首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Atypical spermatogenesis in Murex brandaris occurs in specific areas of the testis. In the early stages, a large nucleus is apparent, together with highly electrodense drops of chromatin spread through the caryoplasm. Many mitochondria are observed at one point of the cytoplasm, and numerous centrioles appear. In the following stages, vacuolization and degeneration of the chromatin are apparent, producing half moon-shaped chromatinic bodies called "caryomerites." Some degeneration of the chondrioma occurs too. Later, the chromatin degenerates and the centrioles produce cilia, while the Golgi body, highly secretory at this stage, produces electrodense PAS- as well as Thiery-positive granules. Moreover, a vermiform-shaped anuclear ciliated cell with numerous electrodense cytoplasmic granules is observed in the M. brandaris atypical spermatozoon.  相似文献   

2.
During the process of chromatin cndensation in the spermiogenesis of the neogastropod mollusc Murex brandaris, the nuclear protein complement undergoes a complex series of changes. These changes lead to the appearance of three small protamines in the ripe sperm nuclei. We have characterized this system electrophoretically and at the compositions with antibodies elicited against a specific spermatozoan protamine. Our results indicate that the complex pattern of chromatin condensation during spermiogenesis in this species (M. brandaris) may be modulated by a series of post-translational (and intranuclear) modifications of DNA-interacting proteins, such as precursors to the sperm protamines. The amino acid composition of each sperm protamine is remarkably simple (lys + arg + gly ≥96 mol%). This system of spermiogenic/spermatozoal proteins in the neogastropod M. brandaris clearly differs from that in patellogastropods and archaeogastropods, and it may be helpful in understanding evolutionary changes in the chromatin condensation pattern during the spermiogenesis of gastropod molluscs. © 1994 Wiley-Liss, Inc.  相似文献   

3.
N Agell  M Chiva  C Mezquita 《FEBS letters》1983,155(2):209-212
Electrophoretic analysis of acid-soluble chromosomal proteins isolated from rooster testis cell nuclei at different stages of spermatogenesis, revealed that the nuclear content of a protein identified by its solubility, electrophoretic mobility and amino acid analysis as the protein conjugate histone H2A-ubiquitin (uH2A, A24) changed markedly from meiotic cells to late spermatids. The protein was not detectable in tetraploid primary spermatocytes; it was present in 1.7% of the total amount of nucleosomal core histones in early spermatids and reached its maximum level (3.5% and 11%) at the end of spermiogenesis, when histones are replaced by the protamine galline.  相似文献   

4.
Follitropin receptor (FSHR) in testicular Sertoli cells mediates signaling by pituitary follitropin (FSH) promoting intercellular communication with germ cells for normal spermatogenesis. Using receptor knockout mice we examined changes in sperm nucleoproteins and chromatin architecture. The expressions of transition proteins 1/2 (TP1/2) and protamine-2 (PRM-2) were greatly diminished at 21 days, but returned to normal at 35 days and 3 months after birth. However, protein components in chromatin were quite different. Western blots detected a reduction in PRM1/2 and prolonged retention of mono-ubiquitinated histone 2A (uH2A) in the epididymal sperm from adult mutants. Two forms of mono- and poly-uH2A were present in sonication-resistant testicular spermatids in normal mice, whereas only an elevated mono-uH2A was detectable in mutants. Decrease in PRM1/2 and retention of mono-uH2A was coincident with reduction in TP1/2 in premature spermatids. Thus lack of FSHR signaling impairs expression of TP1/2 and PRM-2 at an early stage of post-natal development causing delayed spermatogenesis. In the adult, absence of FSHR signaling prolongs retention of mono-uH2A, leading to impair transition of basic nucleoproteins and chromatin remodeling during mouse spermatogenesis.  相似文献   

5.
Summary The structure of guppy (Poecilia reticulata) spermatogonia and spermatocytes has been studied using electron microscopy. The spermatogonia, situated at the apex of the seminiferous tubule, are almost all surrounded by a network of Sertoli cells; they have very diffuse chromatin and one or two large nucleoli. The cytoplasm contains relatively few organelles, although annulate lamellae are found. The mitochondria have few cristae and are concentrated at one pole of the cell; they are sometimes found with intermitochondrial cement. These spermatogonia are separated from each other, having no intercellular bridges or inclusion in Sertoli cells, and are relatively undifferentiated; they correspond to stem cells. The spermatogonia beneath the apex are organized into cysts. First-generation spermatogonia are more dense and heterogeneous, their nuclei becoming smaller and their chromatin becoming denser during successive generations. In spermatocytes, the synaptinemal complex exists as a modified form until metaphase. The concentration of organelles in the cytoplasm increases and the organelles become more diversified as spermatogenesis progresses. Many cytoplasmic bridges are observed (several per cell), indicating that the cells remain in contact after several divisions. These changes in germ cell structure have been related to some of the characteristic features of spermatogenesis in guppy, e.g. the large number of spermatogonial generations and the complexity of spermiogenesis.  相似文献   

6.
DNA topology in histone- and protamine-depleted nuclei (nucleoids) from somatic cells, sperm, and spermatogenic cells was studied to determine if the superhelical configuration of DNA looped domains is altered during spermatogenesis. The expansion and contraction of nucleoid DNA was measured with a fluorescence microscope following exposure of nucleoids to different concentrations of ethidium bromide (EB). Nucleoids from Xenopus laevis erythrocytes, primary spermatocytes, and round spermatids, and from Rana catesbeiana sperm all exhibited a biphasic change (condensed-relaxed-condensed) in size as a function of exposure to increasing concentrations (0.5–100 g/ml) of EB, indicating that they contain negatively supercoiled DNA. In contrast, DNA in sperm nucleoids from Xenopus laevis and Bufo fowleri was relaxed and expanded at low (0.5–6 g/ml) EB concentrations, but became gradually condensed as the EB concentration was increased (6–100 g/ml). Nucleoids prepared from all cell types retained the general shape of the nucleus regardless of the superhelical configuration of the nucleoid DNA. Sperm nucleoid DNA condensed by 100 g/ml EB was relaxed by exposure to UV light, DNase I, proteinase K, or 4 M urea, but not by RNase A or 10 mM dithiothreitol. These results demonstrate that the DNA in sperm nucleoids is constrained in domains of supercoiling by nonbasic nuclear proteins. Negatively supercoiled DNA is present in nucleoids from cells with a full complement of histones, including Rana sperm, but not in nucleoids from Xenopus and Bufo sperm in which histones are replaced by intermediate-type protamines. Histone replacement in these species, therefore, is accompanied by unfolding of nucleosomal DNA and active removal of the negative supercoils. Results presented also suggest an important role for the nonbasic nuclear proteins of sperm in the morphogenesis of the nucleus and the arrangement of DNA.  相似文献   

7.
The organization of sperm chromatin in the dasyurid marsupial, Sminthopsis crassicaudata, was investigated using various morphological techniques. Transmission electron microscopy indicates two quite distinct chromatin regions became evident late in spermiogenesis with an outer globular region containing blocks of very electron-dense chromatin. Fluorescent light microscopical studies after staining with DNA dyes and 7-amino actinomycin D of testicular, caput, and cauda epididymal spermatozoa showed that this region fluoresced less brightly than the rest of the nucleus, indicating the presence of fewer DNA binding sites. Freeze fracture showed that the chromatin in most of the nucleus had randomly arranged particles of various sizes, but that of the outer region was composed entirely of small particles. This outer region was more resistant to low concentrations of the ionic detergent, SDS, whereas both guanidine hydrochloride and urea together with sodium chloride generally dispersed all the chromatin except that in the outer globular region and in a localized area of the nucleus beneath the acrosome. This study has thus revealed that the outer globular chromatin of these spermatozoa responds differently to ionic detergents and protein denaturing agents and has a different chromatin organization than most of the rest of the nucleus. The significance of these differences remains, however, to be determined. © 1994 Wiley-Liss, Inc.  相似文献   

8.
9.
10.
Here we report a detailed analysis of waves of histone acetylation that occurs throughout spermatogenesis in mouse. Our data showed that spermatogonia and preleptotene spermatocytes contained acetylated core histones H2A, H2B and H4, whereas no acetylated histones were observed throughout meiosis in leptotene or pachytene spermatocytes. Histones remained unacetylated in most round spermatids. Acetylated forms of H2A and H2B, H3 and H4 reappeared in step 9 to 11 elongating spermatids, and disappeared later in condensing spermatids. The spatial distribution pattern of acetylated H4 within the spermatids nuclei, analyzed in 3D by immunofluorescence combined with confocal microscopy, showed a spatial sequence of events tightly associated with chromatin condensation. In order to gain an insight into mechanisms controlling histone hyperacetylation during spermiogenesis, we treated spermatogenic cells with a histone deacetylase inhibitor, trichostatin A (TSA), which showed a spectacular increase of histone acetylation in round spermatids. This observation suggests that deacetylases are responsible for maintaining a deacetylated state of histones in these cells. TSA treatment could not induce histone acetylation in condensing spermatids, suggesting that acetylated core histones are replaced by transition proteins without being previously deacetylated. Moreover, our data showed a dramatic decrease in histone deacetylases in condensing spermatids. Therefore, the regulation of histone deacetylase activity/concentration appears to play a major role in controling histone hyperacetylation and probably histone replacement during spermiogenesis.  相似文献   

11.
DNase I-hypersensitivity of rat spermatogenic cells was analyzed 1) to establish overall patterns of hypersensitivity in individual cell types, 2) to correlate these patterns with known changes in chromatin organization and function, and 3) to provide a foundation for further analyses examining DNase I-hypersensitivity and the localization of specific genes during spermatogenesis. Parameters for in situ nick translation, using radioactive and fluorescent probes to visualize DNase I-hypersensitive regions (DHR), were established for fixed and sectioned testicular preparations, permeabilized cells, and isolated germ cell nuclei. As anticipated, the pattern of DHR changed in a cell-type specific manner during the course of spermatogenesis, reflective of known stage-dependent alterations in the composition and structure of both the chromatin and the nuclear lamina/matrix as well as changes in gene expression. DHR in preleptotene spermatocytes were primarily peripheral, while in pachytene spermatocytes they were localized along the condensed chromosomes. The pattern of DHR changed from "checkerboard" in steps 7-8 round spermatid nuclei to "lamellar" in steps 10-11 elongating spermatids. In steps 12-13 elongating spermatids. DHR were localized throughout the nuclei or in a graded manner--increasing from anterior to posterior and mirroring the pattern of chromatin condensation. However, unlike the case in other stages, DNA of steps 12-13 elongating spermatids was exquisitely sensitive to nick translation even in the absence of exogenous DNase I. In contrast to the labeling of earlier stages, steps 16-19 spermatids and mature spermatozoa did not demonstrate DNase I-hypersensitivity under any conditions employed. A variety of agents that interact with topoisomerase II and DNA (teniposide, novobiocin, ethidium bromide, and adenosine triphosphate) were tested to determine the basis for the unique sensitivity to nick translation of steps 12-13 elongating spermatids. None of the agents tested, however, affected this unique labeling. The sensitivity of steps 12-13 elongating spermatids to nick translation in the absence of exogenous nuclease indicators the presence of endogenous nicks, which may relieve torsional stress and aid rearrangement as the chromatin is packaged into a form characteristic of the mature spermatozoon.  相似文献   

12.
The number of cells in the S-phase fraction of the cell cycle reflects proliferative activity. Using flow cytometry histograms and the Phoenix M+ cell cycle program, the percent of cells in the S-phase fraction was measured in single cell suspensions prepared from testes of hamsters of different ages. A cyclical pattern with a period of 9 days, superimposed on another rhythm with a 38 day period was observed (p < 0.01) during hamster maturation and it disappeared after the second spermatogenic wave, where the S phase values reached a plateau. It was concluded that maturing animals passed through a stage in which testicular biological rhythm was involved. Therefore it was concluded that it takes approximately two spermatogenic waves before the proliferation rate in the testis reached a steady state.  相似文献   

13.
14.
Changes in the activity of antioxidizing enzymes during spermatogenesis   总被引:1,自引:0,他引:1  
A study was made of variation in the activity of the antioxidant defense system enzymes superoxide dismitase (SOD), glutathione peroxidase, (GP) and glutathione-S-transferase (GT) in the cells of mouse spermatogenic epithelium. The cells were fractionated in the gradient of human serum albumin using the STAPUT system. SOD activity was comparable to that in liver cells, that of GP was one order to magnitude lower, and that of GT one order of magnitude higher than in the liver. Differentiation of spermatogenic cells demonstrated phase variations in the activity of these enzymes, with a maximum seen at the stage of late pachytene spermatocytes. SOD, GP and GT activities were discovered to be reduced in postmeiotic cells (early spermatid fractions) by 2.8-3.5; 1.9-2.3 and 4.2-5.6 times, respectively. That reduction was followed by activation of the enzymes at the late stages of spermiogenesis.  相似文献   

15.
The levels of dolichyl phosphate and 2,3-dehydrodolichyl diphosphate synthase were determined in seminiferous tubules of prepuberal rats to assess any changes occurring during early stages of spermatogenesis. Dolichyl phosphate increased in concentration two- to threefold from Day 10 to Day 23 after birth. A method was optimized to measure dehydrodolichyl diphosphate synthesis from delta 3-[14C]isopentenyl diphosphate and t,t-farnesyl diphosphate in homogenates of seminiferous tubules. Both dehydrodolichyl mono- and diphosphates were observed as products of the in vitro assay. The specific activity of tubular synthase increased twofold between Day 7 and Day 23 and decreased similarly between Day 23 and Day 60. Since there was a parallel increase in the concentration of tubular dolichyl phosphate and dehydrodolichyl diphosphate synthase activity during early stages of spermatogenesis, it is proposed that the level of dolichyl phosphate may be controlled at least in part by the regulation of de novo dehydrodolichyl diphosphate biosynthesis. The synthase was also solubilized from tubular membranes with deoxycholate and partially purified by chromatography.  相似文献   

16.
The effects of ultra wideband pulse radiation on human cells were investigated. The density of the flow of energy on the surface of irradiated object varied from 10−6 to 10−2 W/cm2 with exposure of 10 s. It was shown that heterochromatin granule quantity in cell nuclei increased under the influence of radiation from 10−4 to 10−2 W/cm2. In some intervals the effect increased with irradiation dose. At irradiation intensity 10−3 W/cm2 the process of heterochromatin granule formation was fully reversible after 2 h of recovery; at intensity 10−2 W/cm2 the reversion of irradiation effects was not full. The data obtained indicated the strong biological activity of ultra wideband ultra short pulse radiation.  相似文献   

17.
Summary Using different sources of protoplasts and two complementary techniques, flow cytometry and image analysis, to study the cell-cycle phases, we sought to define the particular protoplast state associated with the disposition to divide. Both inPetunia and inNicotiana plumbaginifolia, tissues with a higher G2 frequency (from different aged plants) yielded protoplasts capable of increased cell division. InSorghum, the age of the plant does not modify the proportion of G2 nuclei in leaf protoplasts, and we used root protoplasts to increase G2 frequencies. InHelianthus annuus, leaf protoplasts did not divide; however, hypocotyl protoplast preparations with relatively high 4C DNA frequencies do divide. Moreover, image analysis of chromatin structure indicated that leaf nuclei were in the G0 phase, unlike those from hypocotyls which were in G1. A high frequency of protoplasts with G2 nuclei appears to be correlated with the ability of a given preparation to undergo division; conversely, the differentiated G0 state is not conducive to division.  相似文献   

18.
Caspase activation throughout the first wave of spermatogenesis in the rat   总被引:7,自引:0,他引:7  
Early in postnatal life, the first wave of spermatogenesis is accompanied by an initial wave of germ cell apoptosis. This may reflect an adjustment in the number of germ cells that can be adequately maintained by Sertoli cells. Two major pathways (intrinsic and extrinsic) are involved in the process of caspase activation and apoptosis in mammalian cells. The extrinsic pathway is characterized by the oligomerization of death receptors such as FAS or tumor necrosis factor, followed by the activation of caspase-8 and caspase-3. The intrinsic pathway involves the activation of procaspase-9, which in turn activates caspase-3. Extensive information is available concerning apoptotic inducers and their possible mechanisms in the adult rat. However, no data exist regarding the molecular and cellular mechanisms governing physiological cell death during puberty in the male rat. We have studied caspase activation throughout the first wave of spermatogenesis in the rat under physiological conditions, by combining the TUNEL procedure with the localization of active caspases in germ cells. We observed TUNEL-positive germ cells in rats of 5–40 days of age, the highest number being found in 25-day-old rats. TUNEL-positive and caspase-3-positive germ cells appeared as long chains of interconnected germ cells in 25-day-old rats. Caspase activation was assayed by either immunohistochemistry with antibodies against active caspase-3, -8, and -9, or by determining enzymatic activity in seminiferous tubules extracts. Both techniques showed activation of caspase-3, -8, and -9 in 25-day-old rats and low enzymatic activity at other ages. Confocal scanning laser microscopy indicated that active caspase-3, -8, and -9 co-localized with TUNEL-positive cells. Thus, caspase-3, -8, and -9 are active in apoptotic germ cells during the first wave of rat spermatogenesis. The extrinsic pathway of apoptosis may therefore play an important role in germ cell apoptosis during puberty in the rat.This work was financed by a research grant from FONDECYT (1040800) to R.D.M.  相似文献   

19.
Accessibility of mouse testicular and vas deferens (vas) sperm cell DNA to acridine orange, propidium iodide, ellipticine, Hoechst 33342, mithramycin, chromomycin A3, 4'6-diamidino-2-phenylindole (DAPI), and 7-amino-actinomycin D (7-amino-AMD) was determined by flow cytometry. Permeabilized cells were either stained directly or after pretreatment with 0.06 N HCl. For histone-containing tetraploid, diploid, and round spermatid cells, HCl extraction of nuclear proteins caused an approximately sixfold increase of 7-amino-AMD stainability but had no significant effect on DAPI stainability. For these same cell types, the stainability with other intercalating (acridine orange, propidium iodide, ellipticine) and externally binding (Hoechst 33342, mithramycin, chromomycin A3) dyes was increased by 1.6- to 4.0-fold after HCl treatment. In sharp contrast, HCl treatment of vas sperm did not increase the staining level of 7-amino-AMD, DAPI, or propidium iodide but did increase the staining level for the other intercalating dyes (1.3- to 1.5-fold) and external dyes (1.3- to 1.9-fold). Elongated spermatids that contain a mixture of protein types including histones, transition proteins, and protamines demonstrated the greatest variability of staining with respect to type of stain and effect of acid extraction of proteins. In general, for nearly all dyes, the round spermatids had an increased level and tetraploid cells had a decreased level of stainability relative to the same unit DNA content of diploid cells. The observed differential staining is discussed in the context of chromatin alterations related to the unique events of meiosis and protein displacement and replacement during sperm differentiation.  相似文献   

20.
Chromatin may be attached to the nuclear envelope through interaction of the nuclear membrane lamins A, B, and C. Such a hypothesis requires that these proteins are present in all cells with chromatin attachment to the nuclear envelope. We have investigated the distribution of the lamins during spermatogenesis in mouse, which exhibits extremes in nuclear envelope structural changes. By immunohistochemical techniques using human auto-antibodies and monoclonal antibodies against these molecules, we found that the lamins persist through all stages of spermatogenesis, though in highly variable amounts. They are also present during meiotic prophase (pachytene) when chromosomes are only locally attached to the nuclear envelope, analogous to the early prophase of somatic cells. Restructuring of the early spermatid nuclear envelope is accompanied by the appearance of a new lamin at the acrosomal fossa. In the epididymal spermatozoon the distribution of different lamins varies markedly over the nucleus suggesting special structural functions. The presence of lamins throughout spermatogenesis supports the concept that they are a general feature of the nuclear envelope structure, even where a lamina is not recognizable ultrastructurally.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号