Ultrastructural observations of lethal yellow (Ay/Ay) mouse embryos

Development ◽  
1976 ◽  
Vol 35 (1) ◽  
pp. 73-80
Author(s):  
Patricia G. Calarco ◽  
Roger A. Pedersen

Ay/Ay embryos were identified by the presence of large excluded blastomeres (Pedersen, 1974) and examined cytologically and ultrastructurally. Cell organelles, inclusions and junctions in the excluded blastomeres were compared with those of non-excluded cells of Ay/Ay embryos and control embryos. Excluded blastomeres always had the fine structural characteristics of earlier developmental stages and may have arrested at the 4- to 8-cell stage or slightly later. Interior cells (inner cell mass) were observed in all mutant blastocysts. Nonexcluded cells of Ay/Ay embryos were normal until degenerative changes appear in the late blastocyst stage. The mode of action of the +Ay gene was not determined, but evidence from this study and others indicates that the effects of +Ay gene action occur over a wide range of time in early cleavage and implantation.

Development ◽  
1992 ◽  
Vol 115 (1) ◽  
pp. 305-312
Author(s):  
M. Aghayan ◽  
L.V. Rao ◽  
R.M. Smith ◽  
L. Jarett ◽  
M.J. Charron ◽  
...  

Two general mechanisms mediate glucose transport, one is a sodium-coupled glucose transporter found in the apical border of intestinal and kidney epithelia, while the other is a sodium-independent transport system. Of the latter, several facilitated transporters have been identified, including GLUT1 (erythrocyte/brain), GLUT2 (liver) and GLUT4 (adipose/muscle) isoforms. In this study, we used Western-blot analysis and high resolution immunoelectron microscopy (IEM) to investigate the stage-related expression and cellular localization of GLUT1, 2 and 4. The Western blot results demonstrate that GLUT1 is detectable in the oocyte and throughout preimplantation development. GLUT2 isoforms were not detectable until the blastocyst stage, while the GLUT4 isoform was undetectable in the oocyte through blastocyst stages. The present findings confirm previous studies at the molecular level which demonstrated that mRNAs encoding the same GLUT isoforms are detectable at corresponding developmental stages. GLUT1 and GLUT2 display different cellular distributions at the blastocyst stage as shown by IEM studies. GLUT1 has a widespread distribution in both trophectoderm and inner cell mass cells, while GLUT2 is located on trophectoderm membranes facing the blastocyst cavity. This observation suggests a different functional significance for these isoforms during mouse preimplantation development.


2013 ◽  
Vol 25 (1) ◽  
pp. 226 ◽  
Author(s):  
E. P. López-Damián ◽  
T. Fiordelisio ◽  
M. A. Lammoglia ◽  
M. Alarcón ◽  
M. Asprón ◽  
...  

Accurate evaluation of bovine embryos for assessing developmental stage and quality is critical to the success of any embryo transfer program. However, this evaluation process has been reported to be highly subjective in Bos indicus (BI) and can vary as much as 23% compared with that of Bos taurus (BT). These differences in assessment may be related to the quantity of lipid droplets (LD) within the embryo, which has been shown to have a negative effect in cryopreserving embryos. The aim of the present study was to characterize the number and size of LD in different developmental stages of fresh embryos from BI and BT and to compare LD across the three different embryo quality grades (1 = excellent or good, 2 = fair, and 3 = poor). Nonsurgical embryo collection was performed 7 days post-insemination in 10 BI and 10 BT females. Forty-eight embryos were evaluated for stage and grade using stereoscopic microscopy, processed for transmission electron microscopy, and stained with Nile red. Digitalized images were analyzed with ImageJ (National Institutes of Health, Bethesda, MD, USA), contour of lipid droplets were designed, and values of perimeter, area, and fluorescence intensity were assessed. Nonparametric statistical analysis (Mann–Whitney) was utilized. There was no difference in LD number for BT or BI for morulae and blastocyst; however, BI morulae presented larger LD compared with blastocyst stage embryos (286 µm2 v. 223 µm2; P < 0.05). Likewise, BI TF cells had more LD compared with inner cell mass (ICM) cells (48 v. 36; P < 0.05). BT TF cells exhibited larger LD compared with ICM cells (149 µm2 v. 128 µm2; P < 0.05), while BI embryos exhibited a larger area of LD in the ICM compared with the TF (591 µm2 v. 472 µm2; P < 0.05). In all embryos, BI contained more lipid droplets than BT (78 v. 49; P < 0.05). Across all quality grades (good, fair, and poor) there was no difference in the number of LD in BT embryos; however, BI grade-3 embryos presented more LD than grade-1 (36 v. 25). BT embryos LD were larger than BI LD (907 µm2 v. 625 µm2; P < 0.05). Fluorescence images showed higher arbitrary units of fluorescence (auf) for LD in BI. Compared with BT embryos (386 auf v. 280 auf; P < 0.05). These results suggest that BI embryos contain more and smaller LD than BT embryos and the LD described for BI embryo quality grade 1 are larger than those of quality grades 2 and 3, and even though the number of LD in morulae and blastocyst stage embryos are not different LD size is reduced as development occurs. Research funding provided by UNAM-DGAPA-PAPIIT IN200810.


2013 ◽  
Vol 25 (1) ◽  
pp. 255
Author(s):  
C. Sauvegarde ◽  
D. Paul ◽  
R. Rezsohazy ◽  
I. Donnay

Hox genes encode for homeodomain transcription factors well known to be involved in developmental control after gastrulation. However, the expression of some of these genes has been detected during oocyte maturation and early embryo development. An interesting expression profile has been obtained for HOXB9 in the bovine (Paul et al. 2011 Mol. Reprod. Dev. 78, 436): its relative expression increases between the immature oocyte and the zygote, further increases at the 5- to 8-cell stage to peak at the morula stage before decreasing at the blastocyst stage. The main objective of this work is to establish the HOXB9 protein profile from the immature oocyte to the blastocyst in the bovine. Bovine embryos were produced in vitro from immature oocytes obtained from slaughterhouse ovaries. Embryos were collected at the following stages: immature oocyte, mature oocyte, zygote (18 h post-insemination, hpi), 2-cell (26 hpi), 5 to 8 cell (48 hpi), 9 to 16 cell (96 hpi), morula (120 hpi), and blastocyst (180 hpi). The presence and distribution of HOXB9 proteins were detected by whole-mount immunofluorescence followed by confocal microscopy using an anti-human HOXB9 polyclonal antibody directed against a sequence showing 100% homology with the bovine protein. Its specificity to the bovine protein was controlled by Western blot on total protein extract from the bovine uterus and revealed, among a few bands of weak intensities, 2 bands of high intensity corresponding to the expected size. Oocytes or embryos were fixed and incubated overnight with rabbit anti-HOXB9 (Sigma, St. Louis, MO, USA) and mouse anti-E-cadherin (BD Biosciences, Franklin Lakes, NJ, USA) primary antibodies and then for 1 h with goat anti-rabbit Alexafluor 555 conjugated (Cell Signaling Technology, Beverly, MA, USA) and goat anti-mouse FITC-conjugated (Santa Cruz Biotechnology Inc., Santa Cruz, CA, USA) secondary antibodies. Embryos were then mounted in Vectashield containing DAPI. HOXB9 is detected from the immature oocyte to the blastocyst stage. At the immature oocyte stage, it is mainly localised in the germinal vesicle with a weak signal in the cytoplasm. At the mature oocyte stage, HOXB9 labelling is present in the cytoplasm. At the zygote stage, a stronger immunoreactivity is observed in the pronuclei than in the cytoplasm. From the 2-cell stage to the morula stage, the presence of HOXB9 is also more important in the nuclei than in the cytoplasm. HOXB9 is also observed at the blastocyst stage where it is localised in the nuclei of the trophectoderm cells, whereas an inconstant or weaker labelling is observed in the inner cell mass cells. In conclusion, we have shown for the first time the presence of the HOXB9 protein throughout early bovine embryo development. The results obtained suggest the presence of the maternal HOXB9 protein because it is already detected before the maternal to embryonic transition that occurs during the fourth cell cycle in the bovine. Finally, the pattern obtained at the blastocyst stage suggests a differential role of HOXB9 in the inner cell mass and trophectoderm cells. C. Sauvegarde holds a FRIA PhD grant from the Fonds National de la Recherche Scientifique (Belgium).


1995 ◽  
Vol 29 (3) ◽  
pp. 320-326 ◽  
Author(s):  
Th. Rülicke ◽  
P. Autenried

Approximately 18% of cryopreserved 2-cell mouse embryos of 26 different batches showed various degrees of morphological damage after the freeze-thaw process. Normal and damaged morphology were assessed by light microscopy and the ability of an embryo to develop in vitro to a blastocyst, or to develop to term, after transfer to foster mothers. Using vital stains such as Fluorescein-diacetate (FDA) and 4',6-Diamidino-2-Phenylindole (DAPI) it was found that in approximately 82% of the cases, both of the 2 blastomeres of the cryopreserved embryos survived the freeze-thaw process; in 10% only one cell survived the process; and in 8% none survived. Normally, only intact 2-cell embryos are considered for transfer. Here it was shown that over 60% of the partially damaged embryos developed in vitro to the blastocyst stage and, of those, 26% developed to term after transfer to suitable foster mothers. Although the inner cell mass (ICM) appeared to remain smaller during culture after the transfer of partially damaged 2-cell stage embryos, no difference during gestation period was found compared with intact embryos.


Zygote ◽  
2009 ◽  
Vol 17 (2) ◽  
pp. 125-130 ◽  
Author(s):  
Lei Lei ◽  
Na Guan ◽  
Yan-Ning Xu ◽  
Qing-Hua Zhang ◽  
Jing-Ling Shen ◽  
...  

SummaryPolyploid mouse embryos are important models for understanding the mechanisms of cleavage and preimplantation development in mammals. In this study, hexaploid (6n) mouse embryos were produced by the electrofusion of blastomeres from diploid (2n) and tetraploid (4n) embryos at the 2-cell stage. Furthermore, the developmental pattern of hexaploid embryos was evaluated by blastocyst rate, cell number, karyotype analysis, cytoskeleton staining and Oct-4 immunofluorescence. The results showed that 72.7% of the hexaploid embryos were able to develop to the blastocyst stage, which is a lower number than that found with normal diploid embryos (98.0%, p < 0.05). The cell number in hexaploid blastocyst was 12.3 ± 2.0, which was less than that found in diploid or tetraploid blastocysts (41.2 ± 7.2; 18.4 ± 3.5). Karyotype analysis confirmed that the number of chromosomes in hexaploid embryos was 120. β-Tubulin and Oct-4 immunofluorescence indicated that the hexaploid blastocysts were nearly lacking inner cell mass (ICM), but some blastomeres did show Oct-4-positive expression.


eLife ◽  
2018 ◽  
Vol 7 ◽  
Author(s):  
Jan J Zylicz ◽  
Maud Borensztein ◽  
Frederick CK Wong ◽  
Yun Huang ◽  
Caroline Lee ◽  
...  

Early mouse development is regulated and accompanied by dynamic changes in chromatin modifications, including G9a-mediated histone H3 lysine 9 dimethylation (H3K9me2). Previously, we provided insights into its role in post-implantation development (Zylicz et al., 2015). Here we explore the impact of depleting the maternally inherited G9a in oocytes on development shortly after fertilisation. We show that G9a accumulates typically at 4 to 8 cell stage to promote timely repression of a subset of 4 cell stage-specific genes. Loss of maternal inheritance of G9a disrupts the gene regulatory network resulting in developmental delay and destabilisation of inner cell mass lineages by the late blastocyst stage. Our results indicate a vital role of this maternally inherited epigenetic regulator in creating conducive conditions for developmental progression and on cell fate choices.


Reproduction ◽  
2017 ◽  
Vol 154 (6) ◽  
pp. 807-814
Author(s):  
L P Sepulveda-Rincon ◽  
N Islam ◽  
P Marsters ◽  
B K Campbell ◽  
N Beaujean ◽  
...  

It has been suggested that first embryo cleavage can be related with the embryonic–abembryonic axis at blastocyst stage in mice. Thus, cells of the 2-cell embryo might be already biased to form the inner cell mass or trophectoderm. This study was conducted to observe the possible effects of embryo biopsy on cell allocation patterns during embryo preimplantation in two different mouse strains and the effects of these patterns on further development. First, one blastomere of the 2-cell embryo was injected with a lipophilic tracer and cell allocation patterns were observed at blastocyst stage. Blastocysts were classified into orthogonal, deviant or random pattern. For the first experiment, embryos were biopsied at 8-cell stage and total cell counts (TCC) were annotated. Furthermore, non-biopsied blastocysts were transferred into foster mothers. Then, pups and their organs were weighed two weeks after birth. Random pattern was significantly recurrent (≈60%), against orthogonal (<22%) and deviant (<22%) patterns among groups. These patterns were not affected by biopsy procedure. However, TCC on deviant embryos were reduced after biopsy. Moreover, no differences were found between patterns for implantation rates, litter size, live offspring and organ weights (lungs, liver, pancreas and spleen). However, deviant pups presented heavier hearts and orthogonal pups presented lighter kidneys among the group. In conclusion, these results suggest that single blastomere removal does not disturb cell allocation patterns during pre-implantation. Nonetheless, the results suggest that embryos following different cell allocation patterns present different coping mechanisms against in vitro manipulations and further development might be altered.


Cells ◽  
2021 ◽  
Vol 10 (9) ◽  
pp. 2232 ◽  
Author(s):  
Claudia Springer ◽  
Valeri Zakhartchenko ◽  
Eckhard Wolf ◽  
Kilian Simmet

The role of the pluripotency factor NANOG during the second embryonic lineage differentiation has been studied extensively in mouse, although species-specific differences exist. To elucidate the role of NANOG in an alternative model organism, we knocked out NANOG in fibroblast cells and produced bovine NANOG-knockout (KO) embryos via somatic cell nuclear transfer (SCNT). At day 8, NANOG-KO blastocysts showed a decreased total cell number when compared to controls from SCNT (NT Ctrl). The pluripotency factors OCT4 and SOX2 as well as the hypoblast (HB) marker GATA6 were co-expressed in all cells of the inner cell mass (ICM) and, in contrast to mouse Nanog-KO, expression of the late HB marker SOX17 was still present. We blocked the MEK-pathway with a MEK 1/2 inhibitor, and control embryos showed an increase in NANOG positive cells, but SOX17 expressing HB precursor cells were still present. NANOG-KO together with MEK-inhibition was lethal before blastocyst stage, similarly to findings in mouse. Supplementation of exogenous FGF4 to NANOG-KO embryos did not change SOX17 expression in the ICM, unlike mouse Nanog-KO embryos, where missing SOX17 expression was completely rescued by FGF4. We conclude that NANOG mediated FGF/MEK signaling is not required for HB formation in the bovine embryo and that another—so far unknown—pathway regulates HB differentiation.


2004 ◽  
Vol 16 (2) ◽  
pp. 144
Author(s):  
P. Kasinathan ◽  
M.F. Nichols ◽  
J.E. Griffin ◽  
J.M. Robl

Chimeras have been used for investigating fundamental aspects of early embryonic development, and differentiation, and for introducing foreign genes into mammals (Robertson et al., 1986 Nature 323, 445–448; Cibelli et al., 1998 Science 280, 1256–1258). The main objective of this study was to determine if the transfer of blastomeres from in vitro-produced (IVP) embryos into cloned, transchromosomic embryos improved the efficiency of producing transchromosomic calves. Cloned embryos were produced using in vitro-matured bovine oocytes and bovine fetal fibroblasts containing a human artificial chromosome (HAC) (Kuroiwa et al., 2002 Nat Biotechnol 20, 889–894). IVP embryos were produced using standard procedures and blastomeres were harvested at the 8–16 cell stage by removing the zona pellucida with protease. Cloned embryos were randomly divided on Day 4 into two groups. One group received 3–4 IVP blastomeres while a second group served as a control (nonmanipulated cloned embryos). After transferring the blastomeres, the chimeric and cloned embryos were placed in culture (Kasinathan et al., 2001 Biol. Reprod. 64, 1487–1493) and on Day 7 development to the blastocyst stage was evaluated. Grades 1 and 2 embryos were transferred; two each per synchronized recipient. Pregnancy maintenance, calving, and calf survival were evaluated in both groups. Presence of a HAC in live calves was evaluated in both fibroblasts and peripheral blood lymphocytes (PBLs) using FISH analysis. Embryo development to the blastocyst stage, maintenance of pregnancy and number of calves born were analyzed using Chi-square. There were no differences in the rate of blastocyst development at day 7 or establishment of pregnancy at 40d (P&gt;0.05). However, pregnancy rate at 120d, and number of calves that developed to term and were alive at birth (chimera 14/54 and clone 4/90), and at 1 month of age (chimera 13/54 and clone 1/90) were lower (P&lt;0.01) for cloned embryos. The proportion of cells containing an HAC in PBLs, was higher in cloned calves (100%) compared to chimeric calves (26%). The HAC retension rates in PBLs in HAC-positive chimeric and cloned calves were 84% and 95%, respectively. These data indicate that, although the proportion of calves retaining an HAC was lower in chimeras compared to clones, more HAC-positive calves were produced in the chimeric treatment from fewer cloned embryos. We speculate that higher rates of development in the chimeras may be related to the normality of the placenta. Future studies will be required to determine the contribution of the IVP blastomeres to both the inner cell mass and trophectoderm. Therefore, a chimeric approach may be useful for improving the efficiency of producing cloned transchromosomic calves.


2013 ◽  
Vol 25 (1) ◽  
pp. 251
Author(s):  
S. Krebs ◽  
A. Graf ◽  
Z. Valeri ◽  
H. Blum ◽  
E. Wolf

In order to provide a comprehensive view of the transcriptome changes during the earliest stages of bovine development, we sequenced the total RNA content of bovine oocytes, 4-cell, 8-cell, and 16-cell embryos and the inner cell mass and trophoblast envelope of expanded blastocysts on the Illumina Genome Analyzer IIx. For each experiment pools of in vitro matured oocytes from the German Simmental cows were fertilized by sperm of a single bull, and 10 oocytes or embryos per developmental stage were collected to generate total RNA pools used for sequencing. Synthesis of cDNA was initiated directly in the cell lysate in order to avoid any losses during RNA preparation and was random primed in order to capture all RNA species. Amplified cDNA and unstranded sequencing libraries were prepared using kits from Nugen (Ovation RNA-Seq, Nugen, San Carlos, CA, USA). Biological replicates were generated by inseminating the oocytes with sperm from the distant breeds Jersey (n = 3) and Brahman (n = 3). This cross-breeding design allowed tracking of single sequencing reads back to the maternal or paternal genome, where breed-specific SNP are present in the expressed transcripts. The analysis of this dataset resulted in monitoring of zygotic genome activation and parent-specific expression for single transcripts, a catalogue of splicing isoforms, novel transcripts, and non-coding RNAs and differentially expressed genes between the single developmental stages. Using the program DESEqn, 2784 genes showed differential expression between any of the stages at a false discovery rate of 1%. Specifically, we found 200 genes differentially expressed between immature and matured oocytes, 209 genes between matured oocytes and 4-cell embryos, 580 genes between the 4-cell and 8-cell stage, 567 genes between the 8-cell and 16-cell stage, 987 genes between the 16-cell stage and the inner cell mass, and 1569 genes between the 16-cell and the trophoblast. Functional analysis revealed stage-specific functions of the differentially expressed genes. In summary, by fully exploiting the single-nucleotide resolution of the RNA-Seq method, this dataset provides an invaluable resource for the study of zygotic genome activation, imprinting, transcript annotation, and gene expression in the earliest developmental stages of bovine embryos.


Sign in / Sign up

Export Citation Format

Share Document