Judith Kasir's research while affiliated with Hebrew University of Jerusalem and other places

What is this page?


This page lists the scientific contributions of an author, who either does not have a ResearchGate profile, or has not yet added these contributions to their profile.

It was automatically created by ResearchGate to create a record of this author's body of work. We create such pages to advance our goal of creating and maintaining the most comprehensive scientific repository possible. In doing so, we process publicly available (personal) data relating to the author as a member of the scientific community.

If you're a ResearchGate member, you can follow this page to keep up with this author's work.

If you are this author, and you don't want us to display this page anymore, please let us know.

Publications (30)


Figure 1. TSC2 or TSC1 deficiency rescues TOP mRNAs from translational repression in amino acid-starved cells. (A) TSC2 + / + and TSC2 2 / 2 as well as TSC1 + / + and TSC1 2 / 2 MEFs, were amino acid-starved for 16 h ( 2 AA), amino acid-starved and then refed for 2 h, or amino acid staved during the last 16 h of 48 h serum starvation ( 2 AA 2 serum). Subsequently cells were harvested and cytoplasmic extracts were prepared. These extracts were centrifuged through sucrose gradients and divided into polysomal (P) and subpolysomal (S) fractions. RNA from equivalent aliquots of these fractions was analyzed by Northern-blot hybridization with cDNAs for rpL32 mRNA (a TOP mRNA) and actin mRNA (a non TOP mRNA) (in the case of TSC2 also with cDNAs corresponding to rpS6 and tubulin). The radioactive signals were quantified, and the relative translational efficiency (% of the P signal relative to the P + S signals) of each mRNA is numerically presented beneath the autoradiograms as percentage of the mRNA engaged in polysomes. These figures are expressed as an average 6 SEM of the number of determinations in parenthesis, or the average with the individual values in parenthesis, if only two determinations are presented. (B) TSC2 + / + , TSC2 2 / 2 MEFs were untreated or amino acid-starved for 16 h and then harvested. The cytoplasmic proteins were subjected to Western blot analysis using the indicated antibodies. doi:10.1371/journal.pone.0109410.g001 
Figure 2. mTOR mediates amino acid-induced translational activation of TOP mRNAs. (A) Kinetics of the effect of rapamycin on mTORC1 activity. 293 cells were amino acid-starved for 2 h and then refed for the indicated time in the presence or absence of 20 nM rapamycin, after which cells were harvested. The cytoplasmic proteins were subjected to Western blot analysis with anti-rpS6 or anti-Phospho-rpS6 antibodies. The chemiluminescent signals of phospho rpS6 were quantified and normalized to those obtained for rpS6 within the same protein extract. The results are numerically presented relative to those obtained for amino acid-starved cells (time zero), which were arbitrarily set at 1. (B) Kinetics of the effect of rapamycin on polysomal association of TOP mRNAs. HEK293 cells were amino acid-starved for 3 h (time zero), and then refed in the absence (open symbols) or presence (filled symbols) of 20 nM rapamycin (rapa). At the indicated times cells were harvested and cytoplasmic extracts were subjected to polysomal analysis. The percentage of mRNA in polysomes at each time point is presented as an average of at least 2 measurements. (C) HEK293 cells were infected with viruses expressing HcRed (Red) shRNA or mTOR shRNA1. Cells were amino acid-starved for 3 h followed by 3 h amino acid stimulation on day 4 post-infection. The abundance of mTOR and its activity were monitored by Western blot analysis of cytoplasmic proteins with the indicated antibodies. (D) Cytoplasmic extracts from cells described in (C) were subjected to polysomal analysis. (E) and (F) HEK293 were transiently transfected with plasmid-based vectors expressing either wild-type (WT) mTOR or enhanced (En) mTOR. 48 h later cells were amino acid- starved for 3 h and harvested. Cytoplasmic proteins were subject to Western blot analysis (E) and cytoplasmic extracts to polysomal analysis (F). The percentage of mRNA in polysomes is presented as an average 6 SEM of three experiments. doi:10.1371/journal.pone.0109410.g002 
Figure 3. Raptor and rictor are dispensable for translational activation of TOP mRNAs by amino acids. (A) HEK293 cells were infected with viruses expressing Red shRNA, raptor shRNA (Rap) or rictor shRNAs (Ric). The abundance of raptor or rictor, as well the phosphorylation status of direct and indirect substrates of the respective complexes, mTORC1 and mTORC2 (left and right, respectively), was monitored by Western blot analysis. (B) HEK293 cells infected with viruses expressing HcRed, raptor or rictor shRNA were amino acid starved for 3 h ( 2 AA) or starved and then refed for 3 h ( 2 AA R + AA). Cytoplasmic extracts from these cells were subjected to polysomal analysis and the data are presented as described in the legend to Fig. 2F. Numbers above bars are individual values, when only two measurements were performed. doi:10.1371/journal.pone.0109410.g003 
Figure 4. The kinase activity of mTOR is essential for translational control of TOP mRNAs. (A) HEK293 cells were transfected with vectors expressing mTOR-wt, mTOR-rr or mTOR-rr-kd, two days later the cells were amino acid-starved for 3 h followed by 3 h refeeding without or with 20 nM rapamycin. Cytoplasmic proteins derived from the cells were subjected to Western blot analysis with the indicated antibodies. B) Cytoplasmic extracts derived from cells treated as described in (A), were subjected to polysomal analysis. C) HEK293 cells were amino acid-starved for 3 h, or amino acid-starved for 3 h followed by 3 h refeeding without or with 50 nM Torin1. Cytoplasmic proteins derived from the cells were subjected to Western blot analysis with the indicated antibodies. D) Cytoplasmic extracts derived from cells treated as described in (C) were subjected to polysomal analysis and the percentage of mRNA in polysomes is presented as an average 6 SEM of three experiments. doi:10.1371/journal.pone.0109410.g004 
Figure 5. Rapamycin represses the translation of TOP mRNAs in an FKBP-12-dependent fashion. (A) HEK293 cells were amino acid-starved for 3 h and then refed for 3 h in the absence or presence of rapamycin (20 nM), FK506 (20 mM), or both. Cytoplasmic proteins were subjected to Western blot analysis. (B) HEK293 cells were amino 

+6

Reassessment of the role of TSC, mTORC1 and microRNAs in amino acids-meditated translational control of TOP mRNAs
  • Article
  • Full-text available

October 2014

·

455 Reads

·

29 Citations

PLOS ONE

PLOS ONE

·

Judith Kasir

·

Rachel Miloslavski

·

[...]

·

TOP mRNAs encode components of the translational apparatus, and repression of their translation comprises one mechanism, by which cells encountering amino acid deprivation downregulate the biosynthesis of the protein synthesis machinery. This mode of regulation involves TSC as knockout of TSC1 or TSC2 rescued TOP mRNAs translation in amino acid-starved cells. The involvement of mTOR in translational control of TOP mRNAs is demonstrated by the ability of constitutively active mTOR to relieve the translational repression of TOP mRNA upon amino acid deprivation. Consistently, knockdown of this kinase as well as its inhibition by pharmacological means blocked amino acid-induced translational activation of these mRNAs. The signaling of amino acids to TOP mRNAs involves RagB, as overexpression of active RagB derepressed the translation of these mRNAs in amino acid-starved cells. Nonetheless, knockdown of raptor or rictor failed to suppress translational activation of TOP mRNAs by amino acids, suggesting that mTORC1 or mTORC2 plays a minor, if any, role in this mode of regulation. Finally, miR10a has previously been suggested to positively regulate the translation of TOP mRNAs. However, we show here that titration of this microRNA failed to downregulate the basal translation efficiency of TOP mRNAs. Moreover, Drosha knockdown or Dicer knockout, which carries out the first and second processing steps in microRNAs biosynthesis, respectively, failed to block the translational activation of TOP mRNAs by amino acid or serum stimulation. Evidently, these results are questioning the positive role of microRNAs in this mode of regulation.

Download
Share

Figure 1 TOP mRNAs are translationally repressed by oxygen deprivation in an mTOR-dependent fashion. (A) Typical polysomal profiles and its
Figure 3 4E-BP deficiency failed to alleviate the translational repression of TOP mRNAs in oxygen-deprived cells. (A and B) 4E-BP WT and 4E-BP DKO MEFs were either untreated (+) or oxygen starved (2) for 12 h (A), or either untreated (+) or serum starved (2) for 48 h (B). Cytoplasmic proteins from these cells were subjected to western blot analysis with the indicated antibodies. (C) Cells treated as described in A and B were harvested and subjected to polysomal analysis with the indicated probes. (D) 4E-BP DKO MEFs were infected with an empty retroviral vector (EV) pBABE-puro or retroviral vector encoding either wild-type 4E-BP (WT) or 4E-BP 4Ala (4Ala). After selection with puromycin, cells were either untreated or treated with 250 nM Torin 1 for 2 h and their cytoplasmic proteins were subjected to western blot analysis with the indicated antibodies. (E) Cells derived as described in D were harvested and subjected to polysomal analysis with the indicated probes.
Figure 4 Translation of cyclin D3 mRNA, but not those encoding rpL32 and actin, is 4E-BP dependent. (A) Wild-type (WT) and 4E-BP DKO (DKO) MEFs were maintained in 0.5% FBS for 14 h and stimulated with 10% FBS in the absence or presence of 250 nM Torin 1 for 2 h. Levels and the phosphorylation status of the indicated proteins were determined by western blotting. (B2D) WT and DKO MEFs were treated as in A and the cytoplasmic extract was size fractionated by sucrose gradient centrifugation (Supplementary Figure S5 for polysomal profiles). Distribution of rpL32 (B), cyclin D3 (C), and b-actin (D) mRNAs among heavy (H, ≥4 ribosomes) and light (L, 2 to 3 ribosomes) polysomal fractions as well as subpolysomal fraction (S) was monitored by reverse transcriptase-quantitative PCR (RT-qPCR). Values are expressed as a percentage of total mRNA loaded onto the sucrose gradient (input). Data are presented as mean + SD (two independent biological replicates). Each biological replicate was carried out in three technical replicates.
Figure 5 Hyperphosphorylation of 4E-BP failed to derepress TOP mRNA translation. (A and B) Dicer +/+ and Dicer 2/2 hemangiosarcoma cells were either untreated (+), oxygen starved (2) for 12 h (A) or treated with 50 nM Torin 1 (+) for 3 h (B) and cytoplasmic proteins were subjected to western blot analysis with the indicated antibodies. (C) Cells treated as described in A and B as well as cells deprived of oxygen for 12 h and then resupplied with oxygen for 3 h were harvested and subjected to polysomal analysis. (D and E) Untreated L2 lymphoblastoids were subjected to western blot (D) and polysomal (E) analyses.
Fig. S4  
Oxygen sufficiency controls TOP mRNA translation via the TSC-Rheb-mTOR pathway in a 4E-BP-independent manner

March 2014

·

293 Reads

·

85 Citations

Journal of Molecular Cell Biology

Cells encountering hypoxic stress conserve resources and energy by downregulating the protein synthesis. Here we demonstrate that one mechanism in this response is the translational repression of TOP mRNA that encode components of the translational apparatus. This mode of regulation involves TSC and Rheb, as knockout of TSC1 or TSC2 or overexpression of Rheb rescued TOP mRNAs translation in oxygen-deprived cells. Stress-induced translational repression of these mRNAs closely correlates with the hypophosphorylated state of 4E-BP, a translational repressor. However, a series of 4E-BP loss- and gain-of-function experiments disprove a cause-and-effect relationship between the phosphorylation status of 4E-BP and the translational repression of TOP mRNAs under oxygen or growth factor deprivation. Furthermore, the repressive effect of anoxia is similar to that attained by the very efficient inhibition of mTOR activity by Torin 1, but much more pronounced than raptor or rictor knockout. Likewise, deficiency of raptor or rictor, even though it mildly downregulated basal translation efficiency of TOP mRNAs, failed to suppress the oxygen-mediated translational activation of TOP mRNAs. Finally, co-knockdown of TIA-1 and TIAR, two RNA-binding proteins previously implicated in translational repression of TOP mRNAs in amino acid-starved cells, failed to relieve TOP mRNA translation under other stress conditions. Thus, the nature of the proximal translational regulator of TOP mRNAs remains elusive.





FIG. 3. Overexpression of Rheb can relieve translational repression of TOP mRNAs in mitotically arrested cells. (A) HEK 293 cells were transiently transfected with expression vectors encoding FLAG- Rheb (WT) or Myc-Rheb-5A, and 24 h later they were serum starved for 36 h. Subsequently, cells were harvested and subjected to polysomal analysis. (B) HEK 293 cells were transiently transfected with expression vectors encoding FLAG-Rheb (WT) or Myc-Rheb (5A), and 24 h later they were serum starved for 36 h. Subsequently, cells were harvested and cytoplasmic proteins were subjected to Western blot analysis using the indicated antibodies. 
FIG. 4. Knockdown of mTOR downregulates the translation efficiency of TOP mRNAs in insulin-treated cells. (A) HEK 293 cells were infected with viruses expressing Red shRNA (shRed) or mTOR shRNA (shmTOR). Cells were 48-h serum starved and then insulin stimulated for 3 h. The abundance of mTOR and its activity were monitored by Western blot analysis of cytoplasmic proteins with the indicated antibodies. The relative abundance of mTOR was normalized to that of actin, whereas the relative abundance of phospho rpS6 (P-rpS6) was normalized to that of rpS6. The results are numerically presented relative to those obtained for infected cells that expressed Red shRNA, which were arbitrarily set at 1. (B) Cytoplasmic extracts from cells described in panel A were subjected to polysomal analysis. 
FIG. 5. The translation efficiency of TOP mRNAs does not rely on insulin-treated cells. (A) HEK 293 cells were infected with viruses expressing Red shRNA (shRed) or raptor shRNA (shRaptor). Cells were treated as described in the legend to Fig. 4. The abundance of raptor- and mTOR-dependent proteins was monitored by Western blot analysis with the indicated antibodies. The relative abundance of raptor, phospho S6K1(Thr389) and phospho Akt(Ser473) was normalized to that of actin, whereas the relative abundance of phospho rpS6(Ser235/236) was normalized to that of rpS6. The results are numerically presented relative to those obtained with Red shRNA-expressing cells that were arbitrarily set at 1. (B) Cytoplasmic extracts from cells described in panel A were subjected to polysomal analysis. (C) iRapKO cells were either untreated ( Ϫ ) or treated ( ϩ ) with tamoxifen for 4 days and were then further incubated for 3 h in the absence ( Ϫ ) or presence ( ϩ ) of rapamycin. The relative abundance of the indicated proteins was assessed as described in panel A with the exception that results are numerically presented relative to those obtained with cells untreated with either tamoxifen or rapamycin. (D) iRapKO cells were either treated or untreated with tamoxifen (for 4 days) and subsequently were harvested and subjected to polysomal analysis. 
FIG. 7. Rapamycin represses the translation of TOP mRNAs in an FKBP-12-dependent and raptor-sensitive fashion. (A) 3T3-L1 preadipocytes were either untreated (Control), treated for 3 h with rapamycin, FK506, or both drugs, or treated for 0.5 h with LY294002. Subsequently, they were harvested and subjected to polysomal analysis. (B) 3T3-L1 preadipocytes were treated for 3 h with rapamycin, FK506, or both drugs. The cytoplasmic proteins were subjected to Western blot analysis. (C) HEK 293 cells were transiently transfected with pcDNA3-6myc-hFKBP12 (myc-FKBP12) or pcDNA3-6myc (myc tag). Forty-eight hours later, cells were incubated without ( Ϫ ) or with ( ϩ ) rapamycin and were subsequently harvested and subjected to polysomal analysis. 
The TSC-mTOR Pathway Mediates Translational Activation of TOP mRNAs by Insulin Largely in a Raptor- or Rictor-Independent Manner

February 2009

·

146 Reads

·

121 Citations

The stimulatory effect of insulin on protein synthesis is due to its ability to activate various translation factors. We now show that insulin can increase protein synthesis capacity also by translational activation of TOP mRNAs encoding various components of the translation machinery. This translational activation involves the tuberous sclerosis complex (TSC), as the knockout of TSC1 or TSC2 rescues TOP mRNAs from translational repression in mitotically arrested cells. Similar results were obtained upon overexpression of Rheb, an immediate TSC1-TSC2 target. The role of mTOR, a downstream effector of Rheb, in translational control of TOP mRNAs has been extensively studied, albeit with conflicting results. Even though rapamycin fully blocks mTOR complex 1 (mTORC1) kinase activity, the response of TOP mRNAs to this drug varies from complete resistance to high sensitivity. Here we show that mTOR knockdown blunts the translation efficiency of TOP mRNAs in insulin-treated cells, thus unequivocally establishing a role for mTOR in this mode of regulation. However, knockout of the raptor or rictor gene has only a slight effect on the translation efficiency of these mRNAs, implying that mTOR exerts its effect on TOP mRNAs through a novel pathway with a minor, if any, contribution of the canonical mTOR complexes mTORC1 and mTORC2. This conclusion is further supported by the observation that raptor knockout renders the translation of TOP mRNAs rapamycin hypersensitive.


Mice Deficient in Ribosomal Protein S6 Phosphorylation Suffer from Muscle Weakness that Reflects a Growth Defect and Energy Deficit

February 2009

·

180 Reads

·

96 Citations

PLOS ONE

PLOS ONE

Background: Mice, whose ribosomal protein S6 cannot be phosphorylated due to replacement of all five phosphorylatable serine residues by alanines (rpS6(P-/-)), are viable and fertile. However, phenotypic characterization of these mice and embryo fibroblasts derived from them, has established the role of these modifications in the regulation of the size of several cell types, as well as pancreatic beta-cell function and glucose homeostasis. A relatively passive behavior of these mice has raised the possibility that they suffer from muscle weakness, which has, indeed, been confirmed by a variety of physical performance tests. Methodology/principal findings: A large variety of experimental methodologies, including morphometric measurements of histological preparations, high throughput proteomic analysis, positron emission tomography (PET) and numerous biochemical assays, were used in an attempt to establish the mechanism underlying the relative weakness of rpS6(P-/-) muscles. Collectively, these experiments have demonstrated that the physical inferiority appears to result from two defects: a) a decrease in total muscle mass that reflects impaired growth, rather than aberrant differentiation of myofibers, as well as a diminished abundance of contractile proteins; and b) a reduced content of ATP and phosphocreatine, two readily available energy sources. The abundance of three mitochondrial proteins has been shown to diminish in the knockin mouse. However, the apparent energy deficiency in this genotype does not result from a lower mitochondrial mass or compromised activity of enzymes of the oxidative phosphorylation, nor does it reflect a decline in insulin-dependent glucose uptake, or diminution in storage of glycogen or triacylglycerol (TG) in the muscle. Conclusions/significance: This study establishes rpS6 phosphorylation as a determinant of muscle strength through its role in regulation of myofiber growth and energy content. Interestingly, a similar role has been assigned for ribosomal protein S6 kinase 1, even though it regulates myoblast growth in an rpS6 phosphorylation-independent fashion.


Cyclosporin A-Dependent Downregulation of the Na+/Ca2+ Exchanger Expression

April 2007

·

16 Reads

·

9 Citations

Annals of the New York Academy of Sciences

Cyclosporin A (CsA) is an immunosuppressive drug commonly given to transplant patients. Its application is accompanied by severe side effects related to calcium, among them hypertension and nephrotoxicity. The Na+/Ca2+ exchanger (NCX) is a major calcium regulator expressed in the surface membrane of all excitable and many nonexcitable tissues. Three genes, NCX1, NCX2, and NCX3 code for Na+/Ca2+ exchange activity. NCX1 gene products are the most abundant. We have shown previously that exposure of NCX1-transfected HEK 293 cells to CsA, leads to concentration-dependent reduction of Na+/Ca2+ exchange activity and surface expression, without a reduction in total cell-expressed NCX1 protein. We show now that the effect of CsA on NCX1 protein expression is not restricted to transfected cells overexpressing the NCX1 protein but exhibited also in cells expressing endogenously the NCX1 protein (L6, H9c2, and primary smooth muscle cells). Exposure of NCX2- and NCX3-transfected cells to CsA results also in reduction of Na+/Ca2+ exchange activity and surface expression, though the sensitivity to the drug was lower than in NCX1-transfected cells. Studying the molecular mechanism of CsA-NCX interaction suggests that cyclophilin (Cyp) is involved in NCX1 protein expression and its modulation by CsA. Deletion of 426 amino acids from the large cytoplasmic loop of the protein retains the CsA-dependent downregulation of the truncated NCX1 suggesting that CsA-Cyp-NCX interaction involves the remaining protein domains.


The Structural Basis of Na+‐Ca2+ Exchange Activitya

December 2006

·

15 Reads

·

1 Citation


Post transplant persistence of host cells augments the intensity of acute graft-versus-host disease and level of donor chimerism, an explanation for graft-versus-host disease and rapid displacement of host cells seen following non-myeloablative stem cell transplantation?

October 2006

·

156 Reads

·

8 Citations

Bone Marrow Transplantation

Although the use of non-myeloablative stem cell transplantation (NST) reduces the severity of graft-versus-host disease (GVHD), GVHD remains a major complication following allogeneic transplantation. Since following NST in comparison with myeloablative conditioning, higher proportions of host immunohematopoietic cells may persist while donor-derived alloreactive lymphocytes are being infused, thus possibly serving as host antigen presentation for continuous stimulation of donor T cells, we speculated that GVHD may be similarly amplified by conditioning followed by intentional administration of host cells. This hypothesis was tested in a preclinical animal model. Increased incidence of GVHD, higher mortality and increased levels of chimerism were observed in recipients reconstituted with host cells, particularly with non-irradiated spleen cells. Graft-versus-leukemia effect was not impaired by post transplant cell administration. These results suggest that GVHD may be amplified by recipient cell infusion using either irradiated or viable stimulatory host cells, thus possibly explaining in part higher than anticipated incidence of GVHD and rapid displacement of host cells and conversion to 100% donor type cells following NST. Administration of irradiated host antigen-presenting cells post transplantation may thus represent a potential approach for amplification of the alloreactive capacity of donor lymphocytes following stem cell transplantation.


Citations (21)


... mTOR is a serine/threonine kinase involved in cell growth regulation. It is a component of two distinct multiprotein complexes: mTORC1, which is rapamycin-sensitive, and mTORC2, which was identified as rapamycin-insensitive [20][21][22][23][24]. mTORC1 is activated by nutrients and growth factors and promotes protein synthesis through phosphorylation of the kinase S6rp and the elongation factor 4-EBP1 [25,26]. Akt activates mTORC1 indirectly by phosphorylating and thereby inhibiting the tuberous sclerosis 1/2 (TSC1/2) complex, which leads to stimulation of the small GTPase Rheb and activation of mTORC1 [27][28][29][30]. ...

Reference:

mTORC2 is an important target for simvastatin-associated toxicity in C2C12 cells and mouse skeletal muscle – Roles of Rap1 geranylgeranylation and mitochondrial dysfunction
The TSC-mTOR Pathway Mediates Translational Activation of TOP mRNAs by Insulin Largely in a Raptor- or Rictor-Independent Manner
  • Citing Article
  • March 2009

Molecular and Cellular Biology

... Many factors have been proposed as trans-acting factors that could regulate (positively or negatively) the translation of 5 TOP mRNA, including the T-cell intracellular antigen-1 (TIA1) and TIA-1-related (TIAR) proteins, miR-10A, 4EBP1, and LARP1 [46]. mTORC1 enhance 5 TOP mRNA translation by two mechanisms: phosphorylation of 4EBP1 and/or LARP1 [35,37,39]. ...

Reassessment of the role of TSC, mTORC1 and microRNAs in amino acids-meditated translational control of TOP mRNAs
PLOS ONE

PLOS ONE

... While it has been well established that the 5′TOP motif is both essential and sufficient for rapid mTORC1-mediated translational regulation (9,10), the underlying molecular mechanism has been challenging to resolve (11). Both 4EBP1/2 and LARP1 have been found to contribute to 5′TOP translational inhibition, as loss of either factor partially relieved 5′TOP translational repression in cells acutely treated with the mTOR inhibitor Torin1 (4,5,(12)(13)(14). Although binding of 4EBP1/2 to eIF4E reduces cap-dependent translation of all transcripts, eIF4E may have lower affinity for 5′TOP mRNAs, which could make them more sensitive to mTORC1 inhibition (15,16). ...

Oxygen sufficiency controls TOP mRNA translation via the TSC-Rheb-mTOR pathway in a 4E-BP-independent manner

Journal of Molecular Cell Biology

... Human AdoHcy hydrolase is a cytosolic tetramer of identical subunits requiring NAD + as cofactor (Hershfield et al 1985). Amino acid sequences have been highly conserved over a billion years of evolution (Karis and Hershfield 1989; Hu et al 1999; Kasir et al 1988). In the human gene a 256G>A transition of unknown prevalence has been identified (Karis and Hershfield 1989) and three sequence variants have been reported recently (Gellekink et al 2004). ...

Amino acid sequence of S-adenosyl-L-homocysteine hydrolase from Dictyostelium discoidem as deduced from the cDNA sequence
  • Citing Article
  • May 1988

Biochemical and Biophysical Research Communications

... Under physiological conditions, reaction equilibrium moves toward SAH hydrolysis due to rapid elimination of Hcy, but in in vitro conditions, equilibrium moves toward SAH synthesis (7). SAH is a biologically active substance that is product and inhibitor of the methylation reactions in which S-adenosylmethionine (SAM) acts as methyl group donor (8). SAM has been demonstrated to be an effective sedative and a good sleep modulator (9)(10)(11) and has also exhibited anticonvulsive properties (12). ...

The Gene and Pseudogenes of Rat S ‐Adenosyl‐l‐Homocysteine Hydrolase
  • Citing Article
  • April 1995

European Journal of Biochemistry

... Phosphorylation of eS6 is observed in actively translating ribosomes (Duncan & McConkey, 1982). Knockin mice expressing only the alanine-substituted, nonphosphorylatable version of eS6 exhibit severe whole-body phenotypes, including a reduced size, glucose intolerance, and muscle weakness, along with a higher rate of protein synthesis (Ruvinsky et al., 2005(Ruvinsky et al., , 2009). ...

Mice Deficient in Ribosomal Protein S6 Phosphorylation Suffer from Muscle Weakness that Reflects a Growth Defect and Energy Deficit
PLOS ONE

PLOS ONE

... For instance, the gene expression of RPs can be dynamically regulated by instantaneous translation in response to urgent cellular conditions [7]. In contrast, RP mRNAs undergo excessive production through transcriptional regulation, are stored as inactive messenger ribonucleoprotein particles, and are poised for immediate translation [8]. Dysregulation in translation can lead to abnormal proliferation, cell survival, and immune response, consequently contributing to the development of cancers. ...

The TSC-mTOR Pathway Mediates Translational Activation of TOP mRNAs by Insulin Largely in a Raptor- or Rictor-Independent Manner
Molecular and Cellular Biology

Molecular and Cellular Biology

... However, the frequency of 5mC in mtDNA was significantly higher than that of the corresponding nDNA [66,67]. Nevertheless, CpG methylation in mtDNA was low (3-5%) compared to nDNA [68]. Analysis of methylated cytosines implies that most of them were located outside CpG nucleotides in the L strand (D-loop) [69] and showed high dynamics, for example, in fibroblasts. ...

Methylation pattern of mouse mitochondrial DNA

Nucleic Acids Research

... These regions are essential for ion transport (183,184,195). (B) Subsequent studies (50,60,106,239) revealed that the first helix was a signal peptide not necessary for activity, restricting the mature protein to 11 TMSs with an extracellular N terminus. (C) Cysteine scanning mutagenesis and epitope tagging were then applied to investigate NCX1.1 architecture. ...

The Putative Amino-terminal Signal Peptide of the Cloned Rat Brain Na+-Ca2+ Exchanger Gene (RBE-1) Is Not Mandatory for Functional Expression

Journal of Biological Chemistry

... We show here that KB-R7943 is also a potent MCU inhibitor, an effect which could contribute to its cardioprotective activity. In addition, given that HeLa cells lack any detectable plasma membrane NCX activity (Furman et al., 1993; Low et al., 1993), KB-R7943 could be considered a specific inhibitor of MCU in these cells. We use here this new property of KB-R7943 to show that MCU block inhibits InsP 3 -mediated Ca 2 þ release and [Ca 2 þ ] oscillations in intact HeLa cells. ...

Cloning of the rat heart Na+ -Ca2+ exchanger and its functional expression in HeLa cells
  • Citing Article
  • February 1993