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.
Chinese hamsters from five strains with reciprocal translocations, T(1;3)7Idr, T(1;3)8Idr, T(1;2)9Idr, T(7;9)16Idr, and T(1;5)17Idr, and a karyotypically normal strain, CHS/Idr, were used to look for an interchromosomal effect by chromosomal analysis of meiotic cells and one-cell embryos. The frequencies of nondisjunction at first meiosis in five normal (+/+) males, calculated by doubling the number of hyperhaploid cells, ranged from 0.43% to 1.33%, and there was no significant difference in frequency among individuals. On the other hand, the frequency of hyperhaploid cells in males heterozygous for each translocation ranged from 3.0% to 11.8%, and the frequency of hyperhaploid cells with an extra translocation-unrelated chromosome ranged from 0.2% to 0.4%, which is no different from that estimated from scoring of +/+ males at the second meiotic metaphase. In one-cell embryos from crosses between karyotypically normal females and male heterozygotes for T(1;2)9Idr and T(7;9)16Idr, 1.1% and 0.5% of embryos had an extra translocation-unrelated chromosome. Compared with the control, the frequency of meiotic nondisjunction showed no increase in male heterozygotes for the reciprocal translocations. Therefore, the results suggest that multivalents and rearranged chromosomes existing at first and second meiosis in male Chinese hamsters exert no influence on segregation of normal bivalents and chromosomes unrelated to the rearrangements.  相似文献   

2.
The gametic and zygotic selection of genome imbalance was investigated in the Chinese hamster by direct chromosome analyses of spermatocytes and preimplantation embryos from crosses between chromosomally normal females and males heterozygous for a reciprocal translocation, T(2;10)3Idr, abbreviated here as T3. The karyotypes and the frequencies of embryos observed at the first cleavage in the cross +/+female X T3/+male were consistent with those expected from MII scoring in male T3 heterozygotes. Therefore, it was concluded that there was neither gametic selection against genome imbalance nor zygotic selection from fertilization until the first cleavage metaphase. However, 9.1-10.8% of embryos were arrested at the two-cell stage, and karyotypes of these embryos were confirmed as 22(2,10,10,10(2)), 21(2,10,10), and 21(2,10,10(2)). The common abnormality of these embryos was partial monosomy of chromosome 2. Among day 4 embryos, some chromosomally unbalanced embryos, mainly with a deficiency of other segments of chromosomes 2 and 10, had fewer blastomeres than chromosomally balanced embryos. This finding suggests that cleavage of these embryos had been retarded by day 4 of gestation.  相似文献   

3.
The selection of chromosomally abnormal gametes was investigated in the Chinese hamster by direct chromosome analysis of meiotic cells and one-cell embryos obtained from crossing heterozygotes for two reciprocal translocations, T(1;3)7Idr and T(1;3)8Idr. Expected frequencies of male and female gametes with different chromosome constitutions were estimated by scoring of secondary meiotic metaphase (MII) cells in the translocation heterozygotes. The frequency of gametes with each karyotype that participated in fertilization was investigated in pronuclei from translocation heterozygotes in one-cell embryos obtained from crossing the heterozygotes with karyo-typically normal animals. Compared with the expected frequencies from MII scoring, the frequencies of male pronuclei having some karyotypes in one-cell embryos decreased significantly. The karyotypes of male pronuclei showing a decreased frequency were commonly characterized by a deficiency of the long-arm segment of chromosome 1 (q13----qter) or by a deficiency of almost the whole arms of chromosome 3. On the other hand, the frequencies of female pronuclei with the same karyotypes were all consistent with those estimated from MII scoring. These results suggest that sperm nullisomic for certain segments of some chromosomes may fail to participate in fertilization.  相似文献   

4.
E K?fer 《Mutation research》1986,164(3):145-166
A system of tests with the ascomycete Aspergillus nidulans was devised that can detect 3 primary effects of genotoxic agents: (1) increases in mitotic crossing-over; (2) induced aneuploidy; and (3) clastogenic effects which cause chromosomal imbalance. Conidia of a new diploid tester strain, heterozygous for 4 recessive markers which alter conidial color, are treated and plated onto nonselective media. In cases of induced crossing-over, large color segments are found in normal green colonies, frequently adjacent to reciprocal twin segments. In contrast, both malsegregation and chromosome breakage produce unbalanced types which grow poorly and segregate further. Cases with yellow segregants are replated and their secondary diploid sectors tested for markers which are located on both chromosome arms in coupling with yA. Induced aneuploidy can be distinguished from chromosome breakage by the pattern of marker segregation. Any aneuploid type will produce euploid sectors solely by segregation of whole chromosomes; trisomic colonies (yA / yA / +) will show 1:2 ratios for yellow (homozygous yA) to parental green (yA/+) sectors and have characteristic phenotypes. Other induced unbalanced types, if heterozygous for deletions or aberrations may produce yellow diploid sectors by secondary crossing-over as well as by nondisjunction and such cases show unique patterns of genetic segregation and non- predictable phenotypes. As a complementary test, haploid strains are treated and induced abnormally growing types are replated and classified by phenotype. Aneuploids are unstable and produce many normal sectors, and some of these disomic or trisomic types can be visually identified.In contrast, induced deletions are lethal, and duplications or 'morphological' mutants show much more stable abnormal phenotypes. This test system was used to characterize the primary effects of gamma-rays and chloral hydrate. Results and evidence were as follows: (1) A dose-dependent increase of color segments resulting from reciprocal crossing-over was found after treatment of dividing nuclei in germinating diploid conidia with gamma-rays, but not with chloral hydrate. (2) Highly aneuploid and polyploid types were induced in diploid and haploid germinating conidia by chloral hydrate but not to any significant extent by gamma-rays. (3) gamma-Rays caused a dose- dependent increase off abnormally growing colonies when dormant or germinating diploid conidia were treated. These colonies produced secondary euploid sectors by spontaneous nondisjunction and frequently also by crossing-over, which provided evidence for induced semidominant and recessive lethal mutations of many types.  相似文献   

5.
Previous investigations have shown the sex determination in the monogenic blowfly Chrysomya rufifacies to be controlled by a cytologically not discernible homogamety-heterogamety mechanism in the female. Female-producing (thelygenic) females are assumed to be heterozygous for a dominant female sex realizer (F′) with sex-predetermining properties, while male-producing (arrhenogenic) females as well as males are supposed to be homozygous for the recessive allele (f). In order to identify the genetic sex chromosomes of C. rufifacies among its five pairs of long euchromatic chromosomes (nos.1–5) plus one pair of small heterochromatic ones (no. 6), all chromosomes were marked by reciprocal translocations induced by X-ray treatment of adult males. The inheritance of thirteen different heterozygous translocations has been analyzed. All of the translocations (eleven) between two of the four longer chromosomes did not show sex-linked inheritance, thus demonstrating the autosomal character of the chromosomes nos 1, 2, 3 and 4. The same is true for the translocation T6 (2/6). Therefore the small heterochromatic chromosome no. 6, corresponding to the morphologically differentiated sex chromosomes within the amphogenic calliphorid species, remains without sex determining function in the monogenic fly. This could be confirmed by the analysis of monosomic (monosomy-6) and trisomic (trisomy-6) individuals, which resulted from meiotic non-disjunction in T6/+ translocation heterozygotes. Contrary to these translocations, the heterozygous 5/2 translocation (T14) exhibited sex-linked inheritance: There was but a very low frequency (0,76 per cent) of recombinants resulting from crossing-over between F′/f and the translocation breakage point in thelygenic F1 T14/+ females. The sex-linked inheritance of T14 was confirmed by the progeny of a thelygenic F1 T14/+ female crossed to a homozygous T14/T14 translocation male. Among the offspring of that F1 T14/+ female, which had received the translocation from its father, all of the F2 T14/+ females were thelygenic compared to their arrhenogenic T14/T14 sisters. These results prove that the chromosomes of pair no. 5 genetically act as X′X-XX sex chromosomes in C. rufifacies.  相似文献   

6.
Artificially induced translocation stocks have been used to physically map the barley genome; however, natural translocations are extremely uncommon in cultivated genotypes. Albacete is a barley variety widely grown in recent decades in Spain and carrying a reciprocal translocation which obviously does not affect its agronomical fitness. This translocation has been characterized by a combination of cytological and molecular genetic approaches. Firstly, recombination frequencies between markers on chromosomes 1H and 3H were estimated to determine the boundaries of the reciprocal interchange. Secondly, 1H-3H wheat barley telosome addition lines were used to assign selected markers to chromosome arms. Thirdly, fluorescence in situ hybridization (FISH) with rDNA probes (5S and 18S-5.8S-26S) and microsatellite probes [(ACT)(5), (AAG)(5) and (CAG)(5)] was used to determine the locations of the translocation breakpoints more precisely. Fourthly, fine-mapping of the regions around the translocation breakpoints was used to increase the marker density for comparative genomics. The results obtained in this study indicate that the translocation is quite large with breakpoints located on the long arms of chromosomes 1H and 3H, between the pericentromeric (AAG)(5) bands and above the (ACT)(5) interstitial distal bands, resulting in the reciprocal translocation 1HS.1HL-3HL and 3HS.3HL-1HL. The gene content around the translocation breakpoints could be inferred from syntenic relationships observed among different species from the grass family Poaceae (rice, Sorghum and Brachypodium) and was estimated at approximately 1,100 and 710 gene models for 1H and 3H, respectively. Duplicated segments between chromosomes Os01 and Os05 in rice derived from ancestral duplications within the grass family overlap with the translocation breakpoints on chromosomes 1H and 3H in the barley variety Albacete.  相似文献   

7.
Heterozygous chromosome rearrangements such as reciprocal translocations are most accurately displayed as two-dimensional linkage maps. Standard linkage mapping software packages, such as MapMaker, generate only one-dimensional maps and so reciprocal translocations appear as clusters of markers, even though they originate from two nonhomologous chromosomes. To more accurately map these regions, researchers have developed statistical methods that use the variance in map distance to distinguish among the four segments (two translocation, two interstitial) of the translocation. In this study, we describe modifications to one of these protocols, that proposed by Livingstone et al. (2000). We also introduce QuadMap, a new software application for dissecting heterozygous translocation-affected linkage maps.  相似文献   

8.
A UV-induced sulphite-requiring mutant (sD50) consistently shows mitotic linkage to groups I and VIII in haploids from heterozygous mapping diploids. This linkage was found to be due to a reciprocal translocation T2(I;VIII) which could not be separated from the sulphite requirement in about 100 tested progeny from heterozygous crosses, and both may well have been induced by the same mutational event. T2(I;VIII) is the first case of a reciprocal translocation in Aspergillus which showed meiotic linkages between markers of different linkage groups, and, in addition, involved chromosome arms containing markers suitable for complete mapping by the technique of mitotic recombination in homozygous translocation diploids.-Using various selective markers, haploid segregants and diploid crossovers of all possible types were isolated from homozygous translocation diploids. (1) Haploid segregants showed new linkage relationships in T/T diploids: all available markers of VIII now segregated as a group with the majority of the markers of I, except for the markers of the left tip of I. These formed a separate linkage group and are presumably translocated to VIII. (2) Diploid mitotic crossovers confirmed this information and showed that the orientation of the translocated segments was unchanged. These findings conclusively demonstrate that T2(I;VIII) is a reciprocal translocation due to an exchange of the left tip of group I with the long right arm of group VIII.-Since the position of the break on VIIIR was found to be at sD50 this marker could be used to map the break on IL by meiotic recombination in heterozygous crosses. In addition, such crosses showed reduced recombination around the breaks, so that it was possible to sequence markers which normally are barely linked.  相似文献   

9.
Electron microscopic analysis of synaptonemal complexes and analysis of chiasmata distribution in male mice heterozygous for Robertsonian translocation T(16; 17)7Bnr - (Rb7), for synaptonemal reciprocal translocation T(16;17)43H - (T43), in double heterozygotes for these translocations and in males with partial trisomy of the proximal region of chromosome 17 was carried out. Synaptic disturbances around the breakpoints of the translocations, such as asynapsis of homologous regions of partners and non-homologous synapsis of centromeric regions of acrocentric chromosomes, were revealed. Synaptic regularity in the proximal part of the chromosome 17 appeared to be affected by no t12 haplotype. Good coincidence between sizes of mitotic chromosomes and corresponding lateral elements of synaptonemal complexes was found for all chromosomes, with the exception of Rb7 in trisomics. In the latter karyotype, the proximal part of chromosome 17 involved in Robertsonian fusion seems to be shortened in the course of zygotene and never synapted with homologous segment of neither the acrocentric chromosome 17 nor large product of reciprocal translocation. Drastic increase in chiasmata frequency in the proximal part of chromosome 17 was revealed in heterozygotes for T43H and in trisomics, as compared with the double heterozygotes Rb7/T43. The latter finding was explained by the existence of two independent pairing segments in the former karyotypes.  相似文献   

10.
The influence of age and hormones on chromosome segregation at meiosis I was studied in female mice heterozygous for the T(2;4)1Gö translocation. Females of two age groups (18–22 and 40–56 weeks old) were stimulated for ovulation with different doses of gonadotropins (1.5 IU PMS/1.0 IU HCG or 10 IU PMS/10 IU HCG). Analysis of metaphase II oocytes revealed the highest level of hyperhaploidy (1.8%) and presegregation (4.4%) in the young females receiving the low dose. Presegregation preferentially affected the small 42 marker chromosome. There was no significant interference of the tetravalent with disjunction of the nontranslocated normal bivalents. Moreover, no remarkable difference in the mode of segregation (adjacent I, II or alternate) was observed. Recombination within the interstitial pairing segments of the chromosomes involved in the translocation allowed us to calculate crossover frequencies in ovulated oocytes. For both the large 24 and the small 42 marker chromosomes, this frequency was higher in old than in young T(2;4)1Gö/+ females. Our data do not support the production line hypothesis of Henderson and Edwards (1968) which claims that chiasma frequency in oocytes decreases with maternal age.  相似文献   

11.
The present work reports the observations on numerical and structural changes in 7 embryonic cell lines, derived from two stocks marked by reciprocal and heterozygous translocations between the Y and chromosome 3. Karyological analyses were carried out by using different techniques, such as Q-, C- banding and autoradiography, capable to define the distribution of heterochromatin and to identify individual chromosomes. These procedures, considering also synaptic prophases, provided characteristics for distinguishing each line, besides some very similar features common to all cells. Of particular interest was the appearance of various unusual chromosome morphological variants, characterized by centromeric and interstitial or terminal displaced heterochromatic segments, observed never before in Drosophila lines, whereas the original translocation was not found. Moreover, cell lines were found which had been exclusively polyploid since their establishment, irrespective of the length of their quiescence period. These observations confirm previous findings. The probable origin of the marker chromosomes, as well as a possible correlation between the chromosomal constitution of the cultured cells and the original parental karyotype, are discussed.This paper is dedicated to Prof. Dr. Hans Bauer for his 75th anniversary  相似文献   

12.
In order to investigate wheter colchicine affects crossing-over, rye anthers of an inbred line of rye forming bridges and fragments at anaphase I produced by erroneous chiasmata, and anthers of plants heterozygous for a conspicuous heterochromatin band, were cultured in a medium with colchicine. Anthers planted at zygotene did not show bridges at AI in the inbred line. In the heterozygotes no difference between associated chromatids in respect to the heterochromatin band, resulting from crossingover, were observed. In anthers planted at pachytene both bridges and chromosomes showing difference between associated chromatids were observed at a stage equivalent to AI with the same frequency as in anaphase I cells of untreated anthers. This demonstrates that crossing-over or a prerequisite to crossing-over is established at zygotene, and also that absence of chiasmate association at later stages is not due to precocious slipping off of chiasmata.  相似文献   

13.
The meiosis of an individual of the species Euchorthippus pulvinatus gallicus heterozygous for a reciprocal translocation involving chromosomes 3 and 6 has been analysed using the Giemsa C-banding technique. It is concluded that: (i) Chiasma interference in the quadrivalent seems to act only at the arm level. There is no interference across the translocation break point. No interchromosomal chiasma interference could be demonstrated, (ii) The results concerning the co-orientation of the quadrivalent suggest that the length of the chromosomal segments between two adjacent centromeres at metaphase I is related with their orientation behaviour.  相似文献   

14.
Three types of mitotic recombination can be studied in Aspergillus nidulans and Saccharomyces cerevisiae: (1) The classical type of reciprocal mitotic crossing-over which can be detected when it occurs between non-sister chromatids at the four-strand stage followed by co-segregation of a crossing-over and a non-crossing-over chromatid in the subsequent mitotic division. Consequently, mitotic crossing-over reflects cellular responses to primary genetic damage in the G2 phase of the cell cycle. (2) Mitotic gene conversion is a unidirectional event of a localized transfer of genetic information between non-sister chromatids which in yeast can extend to segments of up to 18 cM and even beyond 22 cM in Aspergillus nidulans. Mitotic gene conversion can also occur between unreplicated chromatids and lead to the expression of the newly created genotype without any need for a subsequent mitotic cell division. It reflects a cellular response in G1. (3) Mitotic sister-strand gene conversion can be studied in a recently constructed strain with the same technical ease as classical non-sister chromatid gene conversion. It can be induced by chemicals which do not induce mutation in the Salmonella system and non-sister chromatid gene conversion. Mitotic segregation in Saccharomyces cerevisiae results almost exclusively from crossing-over and gene conversion whereas mitotic chromosomal malsegregation contributes only very little. In contrast to this, in Aspergillus nidulans, both processes contribute considerably so that mitotic segregants always have to be tested for their mechanistic origin.  相似文献   

15.
Etta K?fer 《Genetics》1976,82(4):605-627
To analyze mitotic recombination in translocation heterozygotes of A. nidulans two sets of well-marked diploids were constructed, homo- or heterozygous for the reciprocal translocations T1(IL;VIIR) or T2(IL;VIIIR) and heterozygous for selective markers on IL. It was found that from all translocation heterozygotes some of the expected mitotic crossover types could be selected. Such crossovers are monosomic for one translocated segment and trisomic for the other and recovery depends on the relative viabilities of these unbalanced types. The obtained segregants show characteristically reduced growth rates and conidiation dependent on sizes and types of mono- and trisomic segments, and all spontaneously produce normal diploid sectors. Such secondary diploid types either arose in one step of compensating crossing over in the other involved arm, or—more conspicuously—in two steps of nondisjunction via a trisomic intermediate.—In both of the analyzed translocations the segments translocated to IL were extremely long, while those translocated from IL were relatively short. The break in I for T1(I;VII) was located distal to the main selective marker in IL, while that of T2(I;VIII) had been mapped proximal but closely linked to it. Therefore, as expected, the selected primary crossover from the two diploids with T2( I;VIII) in coupling or in repulsion to the selective marker, showed the same chromosomal imbalance and poor growth. These could however be distinguished visually because they spontaneously produced different trisomic intermediates in the next step, in accordance with the different arrangement of the aneuploid segments. On the other hand, from diploids heterozygous for T1( I;VII) mitotic crossovers could only be selected when the selective markers were in coupling with the translocation; these crossovers were relatively well-growing and produced frequent secondary segregants of the expected trisomic, 2n+VII, type. For both translocations it was impossible to recover the reciprocal crossover types (which would be trisomic for the distal segments of I and monosomic for most of groups VII or VIII) presumably because these were too inviable to form conidia.—In addition to the selected segregants of expected types a variety of unexpected ones were isolated. The conditions of selection used favour visual detection of aneuploid types, even if these produce only a few conidial heads and are not at a selective advantage. For T2(I;VIII) these "non-selected" unbalanced segregants were mainly "reciprocal" crossovers of the same phenotype and imbalance as the selected ones. For T1(I;VII) two quite different types were obtained, both possibly originating with loss of the small VII–I translocation chromosome. One was isolated when the selective marker in repulsion to T1(I;VII) was used and, without being homo- or hemizygous for the selective marker, it produced stable sectors homozygous for this marker. The other was obtained from both coupling and repulsion diploids and showed a near-diploid genotype; it produced practically only haploid stable sectors of the type expected from monosomics, 2n–1 for the short translocation chromosome.  相似文献   

16.
In order to clarify the relationship between meiotic pairing and progress of spermatogenesis, an analysis of male meiotic pairing was carried out in four reciprocal translocation heterozygotes and two double heterozygotes for two semi-identical reciprocal translocations. The reciprocal translocations were chosen to range from fertility (T70H/+) through almost complete sterility (T31H/+) to complete sterility (T32H/+, T42/H+). If meiotic pairing in the translocation multivalent was incomplete, it concerned terminal or probably more often proximal chromosome segments (Chain IV). If both segments failed to pair the multivalent symbol is Chain III+I. Complete pairing is symbolized by Ring IV. To contrast and complement observations of this type, the double heterozygotes were introduced. Males of this type in theory possess two heteromorphic bivalents with a central area of incomplete meiotic pairing (loop formation). Of the T70H/T1Wa double heterozygotes, 36% of the males are capable of inducing at least one decidual reaction in two females whereas for T26H/T2Wa, 79% of the males can do so. For the reciprocal translocations, it was found that proximity of the multivalent to the sex bivalent during pachytene increased in the order Ring IV, Chain IV, Chain III+I. The degree of spermatogenic impairment as measured from cell counts in histological sections and tubular whole mounts, is positively related to the frequency of proximity between the sex chromosomes and the translocation multivalent and thus to lack of meiotic pairing within the multivalent. The meiotic pairing analysis of the double heterozygotes yielded the following findings. For the long heteromorphic bivalents a true loop was never seen in T70H/T1Wa and only rarely observed in T26H/T2Wa. Small marker bivalents of both types were usually recognizable by the following criteria: (i) pairing confined to distal or proximal segments, (ii) both distal and proximal segments pairing and loop formation and (iii) pairing covering the entire length of both homologues but the longer one often with a thickened lateral element. The same positive correlation between the absence of pairing (proximal, distal or central) and the proximity of the small marker bivalent synaptonemal complex to the sex bivalent has been found as for unpaired segments within reciprocal translocation multivalents. One unexpected finding was the occurrence of diploid spermatids and spermatozoa especially in T32H/+ males (70–91%) but also in T31H/+ (3–39%).  相似文献   

17.
A reciprocal translocation between chromosomes 11 and 22 is a site-specific translocation that has been seen in many families with no common ancestry. This translocation is of particular interest because balanced carriers have a 0.7–3.7% risk of having children with the supernumerary der(22), resulting from a 3:1 segregation. We have used a three color fluorescence in situ hybridization (FISH) with specific DNA probes to determine the chromosome segregation pattern of a male carrier of a translocation t(11;22)(q23;q11). The probes selected included a centromeric marker for chromosome 11, a marker closely linked to the centromere of chromosome 22, and a third probe distal to the translocation breakpoint of chromosome 22. The results showed that 3 : 1 segregation is preferential in this patient, with 40.1% of spermatozoa belonging to this segregation type. Alternate segregation followed with 27.4% of analyzed spermatozoa; 17.6% resulted from adjacent 1 and 12.5% resulted from adjacent 2 segregation. We detected 0.5% of presumably diploid spermatozoa. Complementary adjacent 1 products were observed at statistically different frequencies (P = 0.02). Complementary adjacent 2 products without recombination in the interstitial segments were also seen at different frequencies (P = 0.002). In 3 : 1 segregation, the products containing one chromosome were observed more frequently than those with three chromosomes (P = 0.0001). The 24,+der(22) gamete was seen more frequently than all of the other gametes combined which had 24 chromosomes resulting from 3 : 1 segregation. The results of this study demonstrate that in this t(11;22) carrier, 3 : 1 segregation is preferential but not exclusive. Received: 9 December 1998 / Accepted: 1 March 1999  相似文献   

18.
Meiotic segregation, recombination, and aneuploidy was assessed for sperm from a t(1;10)(p22.1;q22.3) reciprocal translocation carrier, by use of two multicolor FISH methods. The first method utilized three DNA probes (a telomeric and a centromeric probe on chromosome 1 plus a centromeric probe on chromosome 10) to analyze segregation patterns, in sperm, of the chromosomes involved in the translocation. The aggregate frequency of sperm products from alternate and adjacent I segregation was 90.5%, and the total frequency of normal and chromosomally balanced sperm was 48.1%. The frequencies of sperm products from adjacent II segregation and from 3:1 segregation were 4.9% and 3.9%, respectively. Reciprocal sperm products from adjacent I segregation deviated significantly from the expected 1:1 ratio (P < .0001). Our assay allowed us to evaluate recombination events in the interstitial segments at adjacent II segregation. The frequencies of sperm products resulting from interstitial recombination in chromosome 10 were significantly higher than those resulting from interstitial recombination in chromosome 1 (P < .006). No evidence of an interchromosomal effect on aneuploidy was found by use of a second FISH method that simultaneously utilized four chromosome-specific DNA probes to quantify the frequencies of aneuploid sperm for chromosomes X, Y, 18, and 21. However, a significant higher frequency of diploid sperm was detected in the translocation carrier than was detected in chromosomally normal and healthy controls. This study illustrates the advantages of multicolor FISH for assessment of the reproductive risk associated with translocation carriers and for investigation of the mechanisms of meiotic segregation of chromosomes.  相似文献   

19.
Chinese hamster stocks with various structurally abnormal chromosomes have been produced by X irradiation. Among these stocks, 18 with various reciprocal translocations were used to investigate the participation of unbalanced gametes in fertilization and the development of unbalanced embryos. Among males as well as females heterozygous for the same translocation, there is no difference in the frequency of each disjunctional class. The participation of chromosomally unbalanced gametes in fertilization was investigated by chromosomal analysis of meiotic cells in heterozygotes for the 18 reciprocal translocations and pronuclei of fertilized ova obtained from crossing these heterozygotes. Compared with the expected frequencies from MII scoring, the frequencies of male pronuclei having a common deficiency of chromosome 1 (1q17-->1q42) or chromosome 3 (3p23-->3q31) decreased significantly in one-cell embryos. However, the frequencies of male pronuclei with other abnormalities were all consistent with those expected from MII scoring. In contrast, the frequencies of female pronuclei with any karyotype including the same abnormalities as those decreased in male pronuclei from the translocation heterozygotes were all consistent with those estimated from MII scoring. These results revealed clearly that most gametes with nullisomies as well as disomies for any chromosomal segments may participate in fertilization, whereas only male gametes nullisomic for certain segments of chromosomes 1 and 3 failed to participate in fertilization. The zygotic selection of chromosomal imbalance was also investigated by direct chromosomal and morphological analyses of preimplantation embryos from crosses between karyotypically normal females and male heterozygotes from the 18 stocks with various reciprocal translocations. These analyses revealed that some embryos were arrested in development at the two-cell stage. The karyotype of these two-cell embryos had a common deficiency in a segment of chromosome 1 or chromosome 2. Embryos with partial monosomy including chromosomes 1, 3, 4 and 5 showed arrested development at four- to eight-cell stages. Among day 4 embryos, some chromosomally unbalanced embryos, mainly with a deficiency of segments of chromosomes 1p, 1q, 2q, 5q, 7q and 8, had fewer blastomeres than karyotypically normal and balanced embryos. The homology between Chinese hamster and mouse chromosomes relating to abnormal embryogenesis at early stages has been partially confirmed from reported maps of chromosomes. The Chinese hamster is useful for further cytogenetic studies during the stages of meiosis and early embryogenesis.  相似文献   

20.
A. Weith  W. Traut 《Chromosoma》1986,94(2):125-131
Microspread pachytene nuclei of wild-type and W chromosome mutants of the mealmoth Ephestia kuehniella were used to study synaptonemal complex (SC) formation. In structurally heterozygous bivalents, axial elements of considerable length differences were brought to the same length by synaptic adjustment. The adjustment length was a compromise between the mutant and the wildtype homologue length in a structural heterozygote of a W chromosome-autosome translocation, T(A; W). The translocated non-homologous W segment really participated in SC formation as could be seen from the W chromosomal heterochromatin, used as a cytogenetic marker. Pachytene pairing of the wild-type W-Z bivalent extended from about two-thirds to the full length of the W chromosome, though from cytogenetic and genetic evidence W and Z are largely — if not completely — non-homologous. Nonhomologous pairing was even more conspicuous in sex chromosome bivalents containing a deleted W chromosome, Df(W). In one of the pairing configurations the halves of the Z chromosome were synapsed to either side of the Df(W). Thus, one side was pairing with the Df(W) in reversed order. The pairing behavior of the W with homologous chromosome segments was tested by introducing supernumerary W segments via the T(A; W) translocation. Pairing between the W and the translocated homologous W segment never occurred, whereas the Z frequently synapsed with it. Even in T(A; W) homozygotes, pairing between the two translocated W segments was not regularly found while the autosomal parts of the translocation chromosomes were always completely paired. Homologous chromosomes and the ability to form an SC are not sufficient for pairing initiation. Specific loci or sequences are postulated for this function. They are either absent from the W chromosome or are present in only low concentrations.  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