05). The results demonstrate the feasibility of using a convective drying protocol for the successful desiccation of mouse sperm and identifies some of the important parameters required for optimization of the procedure.
More than 300 separated actions have been attributed to prolactin (PRL), which could be correlated to the quasi-ubiquitous distribution of its receptor. Null mutation of the PRL receptor (PRLR) gene leads to female sterility caused by a failure of embryo implantation. Using the PRLR knockout mouse model and the mRNA differential display method, among 45 isolated genes, we identified UA 4 as a PRL and steroids-target gene during the peri-implantation period that encodes the decysin. Hormonally regulated in the uterus during pregnancy, this new member of disintegrin metalloproteinase is present in the uterus at the site of blastocyst apposition in nondifferentiated stromal cells at the antimesometrial pole and, interestingly, is colocalized with the PRLR. At midpregnancy, decysin expression persists specifically at the foeto-maternal junction around vessels. Although it has been previously suggested that decysin expression is related to immune function, its function during pregnancy remains to be clearly established.
Ubiquitin-dependent proteolysis has been implicated in the recognition and selective elimination of paternal mitochondria and mitochondrial DNA (mtDNA) after fertilization in mammals. Initial evidence suggests that this process is contributed to by lysosomal degradation of the ubiquitinated sperm mitochondrial membrane proteins. The present study examined the role of the proteasome-dependent protein degradation pathway of the ubiquitin system, as opposed to lysosomal proteolysis of the ubiquitinated proteins, in the regulation of sperm mitochondrion elimination after fertilization. Boar spermatozoa prelabeled with vital fluorescent mitochondrial probes MitoTracker were used to trace the degradation of paternal mitochondria after in vitro fertilization (IVF) of porcine oocytes. The degradation of sperm mitochondria in the cytoplasm of fertilized oocytes started very rapidly, i.e., within 12–20 h after insemination. Four stages of paternal mitochondrial degradation were distinguished, ranging from an intact mitochondrial sheath (type 1) to complete degradation (type 4). At 27–30 h postinsemination, 96% of zygotes contained the partially (type 3) or completely (type 4) degraded sperm mitochondria. Highly specific peptide inhibitors of the ubiquitin-proteasome pathway, lactacystin (10 and 100 μM) and MG132 (10 μM), efficiently blocked the degradation of the sperm mitochondria inside the fertilized egg when applied 6 h after insemination. Using 10 μM MG132, only 13.6% of fertilized oocytes screened 27–30 h after IVF displayed type 3 sperm mitochondria, and there was no incidence of type 4, completely degraded mitochondria. Although lactacystin is not a reversible agent, the effect of MG132 was fully reversible: zygotes transferred to regular culture medium after 24 h of culture with 10 μM MG132 resumed development and degraded sperm mitochondria within the next cell cycle. Surprisingly, penetration of the zona pellucida (ZP) was also inhibited by MG-132 and lactacystin when the inhibitors were added at insemination. Altogether, these data provide the first evidence of the participation of proteasomes in the control of mammalian mitochondrial inheritance and suggest a new role of the ubiquitin-proteasome pathway in mammalian fertilization.
Spermatogenesis is dependent on a small population of stem cells. Although stem cells are believed to expand infinitely, there is little functional evidence regarding whether spermatogonial stem cells can increase in their number. Using the spermatogonial transplantation technique, we evaluated the proliferative potential of spermatogonial stem cells in two models of regeneration. After busulfan injection to deplete stem cells, the surviving stem cells were able to expand by at least 15.8-fold within 2 mo. On the other hand, a serial transplantation study indicated that one transplanted stem cell was able to expand by 3.8- and 12-fold within 2 and 4 mo, respectively. These results provide direct functional evidence for the expansion of stem cells and establish the basis for further characterization of the stem cell self-renewal process.
Previous investigations have shown that maternal diabetes impairs rodent embryo development during the earliest phase of gestation. Exposure to high concentrations of glucose before implantation results in a decrease in the number of cells per embryo and in a concomitant increase in two nuclear markers of apoptosis: chromatin degradation and nuclear fragmentation. In the present study, we show that caspase-6 is expressed in rat blastocysts, using reverse transcription-polymerase chain reaction (RT-PCR) and immunocytochemistry. Caspase-6 is detected in all cells of the blastocyst and is excluded from the nucleus. To assess the role of caspase-6 in the glucose-induced apoptosis, rat blastocysts were incubated for 24 h in either 6 or 28 mM glucose in the presence or absence of a specific inhibitor of caspase-6 (VEID-CHO, 100 nM). After incubation, blastocysts were examined for the proportion of nuclei showing signs of chromatin degradation and nuclear fragmentation. Addition of VEID-CHO was found to inhibit nuclear fragmentation, but did not prevent the increase in chromatin degradation triggered by excess glucose. Our data indicate that chromatin degradation and nuclear fragmentation are two nuclear damages that are induced separately by high glucose in rat blastocysts. Furthermore, nuclear fragmentation in rat blastocysts is apparently mediated by the activation of caspase-6.
We recently reported an unusual abundance of arginine (4–6 mM) in porcine allantoic fluid during early gestation. However, it is not known whether such high concentrations of arginine are unique for porcine allantoic fluid or whether they represent an important physiological phenomenon for mammals. The present study was conducted to test the hypothesis that arginine is also the most abundant amino acid in ovine allantoic fluid. Allantoic and amniotic fluids, as well as fetal and maternal plasma samples, were obtained from ewes between Days 30 and 140 of gestation. Glycine was the most abundant amino acid in maternal uterine arterial plasma, representing approximately 25% of total α-amino acids. Alanine, glutamine, glycine, plus serine contributed approximately 50% of total α-amino acids in fetal plasma. Fetal:maternal plasma ratios for amino acids varied greatly, being less than 1 for glutamate during late gestation, 1.5–3 for most amino acids throughout gestation, and greater than 10 for serine during late gestation. Marked changes were observed in amino acid concentrations in amniotic and allantoic fluids associated with conceptus development. Concentrations of alanine, citrulline, and glutamine in allantoic fluid increased by 20-, 34-, and 18-fold, respectively, between Days 30 and 60 of gestation and were 24.7, 9.7, and 23.5 mM, respectively, on Day 60 of gestation (compared with 0.8 mM arginine). Remarkably, alanine, citrulline, plus glutamine accounted for approximately 80% of total α-amino acids in allantoic fluid during early gestation. Serine (16.5 mM) contributed approximately 60% of total α-amino acids in allantoic fluid on Day 140 of gestation. These novel findings of the unusual abundance of traditionally classified nonessential amino acids in allantoic fluid raise important questions regarding their roles in ovine conceptus development.
The mastomys is a small laboratory rodent that is native to Africa. Although it has been used for research concerning reproductive biology, in vitro fertilization (IVF) and intracytoplasmic sperm injection are very difficult in mastomys because of technical problems, such as inadequate sperm capacitation and large sperm heads. The present study was undertaken to examine whether mastomys spermatids could be used to fertilize oocytes in vitro using a microinsemination technique, because spermatids are more easily injected than mature spermatozoa into oocytes. Most mastomys oocytes (80%–90%) survived intracytoplasmic injection with either round or elongated spermatids. Round spermatids had little oocyte-activating capacity, similar to those of mice and rats, and exogenous stimuli were needed for normal fertilization. Treatment with an electric pulse in the presence of 50 μM Ca2 followed by culture in 10 mM SrCl2 led to successful oocyte activation. After injection of round spermatids into preactivated oocytes, 93% of oocytes were normally fertilized (male and female pronuclei formed), and 100% of cultured oocytes developed to the 2-cell stage. However, none reached term after transfer into recipient females. Elongated spermatids, which correspond to steps 9–11 in rats, activated oocytes on injection without additional activation treatment. After embryo transfer, five offspring (6% per transfer) developed to term. These results indicate that microinsemination with spermatids is a feasible alternative in animal species that are refractory to IVF and sperm injection and that using later-stage spermatids may lead to increased production of viable embryos that can develop into normal offspring.
Simultaneously evaluating postthaw viability and acrosome integrity of spermatozoa by flow cytometry would provide a valuable testing tool in both research and routine work. In the present study, a new triple-stain combination was developed for the simultaneous evaluation of viability and acrosome integrity of bovine sperm processed in egg yolk-based extender by flow cytometer. SYBR-14 and propidium iodide (PI) enabled the discrimination of sperm cells from egg yolk and debris particles, which was instrumental for the flow cytometric analyses of frozen-thawed bovine sperm, because it implied that washing steps to remove egg yolk were no longer required. In addition, phycoerythrin-conjugated peanut agglutinin (PE-PNA) was used to discriminate acrosome-damaged/reacted sperm cells from acrosome-intact cells. Repeatability was calculated using two processed ejaculates of 10 bulls. Three straws per batch were analyzed in duplicate measurements. Method-agreement analysis between the SYBR-14/PE-PNA/PI and fluorescein isothiocyanate (FITC)-conjugated PNA was performed, with FITC-PNA/PI staining being carried out on 14 frozen-thawed semen samples immediately after thawing and after a 3-h incubation at 37°C. The British Standards Institution repeatability index of the SYBR-14/PE-PNA/PI combination was 2.6%. On average, the FITC-PNA/PI method showed a 6.3% overestimation of the live and acrosome-intact sperm cell subpopulation. In conclusion, the new triple-stain combination is highly repeatable and easy to use in routine application, and it provides a more precise estimate for the rate of sperm cells with intact head membrane and acrosome compared to the generally used and validated FITC-PNA/PI staining.
Ana Maria S. Assreuy, Nylane M. N. Alencar, Benildo S. Cavada, Duílio R. Rocha-Filho, Regina F. G. Feitosa, Fernando Q. Cunha, Juan J. Calvete, Ronaldo A. Ribeiro
The complex of porcine seminal plasma heterodimers I and II (PSP-I/PSP-II), which are heterodimers of glycosylated spermadhesins, is the major component of porcine seminal fluid. The proinflammatory and immunostimulatory activities of this spermadhesin complex suggest its participation in modulation of the uterine immune activity that may ensure reproductive success. Spermadhesin PSP-I/PSP-II induced the migration of neutrophils into the peritoneal cavity of rats via activation of resident cells. In the present study, we have investigated the involvement of macrophages and mast cells in the neutrophil chemotactic activity of PSP-I/PSP-II and the underlying mechanism. Macrophages and mast cells were isolated, cultured, and stimulated with purified PSP-I/PSP-II. Pharmacological modulation was performed using the glucocorticoid dexamethasone, indomethacin (cyclooxygenase inhibitor), MK886 (leukotriene inhibitor), and the supernatant of spermadhesin-stimulated mast cells. Macrophages stimulated with PSP-I/PSP-II released into the culture supernatant a neutrophil chemotactic substance. This activity was partly inhibited by both dexamethasone (85%) and the supernatant of spermadhesin-stimulated mast cells (74%) but not by indomethacin and MK886. An anti-tumor necrosis factor (TNF) α antibody neutralized (by 68%) the neutrophil chemotactic activity of PSP-I/PSP-II-stimulated macrophages. An anti-interleukin (IL)-4 antibody blocked the inhibitory activity of spermadhesin-stimulated mast cells on release of a neutrophil chemotactic substance by PSP-I/PSP-II-stimulated macrophages. As a whole, these data indicate that the neutrophil migration-inducing ability of spermadhesin PSP-I/PSP-II involves the release of the inflammatory cytokine TNFα by stimulated macrophages and that this activity is modulated by the lymphokine IL-4 liberated by mast cells. The balance between these two cytokines may control onset of the local inflammatory reaction, avoiding excessive neutrophil recruitment that would lead to tissue damage.
We have previously reported increased protein expression of sarcoplasmic reticulum calcium ATPase (SERCA) 2b in myometrium from women in labor at term, but the stimulus for this change is unknown. Proinflammatory cytokines have been implicated in the cascade of events leading to preterm and term labor, and we hypothesize that interleukin (IL)-1β may induce changes in key calcium homeostatic mechanisms and, in turn, augment myometrial contractility before labor. The aim of the present study was to investigate the long-term effects of IL-1β on SERCA 2b protein expression, calcium mobilization from intracellular stores, and store-operated calcium entry. Myometrial biopsies were obtained, with consent, from women undergoing elective cesarean section at term. Primary-cultured human myometrial smooth muscle (HMSM) cells were exposed to IL-1β (10 ng/ml) for 24 h or to culture medium alone (control). Cells were subsequently used in Western blot studies or loaded with fura-2 to assess calcium dynamics using fluorescent digital imaging. The present study clearly demonstrated that IL-1β significantly increased SERCA 2b protein expression in HMSM cells. Cyclopiazonic acid-induced calcium transients were also augmented, predominantly by activation of lanthanum-sensitive, store-operated calcium entry. HMSM cell excitability was enhanced, as evidenced by increased basal calcium entry and the initiation of spontaneous calcium transients in 37% of IL-1β-treated cells. IL-1β modulation of calcium mobilization may be an important mechanism in the cascade of events preparing the pregnant uterus for labor.
Receptor-mediated endocytosis of yolk precursors via clathrin-coated structures is the key mechanism underlying rapid chicken oocyte growth. In defining oocyte-specific components of clathrin-mediated events, we have to date identified oocyte-specific yolk transport receptors, but little is known about the oocytes' supporting endocytic machinery. Important proteins implicated in clathrin-mediated endocytosis and recycling are the endophilins, which thus far have been studied primarily in synaptic vesicle formation; in the present study, as a different highly active endocytic system, we exploit rapidly growing chicken oocytes. Molecular characterization of the chicken endophilins I, II, and III revealed that their mammalian counterparts have been highly conserved. All chicken endophilins interact via their SH3 domain with the avian dynamin and synaptojanin homologues and, thus, share key functional properties of mammalian endophilins. The genes show different expression patterns: As in mammals, expression is low to undetectable in the liver and high in the brain; in ovarian follicles harboring oocytes that are rapidly growing via receptor-mediated endocytosis, levels of endophilins II and III, but not of endophilin I, are high. Immunohistochemical analysis of follicles demonstrated that endophilin II is mainly present in the theca interna but that endophilin III predominates within the oocyte proper. Moreover, in a chicken strain with impaired oocyte growth and absence of egg-laying because of a genetic defect in the receptor for yolk endocytosis, endophilin III is diminished in oocytes, whereas endophilin III levels in the brain and endophilin II localization to theca cells are unaltered. Thus, the present study reveals that the endophilins differentially contribute to oocyte endocytosis and development.
Tesmin is a testis-specific protein. Four mouse tesmin cDNAs so far reported encode a testis-specific, metallothionein-like, 30-kDa protein (tesmin-30). An antibody against tesmin-30, however, detected a protein of 60 kDa (tesmin-60) from the mouse testis. To resolve the relationship between the two, the immunoprecipitated native tesmin-60 was sequenced. The result indicated that tesmin-30 is not full-length but is part of the C-terminal half of tesmin-60. The full-length cDNA (2.2 kilobases [kb]) encoding tesmin-60 (475 amino acid residues) and its genomic DNA (23 kb) were cloned and sequenced. A search of databases indicated that tesmin is a member of the CXC-hinge-CXC family. Immunohistochemistry indicated that tesmin exhibits dynamic subcellular localization changes during spermatogenesis. Before meiosis, it was localized in the cytoplasm of early to late spermatocytes and then translocated into the nucleus just before meiotic division. After meiosis, it appeared in spermatids, starting from the acrosomal vesicles, moving to the nuclear membrane and then to the caudal end as the spermatids elongated, and finally relocating into the cytoplasm. Oxidative stress by cobalt chloride, as well as by diethylmaleate, induced both premature translocation of tesmin from the cytoplasm to the nucleus and apoptotic signals in spermatocytes. The persistent existence of tesmin and its temporally and spatially dynamic localization suggest that tesmin is involved in multiple stages of spermatogenesis and spermiogenesis, possibly during sperm maturation and/or morphogenesis.
In this study, we examined the localization of vascular endothelial growth factor (VEGF) and the changes in VEGF mRNA expression in various regions of the oviduct in fertile women throughout the ovulatory cycle. Oviduct tissue was collected from 22 women undergoing laparoscopic tubal sterilization or hysterectomy for a benign gynecological condition. Oviduct sections were divided into isthmus, ampullary, and infundibular regions. Serial cross sections were analyzed for the presence of VEGF by specific immunohistochemical staining. The mucosal layer was isolated, and a semiquantitative reverse transcription polymerase chain reaction was performed. Immunohistochemical study revealed VEGF in the oviduct luminal epithelium, smooth muscle cells, and blood vessels within the oviduct. VEGF mRNA expression in oviduct was the highest during the periovulatory stage, and the expression in the ampullary and infundibular regions was higher than that in the isthmus. There was a significant positive correlation between serum FSH and LH concentrations and VEGF mRNA expression. There was no significant correlation between serum estradiol and progesterone concentrations and VEGF mRNA expression. These results suggest that VEGF in human oviduct may play an important role related the early reproductive events, which occur predominantly in the ampulla during the periovulatory phase when serum FSH and LH concentrations are high.
In the pituitary, activin stimulates the synthesis and release of FSH. However, the activin receptor signaling pathways that mediate these effects are poorly known. We investigated these mechanisms in primary ovine pituitary cells (POP) and in the murine LβT2 gonadotrope cell line. POP cells and LβT2 cells express the different activin receptors (types IA, IB, IIA, and IIB) and the Smad proteins (Smad-2, -3, -4, and -7). In both POP and LβT2 cells, activin activated several signaling pathways: Smad-2, extracellular regulated kinase-1/2 (ERK1/2), p38, and phosphatidylinositol 3′-kinase (PI3K)/Akt. Phosphorylation of ERK1/2 and p38 were stimulated (3- to 6-fold) rapidly in 5 min, whereas activation of both Smad-2 and Akt (3- to 5-fold) occurred later, in 60 min. Activin also increased the association of activin receptor IIB with PI3K. Using specific inhibitors, we demonstrated that the activation of Smad-2 was partially blocked by the inhibition of PI3K but not by the inhibition of ERK1/2 or p38, suggesting a cross-talk between the Smad and PI3K/Akt pathways. In both POP and LβT2 cells, FSH expression and secretion in response to activin were not altered by the inhibition of PI3K/Akt, ERK1/2, or p38 pathways, whereas they were reduced by about 2-fold by expression of a dominant negative of Smad-2 or the natural inhibitory Smad-7 in LβT2 cells. These results indicate that activin activates several signaling pathways with different time courses in both POP and LβT2 cells, but only the Smad-2 pathway appears to be directly implicated in FSH expression and release in LβT2 cells.
Tissue factor pathway inhibitor 2 (TFPI-2), a Kunitz-type proteinase inhibitor, might play an important role during placenta growth by regulating trophoblast invasion and differentiation. Many TFPI-2 transcripts have been detected in syncytiotrophoblast cells, but conflicting results have been reported concerning TFPI-2 synthesis by the cytotrophoblast. To address this issue, we developed a method to isolate pure preparations of human villous cytotrophoblast cells from normal term placentas, and the synthesis of tissue factor, TFPI-1, and TFPI-2 mRNAs was then evaluated. Cells were isolated by trypsin-DNase-EDTA digestion, followed by Percoll gradient separation and immunodepletion of human leukocyte antigen-positive cells. The quality of villous cytotrophoblast cells was verified by electron microscopy. Purity of cell preparations was assessed by labeling cells with GB25, a monoclonal antibody specific to villous trophoblast cells, and by checking the absence of contaminating cells using anti-CD9 antibody. The lack of hCG, CD32 mRNA, and tissue factor mRNA also indicated the absence of contaminating cells. Using competitive reverse transcription polymerase chain reaction, we showed that freshly isolated villous cytotrophoblast cells synthesized significant levels of TFPI-1 mRNA and larger amounts of TFPI-2 mRNA. TFPI-1 and TFPI-2 mRNA synthesis remained unchanged when cytotrophoblast cells were cultured in complete medium and evolved as a multinucleated syncytiotrophoblast. These results indicate that the villous cytotrophoblast and syncytiotrophoblast are both important sites of TFPI-2 synthesis in the human placenta. This study also indicates that tissue factor detection should be used systematically to check the purity of cytotrophoblast cell preparations because it allows detection of contamination by monocytes/macrophages and by syncytial fragments.
An increase in metallothionein 1 (MT-1) mRNA was detected in the ovaries of immature Wistar rats that were primed with s.c. injection of 10 IU eCG followed 48 h later by 10 IU hCG s.c. to initiate the ovulatory process. Ovarian RNA was extracted at 0, 2, 4, 8, 12, 24, 72, 144, and 288 h after the primed animals were injected with hCG. These extracts were used for reverse transcription polymerase chain reaction (RT-PCR) differential display and Northern analyses that yielded complementary gene fragments for MT-1. Expression of MT-1 mRNA increased significantly by 24 h after hCG treatment and reached a peak at 144 h after hCG. In contrast, a disintegrin and metalloproteinase with thrombospondin motifs and a tissue inhibitor of metalloproteinase 1, which were also detected by the RT-PCR differential display procedure, reached a peak at 12 h after hCG and returned to control levels in the ovaries by 72 h after hCG. In situ hybridization indicated that most of the MT-1 mRNA was expressed in the vicinity of the theca interna of preovulatory follicles and in the lutein granulosa of postovulatory follicles. Thus, MT-1 mRNA expression is primarily in the vicinity of steroid-secreting areas of the ovary. The substantial increase in MT-1 mRNA expression might be important in protecting the ovarian tissues from oxidative stress generated by ovarian inflammatory events during the ovulatory process and luteinization.
The efficiency of intracytoplasmic sperm injection (ICSI)-mediated transgenesis is often limited by poor embryo development. Because our previous work indicated that impairment of embryo development is frequently related to chromosomal abnormalities, we hypothesized that foreign DNA and/or conditions used to enhance integration of the DNA might induce chromosome damage. Therefore, we examined the chromosomes of mouse embryos produced by transgenesis with the EGFP gene. Spermatozoa were processed with three methods that cause membrane disruption: freeze-thawing, Triton X-100, or Triton X-100 followed by a sucrose wash. Membrane-disrupted spermatozoa were mixed with EGFP plasmids and injected into metaphase II oocytes. Three endpoints were evaluated: paternal chromosomes of the zygote, embryo capacity to develop in vitro, and expression of the transgene at the morula/blastocyst stage. In all pretreatments, we observed a significant decrease (approximately 2-fold) in the frequency of normal karyoplates when spermatozoa were incubated with exogenous DNA as compared with the treatment when no DNA was added. As predicted, embryo development was correlated with the integrity of the paternal chromosomes of the zygote. Searching for the possible mechanism of chromosome degradation, we used the ion chelators EGTA and EDTA and found that they neutralize the harmful effect of the transgene and stabilize the paternal chromosomes. In the presence of chelating agents, however, the number of embryos expressing EGFP produced with ICSI-mediated transgenesis decreased significantly. The results suggest that treatment of spermatozoa with exogenous DNA leads to paternal chromosome degradation in the zygote. Furthermore, the mechanisms of disruption of paternal chromosomes and the integration of foreign DNA may be closely related.
In the adult ovary, pituitary FSH via interaction with its receptor (FSHR) is required for follicular maturation and granulosa cell development. In humans and nonhuman primates, the pool of follicles available for adult ovarian function is established in utero. However, our understanding of the ontogeny and developmental regulation of FSHR in the ovary of the primate fetus is incomplete. Our goal was to determine whether the baboon fetal ovary expresses the full-length FSHR mRNA transcript and whether levels are developmentally regulated. Fetal ovaries were obtained at mid (Day 100) and late (Day 165) gestation (term = Day 184) from untreated baboons and on Day 165 from baboons in which fetal estrogen levels were either decreased by >95% by treatment with the aromatase inhibitor CGS 20267 or restored to 30% of normal by treatment with CGS 20267 plus estradiol benzoate administered s.c. to the mother on Days 100–164. The full-length 2088-base pair FSHR mRNA transcript was expressed in ovaries of adult and fetal baboons untreated or treated with CGS 20267 or CGS 20267 and estrogen. Mean (±SEM) FSHR mRNA levels (ratio of FSHR mRNA:18S rRNA), quantified by reverse transcription polymerase chain reaction, were increased (P < 0.05) 2-fold between mid (0.34 ± 0.06) and late gestation (0.76 ± 0.07), an increase prevented (P < 0.05) in estrogen-depleted baboons (0.44 ± 0.10) and partially restored by treatment with CGS 20267 and estrogen (0.58 ± 0.16). We previously showed that the number of follicles/0.33 mm2 in fetal ovaries of untreated baboons in late gestation was reduced 50% by treatment with CGS 20267 and restored to normal in baboons treated with CGS 20267 and estrogen. Thus, when corrected for the number of follicles/0.33 mm2, FSHR mRNA levels were similar in baboon fetal ovaries untreated (0.010 ± 0.001) or treated with CGS 20267 (0.009 ± 0.002) or CGS 20267 and estrogen (0.007 ± 0.003). We conclude that estrogen plays a major role in regulating ovarian FSHR mRNA expression in the primate fetus, and that the developmental increase in FSHR mRNA levels reflects the estrogen-dependent increase in folliculogenesis (i.e., increased number of granulosa cells and oocytes).
To investigate alterations in the molecular weight forms of inhibin in bull testis from the infantile (4–5 wk of age) to postpubertal (49–56 wk of age) periods, testicular homogenates were obtained from animals of various ages and fractionated by a combination of immunoaffinity chromatography and SDS-PAGE. Subsequently, the fractions eluted from the SDS gels were assayed for total inhibin, inhibin A, and inhibin B by fluoroimmunoassay or immunofluorometric assays (IFMAs) and for inhibin bioactivity by an in vitro bioassay. The molecular mass patterns of inhibin A and inhibin B in the testis, as determined by the dimer-specific IFMAs, showed the presence of a peak of approximate 47 kDa until 21–26 wk of age. However, the peak disappeared after 31–32 wk of age. As bulls aged, especially after 31–32 wk of age, inhibin A and inhibin B levels increased in the molecular mass region of 27–34 kDa. Total inhibin showed two peaks, of between 20 and 26 kDa and at approximately 47 kDa, until 21–26 wk of age and a single peak between 20 and 30 kDa after 31–32 wk of age. The eluted fractions corresponding to 29, 31, or 47 kDa gave a dose-response curve that was parallel to the curve generated with 32-kDa inhibin A or 29-kDa inhibin B standard in the IFMA for inhibin A or inhibin B. The fractions corresponding to 29 and 31 kDa suppressed basal release of FSH from rat pituitary cells, but the 47-kDa fraction had a lower FSH-suppressing activity. In the testes of older bulls, immunoblot analysis revealed the presence of a 29-kDa band cross-reacting with inhibin α and inhibin βB antibodies and of a 31-kDa band cross-reacting with inhibin α and inhibin βA antibodies. The 47-kDa band was recognized by the α, βA, and βB antibodies. Immunohistochemisty of the testis at each age showed that inhibin α subunits were found exclusively in Sertoli cells, but the intensity of immunostaining diminished in older bulls, in parallel with the decrease in the testicular concentrations of total inhibin. We conclude that 1) bovine Sertoli cells produce both inhibin A and inhibin B, 2) inhibin production in Sertoli cells during the prepubertal period is characterized by the 47 kDa inhibin-related material that contains precursor forms of inhibin A and inhibin B, and 3) the proportion of the mature forms of inhibin A and inhibin B increases as bulls age, although total inhibin production in Setroli cells decreases.
Expression of vascular endothelial growth factor (VEGF) isoforms and its receptors, Flt-1 and KDR, was investigated during the period of peri-implantation in mink, a species that displays obligate embryonic diapause. Uterine samples were collected during diapause, embryo activation, and implantation from pseudopregnant and anestrous animals and analyzed by semiquantitative reverse transcription polymerase chain reaction and immunohistochemistry. The abundance of mRNA of VEGF isoforms 120, 164, and 188 was highest during late embryo activation and at implantation. VEGF protein was localized to the glandular epithelium at all stages of peri-implantation, whereas the luminal epithelium lacked VEGF reactivity during diapause. Endometrial stroma and luminal and glandular epithelia were positive for VEGF in implanted uteri. The invasive trophoblast cells of the implanting embryo were intensively stained. High levels of VEGF mRNA in pseudopregnant uteri indicates that VEGF upregulation leading to implantation is dependent upon maternal rather than embryonic factors. The abundance of the two receptors, KDR and Flt-1, increased in the uterus during implantation. Low levels of the receptors in pseudopregnant uteri compared with those containing activated or implanted embryos indicates that the embryo regulates receptor expression. These results demonstrate VEGF and VEGF receptor expression during early gestation in mink and suggest that maternal and embryonic input regulates different aspects of the angiogenic process.
After in vitro fertilization with spermatozoa from bulls with high in vitro fertility, a beneficial paternal effect is manifested during the G1 phase of the first cell cycle. This benefit determines an earlier onset of the first S phase, and then a successful morula-blastocyst transition 7 days later. We hypothesized that the origin of the paternal effect could be a shift of the metabolism of the fertilized oocyte, because in mice, sperm decondensation is responsible for a dramatic increase in glucose metabolism. In this study we investigated the interaction between both pronuclei and compared glycolysis and pentose phosphate pathway (PPP) activities in bovine oocytes fertilized with spermatozoa from bulls of high or low fertility. Here we demonstrate that male pronucleus formation is necessary for the onset of the S phase in the female pronucleus, and that the component promoting an early S phase in both pronuclei is metabolic and linked to an up-regulation of the PPP during the male pronucleus formation. This long-lasting paternal effect is more evidence of the important role of epigenetic control during early embryo development.
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