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.
Lymphocyte chromosomes from a cercopithecoid species, Macaca mulatta, were studied for the occurrence of lateral asymmetry in constitutive heterochromatin. The technique consisted of growing the lymphocytes for one cell cycle in BrdUrd, staining with 33258 Hoechst, exposing them to UV light, treating them with 2 SSC and staining with Giemsa. This procedure revealed asymmetric staining in the region of constitutive heterochromatin of the nucleolar organizer marker chromosome (no. 13 of the complement). In these chromosomes, the darkly staining region was confined at any given point to a single chromatid, while the corresponding region on the sister chromatid was lightly stained. This pattern of asymmetric staining in the constitutive heterochromatic region was not observed in any other chromosome of Macaca mulatta. The lateral asymmetry of constitutive heterochromatin in this species is presumed to reflect the strand bias in the distribution of thymine in the alphoid DNA fractions.  相似文献   

2.
Lymphocyte chromosomes from a cercopithecoid species, Macaca mulatta, were studied for the occurrence of lateral asymmetry in constitutive heterochromatin. The technique consisted of growing the lymphocytes for one cell cycle in BrdUrd, staining with 33258 Hoechst, exposing them to UV light, treating them with 2 SSC and staining with Giemsa. This procedure revealed asymmetric staining in the region of constitutive heterochromatin of the nucleolar organizer marker chromosome (no. 13 of the complement). In these chromosomes, the darkly staining region was confined at any given point to a single chromatid, while the corresponding region on the sister chromatid was lightly stained. This pattern of asymmetric staining in the constitutive heterochromatic region was not observed in any other chromosome of Macaca mulatta. The lateral asymmetry of constitutive heterochromatin in this species is presumed to reflect the strand bias in the distribution of thymine in the alphoid DNA fractions.  相似文献   

3.
Spermatogenesis involving an additional chromosome reduplication during zygotene in sporadic males and intersexes of the thelytokous phasmid Carausius morosus Br. has been examined using differential staining of chromatids after 5-bromodeoxycytidine incorporation. After reduplication autobivalents are formed by synapsis between identical sister chromosomes. Chiasmata are only formed after reduplication; they do not occur in constitutive heterochromatin, but can be formed in facultative heterochromatin, dependent on heteropycnosis and sex. Quadrivalents and U-type exchanges occur. In spermatogonia and spermatocytes the number of differentially stained chromosomes varies considerably; sister chromatid exchanges hardly appear. Sex bivalents with differentially stained chromosomes have a lower chiasma frequency than normally stained sex bivalents. Bivalents show reduced staining of all four, two outer, or one inner chromatid. Autobivalents arise in the same way as diplochromosomes; chromatids with the oldest DNA sub-units remain together during reduplication and are thus involved in sister chromosome pairing. The additional reduplication begins 7 days after the premeiotic S-phase, first metaphase after 19 days. Spermatogenesis is abnormal from first anaphase onwards.  相似文献   

4.
Summary Variations in lateral asymmetry of constitutive heterochromatin were studied in 30 normal individuals with reference to the chromosomal regions 1q12, 9q12, 15p11, 16q12 and Yq12. The technique consisted of growing human lymphocytes for one cell cycle in BrdU, staining with 33258 Hoechst, exposing them to UV light, treating them with 2 x SSC, and staining with Giemsa. This procedure revealed asymmetric staining in the region of constitutive heterochromatin in these chromosomal regions. Chromosomes 15, 16, and Y showed simple lateral asymmetry, whereas chromosome 1 showed both simple and compound asymmetry. In 15 cases, compound lateral asymmetry was evident in both homologues of chromosome 1, 12 cases showed compound lateral asymmetry in one homologue and simple lateral asymmetry in the other, and the remaining three cases showed simple lateral asymmetry in both the homologues. The centromere region of chromosome 9 stained symmetrically with this technique. The lateral asymmetry is presumed to reflect the strand bias in the distribution of thymine in satellite DNA fractions.  相似文献   

5.
M A Kim 《Humangenetik》1975,28(1):57-63
Using the fluorochrome "Hoechst 33258", intensity of fluorescence was found to differ distinctly between the sister chromatids in the paracentric regions of chromosomes 1, 16, and 19, after one round of replication in medium containing BUdR. Thus the effect of fluorescence asymmetry is not limited to the part of the Y chromosomes that fluoresces intensely with quinacrine; it can also be determined in the weakly Q-fluorescent pericentric regions of chromosomes, which are known to be the sites where highly reiterated sequences of satellite DNA are located. However, an exception is the paracentric region of chromosome 9 which does not show the effect of lateral asymmetry. The difference of fluorescence intensity in the heterochromatic regions of the sister chromatids of human chromosome 1 is measured by densitometric tracement along the long axes of chromosomes; this is obtained from two individuals with an "uncoiler" heterchomatic block (type III) having a relative intensity of 1:1.93 in an average of the total measured blocks. This corresponds to the uneven distribution of thymine base of 22.8 and 43.2 in the two strands of the DNA double hexlix. A chromatid exchange rate of 9 in 100 metaphases per cell cycle was found within the uncoiler region of chromosome 1.  相似文献   

6.
Chinese hamster strain cells were cultured in the presence of BUdR and air-dried on slides. The chromosome preparations were incubated in 1 M NaH2PO4 at 88 °C for 4–6 min and stained with Giemsa. The reverse type of sister chromatid differential staining occurred, in which unifilarly BUdR-substituted chromatids stained faintly and bifilarly substituted chromatids stained darkly. Feulgen reaction performed on the same chromosomes after removing Giemsa stain showed the same type of differential staining.  相似文献   

7.
The frequency of sister chromatid exchanges (SCEs) was determined for the chromosomes (except Y2) of the Indian muntjac stained by the fluorescence plus Giemsa (FPG) or harlequin chromosome technique. The relative DNA content of each of the chromosomes was also measured by scanning cytophotometry. After growth in bromodeoxyuridine (BrdU) for two DNA replication cycles. SCEs were distributed according to the Poisson formula in each of the chromosomes. The frequency of SCE in each of the chromosomes was directly proportional to DNA content. A more detailed analysis of SCEs was performed for the three morphologically distinguishable regions of the X-autosome composite chromosome. The SCE frequency in the euchromatic long arm and short arm were proportional to the amount of DNA. In contrast, the constitutive heterochromatin in the neck of this chromosome contained far fewer SCEs than expected on the basis of the amount of DNA in this region. A high frequency of SCE, however, was observed at the point junctions between the euchromatin and heterochromatin.  相似文献   

8.
The experiments described were directed toward understanding relationships between mouse satellite DNA, sister chromatid pairing, and centromere function. Electron microscopy of a large mouse L929 marker chromosome shows that each of its multiple constrictions is coincident with a site of sister chromatid contact and the presence of mouse satellite DNA. However, only one of these sites, the central one, possesses kinetochores. This observation suggests either that satellite DNA alone is not sufficient for kinetochore formation or that when one kinetochore forms, other potential sites are suppressed. In the second set of experiments, we show that highly extended chromosomes from Hoechst 33258-treated cells (Hilwig, I., and A. Gropp, 1973, Exp. Cell Res., 81:474-477) lack kinetochores. Kinetochores are not seen in Miller spreads of these chromosomes, and at least one kinetochore antigen is not associated with these chromosomes when they were subjected to immunofluorescent analysis using anti-kinetochore scleroderma serum. These data suggest that kinetochore formation at centromeric heterochromatin may require a higher order chromatin structure which is altered by Hoechst binding. Finally, when metaphase chromosomes are subjected to digestion by restriction enzymes that degrade the bulk of mouse satellite DNA, contact between sister chromatids appears to be disrupted. Electron microscopy of digested chromosomes shows that there is a significant loss of heterochromatin between the sister chromatids at paired sites. In addition, fluorescence microscopy using anti-kinetochore serum reveals a greater inter-kinetochore distance than in controls or chromosomes digested with enzymes that spare satellite. We conclude that the presence of mouse satellite DNA in these regions is necessary for maintenance of contact between the sister chromatids of mouse mitotic chromosomes.  相似文献   

9.
A sister chromatid differential staining pattern is observed if chromosomes replicate for two cycles in the presence of 5-bromodeoxyuridine (BUdR) and are subsequently stained in Hoechst 33258, irradiated with black light, and then stained in Coomassie Brilliant Blue R-250. In this pattern the chromatids containing DNA that is bifilarly substituted with BrdUrd are darkly stained and the chromatids with DNA that is unifilarly substituted are lightly stained. This staining pattern is the reverse of that found when slides are stained in Hoechst plus Giemsa. Slides stained with either Giemsa or Coomassie Blue can be destained and restained repeatedly with the other stain to alternate the pattern observed.  相似文献   

10.
We have studied two aspects of the process of sister chromatid separation in the Drosophila melanogaster neuroblasts. First, we analyzed the requirement of a functional spindle for sister chromatid separation to take place using microtubule depolymerizing drugs such as colchicine or a reversible analogue (MTC). Incubation of this tissue in colchicine causes the cells to block irreversibly at metaphase and no significant levels of sister chromatid separation were observed even after long periods of incubation. Exposure of neuroblasts to MTC also causes cells to block at metaphase, but after reversion most of the cells enter anaphase and are thus able to complete sister chromatid separation. These results imply that a functional spindle is required for sister chromatid separation. Second, we studied the role of heterochromatin during chromatid pairing and subsequent separation in chromosomes which carry either one or two extra pieces of heterochromatin. The results indicate that sister chromatids establish strong pairing along the translocated heterochromatin. During the early stages of anaphase, these chromosomes separate first the centromeric region and later the regions bearing extra heterochromatin. These results indicate that constitutive heterochromatin plays an important role for sister chromatid pairing and might be involved in the process of separation.  相似文献   

11.
Summary Second generation BrdU-labeled acrocentric chromosomes exhibit NOR lateral asymmetry (NLA) in metaphases that have been sequentially stained with silver and the Hoechst-Giemsa sister chromatid differential (SCD) technique. The NLA presumably results from suppression of NOR activity in the doubly-substituted chromatid. Examination of single chromatid (NOR) associations in pairs of acrocentrics reveals that light chromatids associate less frequently than dark chromatids and that the frequency distribution of dark and light alignment configurations can be explained by this differential tendency to associate. Thus, it appears that a hypothesis of non-random chromatid segregation as an explanation for non-random chromatid alignments in associating acrocentric chromosomes is unwarranted.This work is a joint project of The University of Texas M.D. Anderson Hospital and Tumor Institute and the John S. Dunn Research Foundation of Houston, Texas  相似文献   

12.
The fluorescence of human lymphocyte chromosomes stained with sulfhydryl group-specific fluorochromes is markedly enhanced by a mild near-ultraviolet irradiation pretreatment, indicating breakage of protein disulfide bonds. When metaphase preparations of cells cultured in the presence of BrdU during two cell cycles are irradiated and subsequently stained with the sulfhydryl group-specific fluorescent reagents used in this study, a differential fluorescence of sister chromatids is observed. After staining with the DNA-specific fluorochrome DAPI an opposite pattern of lateral differentiation appears. It can be concluded that the chromatid containing bifilarly BrdU-substituted DNA has a higher content of sulfhydryl groups than the chromatid containing unifilarly BrdU-substituted DNA. This implies a more pronounced effect of breakage of disulfide bonds in the chromatid with the higher degree of BrdU-substitution. BrdU-containing chromosomes pretreated with the mild near-ultraviolet irradiation procedure used by us, do not show any differentiation of sister chromatids after Feulgen staining. Using sulfhydryl group-specific reagents, differential fluorescence of sister chromatids could still be induced by irradiation with near-ultraviolet light after the complete removal of DNA from the chromosomes by incubation with DNase I. Thus, the protein effect of irradiation of BrdU-containing chromosomes takes place independently of what occurs to DNA.Our results indicate that subsequent to the primary alteration of chromatin structure caused by the incorporation of BrdU into DNA, breakage of disulfide bonds of chromosomal proteins might play an important role in bringing about differential staining of sister chromatids, at least for those procedures that use irradiation as a pretreatment or prolonged illumination during microscopic examination.  相似文献   

13.
Summary Isolabeling observed during sister chromatid differentiation (SCD) was studied from human skin fibroblasts by the fluorescence-plus-Giemsa (FPG) technique. Bromodeoxyuridine (BrdU) was fed to exponentially dividing cells for 52 h to enable completion of two consecutive cycles of DNA replication. During this period, the late-replicating regions of some chromosomes were able to go through three replication cycles. These chromosome regions had evidently incorporated BrdU bifiliarly in both chromatids and hence, on staining with FPG, appeared isostained (isolabeled). Thus, incubation of exponentially dividing cells with BrdU for a period longer than that required for two cell cycles appears to be a suitable method for revealing the late-replicating regions of the genome, such as the X chromosome in a human female, as isolated.In another experiment with Indian muntjac chromosomes, isolabeled segments were darkly stained, which suggested unifilar incorporation of BrdU. In this case, unequal crossing-over or an unequal distribution of thymine residues probably is responsible for the isolabel.  相似文献   

14.
Lateral asymmetry refers to unequal fluorescent intensity between adjacent regions of sister chromatids. It has been observed in the centromeric regions of mitotic chromosomes of mouse or human origin when cells are grown in 5-bromo-2-deoxyuridine (BrdU) for a single round of DNA synthesis. The chromosome-orientation fluorescence in situ hybridization (CO-FISH) technique was used with pseudodiploid mouse cells to show that the regions of asymmetrical brightness coincide with major satellite repetitive DNA, and that the more heavily BrdU-substituted chromatid is the one that fluoresces less brightly. These observations support a 20 year old hypothesis on the origin of lateral asymmetry. Other observations suggest that differential loss of DNA from the heavily substituted chromatid also contributes to lateral asymmetry.  相似文献   

15.
At the metaphase/anaphase transition in the mouse and rat male germ lines during the perinatal period, sister centromeres separate before sister chromatids. This gives the chromosomes an unusual appearance that resembles the premature centromere division described in some human pathological conditions such as Roberts syndrome. At the same period, there is also an unusual pattern of DNA methylation, with strongly demethylated heterochromatin and methylated euchromatin. This suggests that chromosome DNA methylation may modulate chromatid and centromere splitting, without altering normal chromosome segregation.  相似文献   

16.
Hoechst 33258 fluorescent staining of bromodeoxyuridine substituted chromosomes provided a high resolution technique for following the segregation of replicated chromosomal DNA (Latt, 1973). Modifications have produced the same results after Giemsa staining (Wolff and Perry, 1975). Since this does not necessarily require Hoechst (Korenberg and Freedlander, 1975), we call this bromodeoxyuridine-Giemsa banding (BG-banding). We here describe a further modification which allows one to follow the T-rich strand of the AT-rich satellite DNA of C-band heterochromatin. We call this TC-banding. This technique was used to examine metacentric marker chromosomes found in mouse L-cells that contain many interstitial blocks of centromeric-type heterochromatin in each arm plus the usual two blocks of centromeric heterochromatin. One of the advantages of this technique for such chromosomes is that it is possible to distinguish first from second cell cycle sister chromatid exchange and unambiguously detect centromeric sister chromatid exchange. We found some chromosomes to have high rates of centromeric sister chromatid exchange. After one cycle in bromodeoxyuridine we could examine the satellite polarity of the heterochromatic DNA. Since there was no change in satellite polarity in any of the heterochromatic blocks, marker chromosomes could not have been formed by paracentric inversions, inverted insertions or inverted translocations. These results allow the formulation of several rules of chromosome organization.  相似文献   

17.
Structural organization of the heterochromatic region of human chromosome 9   总被引:2,自引:0,他引:2  
Giemsa-11, G-banding and Lateral Asymmetry staining techniques were used to define the substructure of the C-band heterochromatin of human chromosome 9, in a sample of 108 different chromosomes 9, from 54 individuals. In this sample, the juxtacentromeric portion of the C-band region stained positive by the G-banding technique while Giemsa-11 delineated a more distally located block. Examination of the pericentric inversions generally revealed that the entire C-band region is changed with the substructural organization left intact; i.e. the G-band is proximal, the G-11 distal to the centromere. The partial pericentric inversions were found to have larger than average amounts of G-band heterochromatin on the short arm. The G-11 staining was in its usual position on the long arm with none on the short arm. Such apparent inversions therefore may not represent true inversions. — Long heterochromatic regions frequently had a segmented appearance when stained with G-11; there was a dark G-band within the pale heterochromatic region when stained with the G-banding technique which corresponded in location to the achromatic gap produced by G-11. This extra G-band may have been derived from the juxtacentromeric G-band by processes analogous to unequal crossing over. — Simple lateral asymmetry was consistently present only in the G-band heterochromatin of those chromosomes 9 containing large blocks of G-band positive material. Examination of the portion of the C-band which would correspond to the G-11 positive material revealed no consistent patterns of asymmetry. Usually both strands were heavily stained and symmetrical but occasionally there were light areas present on one strand suggestive of compound lateral asymmetry.  相似文献   

18.
Mitotic chromosome structure and pathways of mitotic condensation remain unknown. The limited amount of structural data on mitotic chromosome structure makes it impossible to distinguish between several mutually conflicting models. Here we used a Chinese hamster ovary cell line with three different lac operator-tagged vector insertions distributed over an ∼1 μm chromosome arm region to determine positioning reproducibility, long-range correlation in large-scale chromatin folding, and sister chromatid symmetry in minimally perturbed, metaphase chromosomes. The three-dimensional positions of these lac operator-tagged spots, stained with lac repressor, were measured in isolated metaphase chromosomes relative to the central chromatid axes labeled with antibodies to topoisomerase II. Longitudinal, but not axial, positioning of spots was reproducible but showed intrinsic variability, up to ∼300 nm, between sister chromatids. Spot positions on the same chromatid were uncorrelated, and no correlation or symmetry between the positions of corresponding spots on sister chromatids was detectable, showing the absence of highly ordered, long-range chromatin folding over tens of mega-basepairs. Our observations are in agreement with the absence of any regular, reproducible helical, last level of chromosome folding, but remain consistent with any hierarchical folding model in which irregularity in folding exists at one or multiple levels.  相似文献   

19.
Paulo R. Alves 《Chromosoma》1982,85(5):713-722
A mouse cell line of C57B1/6J spontaneous melanoma (clone PG 19), and a C-type virus transformed cell line (G-8 clone 124) originating from normal Balb/c mice were used in a study of the centromeric heterochromatin region of BrdU-labelled chromosomes stained by the Giemsa pH 10.4 method. Three possible explanations for the generation of compound lateral asymmetry within the centromeric heterochromatin region of the laboratory mouse are discussed: 1) inverted translocation; 2) centric fusion followed by paracentromeric fission and 3) inversion of part of the centromeric satellite DNA. These processes could be of considerable genetic and evolutionary significance. The non-random spatial position of unstained and dark stained C-bands in BrdU-labelled diplochromosomes of endoreduplicated cells can be explained as being due to the localization of the old and new DNA chains in a unineme chromatid model. The late replicating regions are shown to be located on the inside of the half-chromatid close to the axial symmetry axis of the metaphase chromosome.  相似文献   

20.
A benzimidazole derivative, Hoechst 33258 can induce decondensation of constitutive heterochromatin in the mouse derived L cell chromosomes when the compound is given in sufficiently high concentration (40 micrograms/ml) to the L cell culture. Hoechst 33258 at low concentration (1 micrograms/ml, 16 h) cannot produce this effect on L cell chromosomes. Bromodeoxyuridine (BUdR) incorporation for one cell cycle simultaneous with the Hoechst 33258 treatment at low concentration could decondense heterochromatin segments in metaphase chromosomes. The heterochromatin decondensation, however, was asymmetric; it was observed only on one chromatid and the other of a chromosome remained in condensed state. The observation of asymmetric decondensation of heterochromatin by Hoechst 33258 after BUdR incorporation for one cell cycle, the association of A-T rich satellite DNA to mouse heterochromatin, and available data on the specific binding of Hoechst 33258 to A-T base pairs of DNA and on the higher affinity of the compound to BUdR substituted DNA than to ordinary DNA implied that the binding of Hoechst 33258 molecules to A-T rich satellite DNA is the cause of heterochromatin decondensation.  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