Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 42
Filter
1.
J Physiol Pharmacol ; 56 Suppl 2: 77-88, 2005 Mar.
Article in English | MEDLINE | ID: mdl-16077192

ABSTRACT

Ageing of organisms is among the most complex processes currently known. Understanding the molecular mechanism of physiological ageing is one of the most essential issues in biology and medicine because it is not possible to predict when and how a certain individual will start ageing. In the past centuries human life expectancies increased. Extension of life span is associated with increased susceptibility to a number of chronic diseases. Insight into the cellular and molecular targets of the ageing process would offer the opportunity to prevent at least some of the destructive processes. In the present paper the involvement of two tumor suppressor proteins: wild-type p53 and poly(ADP-ribose)polymerase-1 (PARP-1) in the regulation of cellular senescence and physiological ageing was reviewed. Moreover, the interaction and cross-talk between p53 and PARP1-1 was discussed.


Subject(s)
Aging/metabolism , Cellular Senescence , Poly(ADP-ribose) Polymerases/metabolism , Signal Transduction , Tumor Suppressor Protein p53/metabolism , Aging/genetics , Aging/pathology , Animals , Apoptosis , Basal Metabolism , Cell Cycle , Cellular Senescence/genetics , Humans , Longevity , Mutation , Signal Transduction/genetics , Tumor Suppressor Protein p53/genetics
2.
J Physiol Pharmacol ; 56(1): 15-28, 2005 Mar.
Article in English | MEDLINE | ID: mdl-15795472

ABSTRACT

In this study we investigated the function of p53 as a regulator of cell cycle progression in cycling and senescent cells. Using the conditional temperature-sensitive (ts) mutant we could prevent the detrimental effect of constitutive expression of high levels of wt p53 protein. High levels of wt p53 inhibited cell proliferation by blocking the cells to progress from G1 to S phase of the cell cycle. Flow cytometric analysis revelaed a maintenance of G1 cell population for a longer time depending on the prolonged expression of wt p53 protein. The p53 mediated inhibition of cell proliferation and of the cycle was reversible. However, a spontaneous increase of wt p53 occurring in ageing normal human MRC-5 fibroblasts was associated with irreversible reduction of proliferative potential. The accumulation of G1 cells was detected by flow cytometry. By the measurement of DNA content it is not possible to discriminate between cells arrested in G1 and G0 phase, therefore, the expression of G1 markers was determined. Analysis of the expression of distinct cell cycle regulators revealed that quiescent MRC-5 cells were in G0 phase. Our results indicate that cell cycle arrest occurring in senescent cells is associated with the G0 transition.


Subject(s)
Cell Cycle/physiology , Cellular Senescence/physiology , Tumor Suppressor Protein p53/physiology , Animals , Apoptosis/physiology , Cell Death/physiology , Cell Line, Transformed , Cell Line, Tumor , Humans , Mice , Rats , Resting Phase, Cell Cycle/physiology , Tumor Suppressor Protein p53/biosynthesis
3.
Pol J Pharmacol ; 56(5): 635-41, 2004.
Article in English | MEDLINE | ID: mdl-15591654

ABSTRACT

Cyclin-dependent kinases (CDKs) are serine/threonine kinases that play a key role in the regulation of the cell cycle progression. In proliferating cells, distinct CDKs activated upon complexing with specific cyclins and upon site-specific phosphorylation coordinate in an orchestrated way the appropriate transition between consecutive phases of the cell cycle. Aberrant expression or altered activity of distinct CDK complexes results in escape of cells from the cell cycle control and leads to malignant transformation. Therefore, the inhibition of CDKs in malignant cells provides a new strategy in the fight against cancer. Recently, selective CDK inhibitors targeting distinct CDKs were developed. They represent promising anti-cancer drugs due to their strong anti-proliferative efficacy combined with a relative low direct cytotoxicity. The aim of this study was to compare the effect of two related CDK inhibitors: roscovitine (ROSC) and olomoucine (OLO) on the cell cycle progression in human breast cancer MCF-7 cells. Both examined CDK inhibitors differentially affected the cell cycle progression in MCF-7 cels. Whereas ROSC arrested cells in G(2)/M, OLO inhibited cells at S to G(2) transition and increased the number of cells residing in the S-phase. Moreover, both CDK inhibitors modulated the cell cycle progression with distinct kinetics. Accumulation of G(2)/M-arrested cells beginning 6 h after exposure of cells to ROSC coincided with a strong up-regulation of the p53. Interestingly, ROSC triggered apoptosis in MCF-7 cells by activation of mitochondrial pathway. Loss of the integrity of mitochondrial membrane observed after exposure of cells to ROSC for 6 h led to release of distinct mitochondrial proteins, e.g. apoptosis inducing factor (AIF). In contrast to ROSC, OLO-induced cell cycle changes could be detected after 12 h of the treatment. OLO did not up-regulate p53 protein. It indicates that both examined CDK inhibitors are selective and block the cell cycle progression of human breast carcinoma cells at different phases.


Subject(s)
Breast Neoplasms/enzymology , Cyclin-Dependent Kinases/antagonists & inhibitors , Protein Kinase Inhibitors/pharmacology , Purines/pharmacology , Breast Neoplasms/drug therapy , Cell Cycle/drug effects , Cell Cycle/physiology , Cell Line, Tumor , Cyclin-Dependent Kinases/metabolism , Humans , Kinetin , Protein Kinase Inhibitors/therapeutic use , Purines/therapeutic use , Roscovitine
4.
Drugs Exp Clin Res ; 29(2): 53-67, 2003.
Article in English | MEDLINE | ID: mdl-12951835

ABSTRACT

CHS 828, a novel cyanoguanidine, represents a new class of drugs for cancer therapy, with an unknown primary mechanism of action. It is generally known that anticancer drugs induce p53 response thereby triggering cell cycle arrest or apoptosis. We investigated the effect of CHS 828 on p53 response in normal and tumor cells and compared this effect with that exerted by conventional anticancer drugs. After 24 h of treatment with CHS 828, we observed a dose-dependent up-regulation of wild type (WT) p53 protein in human breast carcinoma MCF-7 cells as well as in normal human and mouse fibroblasts. The highest p53 increase was observed at 300 nM to 1 microM CHS 828. CHS 828 induced phosphorylation of p53 protein at Ser-15 in normal cells. However, the drug failed to induce p53 protein in mouse cells in which the poly(ADP-ribose)-1 gene (PARP-1) was disrupted even at a 30-fold higher dose and after prolonged treatment. Combined treatment of PARP-1 -/- cells by multidrug resistance modulators did not alter p53 expression. CHS 828 inhibited cell proliferation and DNA replication in the tested cells. Interestingly, DNA synthesis as well as proliferation of PARP-1 deficient cells was inhibited by drug concentrations that were approximately 3-fold lower than their conventional counterparts. Treatment of cells with CHS 828 for 48 h did not induce apoptosis.


Subject(s)
Antineoplastic Agents/pharmacology , Cyanides/pharmacology , Guanidines/pharmacology , Tumor Suppressor Protein p53/biosynthesis , Animals , Apoptosis/drug effects , Cell Division/drug effects , Cell Line, Tumor , Cells, Cultured , DNA/biosynthesis , Dose-Response Relationship, Drug , Drug Resistance, Multiple , Drug Resistance, Neoplasm , Fibroblasts/cytology , Fibroblasts/drug effects , Humans , Immunoblotting , Mice , Phosphorylation , Poly Adenosine Diphosphate Ribose/deficiency , Poly Adenosine Diphosphate Ribose/genetics , Tumor Suppressor Protein p53/genetics , Tumor Suppressor Protein p53/metabolism , Up-Regulation
5.
Ann N Y Acad Sci ; 973: 258-64, 2002 Nov.
Article in English | MEDLINE | ID: mdl-12485873

ABSTRACT

The nucleolus represents a highly dynamic nuclear compartment of the interphase nucleus. It plays a key role in ribosome biogenesis. The number of nucleoli, their size, and their activity increase in exponentially growing cells; therefore these parameters reflect the proliferating activity of the cells. A variety of staining techniques have been employed to vizualize nucleolar changes in malignant cells. Staining of so-called nucleolar organizer regions (NORs), based upon a strong avidity of nucleolar proteins to bind silver ions, represents the technique most frequently used by pathologists. Nucleolar changes and pleomorphism associated with overt proliferation of tumor cells have also been documented by immunohistochemical and ultrastructural studies. Contrary to cell proliferation, cytostatics-induced changes of nucleolar phenotype in malignant cells point to a potential role of nucleolar components in the execution of active cell death. Recent studies have provided direct clues that so-called death domains and other apoptosis-related proteins are accumulated in nucleoli upon induction of active cell death. It can be concluded that the plurifunctionality of nucleoli regarding cell proliferation and apoptosis could open new vistas toward understanding dysregulation in malignant cells.


Subject(s)
Apoptosis/physiology , Cell Nucleolus/ultrastructure , Cell Division , Cell Nucleolus/pathology , Cell Nucleolus/physiology , Humans , Neoplasms/pathology , Signal Transduction/physiology
6.
Cell Mol Biol Lett ; 6(2): 117-40, 2001.
Article in English | MEDLINE | ID: mdl-11544635

ABSTRACT

We investigated the interaction between poly(ADP-ribose) polymerase-1 (PARP-1) and the product of the tumor suppressor gene p53 using two different approaches. In the first approach, we used primary and immortalized cells derived from wt and PARP-1 -/- mice. We examined whether PARP-1 deficiency would affect the expression of the wild-type (wt) p53 protein. The inactivation of the PARP-1 gene markedly affected the constitutive expression of the wt p53 protein. Interestingly, only the regularly spliced form of wt p53 was reduced to a barely detectable level in consequence to an approximately 8-fold shortening of its half-life, whereas the level of alternatively spliced p53 remained unchanged. Moreover, reconstitution of cells lacking the PARP-1 gene with the human counterpart restored the normal stability of the regularly spliced p53 protein. In the second approach, we performed experiments with c-Ha-ras transformed primary rat cells overexpressing the p53135val mutant alone or in combination with PARP-1. The advantage of this temperature sensitive p53135val mutant is its oncogenic character at 37 degrees C, connected with cytoplasmic localization of p53, and its tumor suppressor activity at 32 degrees C, accompanied by p53 translocation into the nucleus. No noticeable differences in proliferation and G1 accumulationwere observed between cells expressing p53135val with or without PARP-1. On the other hand, a comparison of the recovery of G1 arrested cells after a shift up to 37 degrees C for both cell lines showed dramatic differences in the kinetics. While cells expressing p53135val rapidly reached the characteristic S-phase level after a shift up to basal temperature, cells additionally expressing PARP-1 rested in G1 despite the temperature elevation. This coincided with exclusively cytoplasmic p53 protein in cells expressing p53135val and predominantly nuclear localization of p53 in p53135val +PARP-1 cells, as evidenced by immunostaining. Determination of the p53 level during the maintenance of cells at 32 degrees C revealed a marked decrease in the level of p53 in cells expressing p53135val alone, whereas in cells coexpressing PARP-1, the level of p53 remained largely unaffected. This indicates that the stability of wild-type p53 greatly differed between both cell lines. Furthermore, the inhibition of PARP-1 activity in G1 arrested cells by 3-aminobenzamide abolished its stabilizing effect on the wild-type p53 protein. Taken together, our results indicate that PARP-1 regulates the stability of the wt p53 protein and that its enzymatic activity is necessary for this stabilizing action.


Subject(s)
Poly(ADP-ribose) Polymerases/metabolism , Tumor Suppressor Protein p53/metabolism , Alternative Splicing/genetics , Animals , Blotting, Western , Cell Cycle/genetics , Cell Division/genetics , Cell Nucleus/metabolism , Cells, Cultured , Doxorubicin/pharmacology , Fibroblasts , Half-Life , Humans , Mice , Mice, Knockout , Microscopy, Fluorescence , Mutation/genetics , Phenotype , Poly(ADP-ribose) Polymerase Inhibitors , Poly(ADP-ribose) Polymerases/deficiency , Poly(ADP-ribose) Polymerases/genetics , RNA, Messenger/genetics , RNA, Messenger/metabolism , Rats , Transcription, Genetic/drug effects , Transcriptional Activation/drug effects , Tumor Suppressor Protein p53/genetics
7.
J Cell Biochem Suppl ; Suppl 36: 1-11, 2001.
Article in English | MEDLINE | ID: mdl-11455565

ABSTRACT

The Harvey-ras gene encodes small guanine nucleotide binding proteins, mutant forms of which are associated with a number of human malignancies. Based on studies with truncated forms of the protein it is known that correct post-translational processing of Ras is essential for cytoplasmic membrane localization and function. Surprisingly, immunofluorescence analysis provided evidence that in addition to its cytosolic localization, activated H-Ras(Val 12) was also localized in the nuclei of transformed cells both in vitro and in vivo. Immunoblot analysis of nuclear fractions was consistent with results found by immunohistochemistry. Moreover, inhibition of protein farnesylation prevented the nuclear targeting of activated H-Ras(Val 12) and NFkappaB. Alterations in subcellular distribution pattern and phosphorylation of the cell cycle inhibitor p27, which is involved in Ras driven tumor growth, coincided with nuclear localization of H-Ras(Val 12). Proteins are often not functional until they are transported to their final destination. Indeed, Ras was found to complex with NTF2 a factor involved in nuclear protein import and export. Therefore it is suggested that NTF2 is the actual carrier for oncogenic Ras. In view of these observations the question arises whether the nuclear localization of H-Ras(Val 12) in tumors is important in oncogenic activation or whether it is a response to apoptosis. J. Cell. Biochem. Suppl. 36: 1-11, 2001.


Subject(s)
Nuclear Proteins/metabolism , Nucleocytoplasmic Transport Proteins , Oncogene Proteins/metabolism , Subcellular Fractions/metabolism , Tumor Suppressor Proteins , ras Proteins/metabolism , Alkyl and Aryl Transferases/antagonists & inhibitors , Animals , Blotting, Western , Carrier Proteins/metabolism , Cell Cycle Proteins/metabolism , Cell Line , Cyclin-Dependent Kinase Inhibitor p27 , Diethylnitrosamine , Enzyme Inhibitors/pharmacology , Fluorescent Antibody Technique , Liver/metabolism , Liver/ultrastructure , Liver Neoplasms, Experimental/chemically induced , Liver Neoplasms, Experimental/metabolism , Mice , Mice, Inbred C3H , Microscopy, Electron , NF-kappa B/metabolism , Precipitin Tests , Protein Processing, Post-Translational , Rats
8.
J Hepatol ; 34(3): 366-72, 2001 Mar.
Article in English | MEDLINE | ID: mdl-11322196

ABSTRACT

BACKGROUND/AIMS: Antibodies against nuclear pore complexes (NPCs) have been detected in primary biliary cirrhosis (PBC), but their clinical relevance is still unsettled. METHODS: We tested sera from 171 consecutive PBC patients and 230 control subjects (149 with autoimmune or viral liver diseases, 28 with systemic lupus erythematosus, and 53 healthy) by immunoblotting for antibodies against purified human NPCs. RESULTS: Antibodies to NPCs were detected in 27% of the patients with PBC, were highly specific (97%), and were not associated with antimitochondrial antibodies. Their prevalence was higher in symptomatic patients (36 vs. 16%, P < 0.01) and was associated (P < 0.001) with more severe disease, as assessed by the presence of cirrhosis or its complications (13% prevalence in patients without cirrhosis, 31% in uncomplicated, and 54% in complicated cirrhosis), or by the application of the Mayo prognostic model (12% in the lowest, 21% in the median, 47% in the highest score tertile). Positive patients had higher levels of serum bilirubin (2.2 +/- 3.7 vs. 1.0 +/- 1.1 mg/dl, P < 0.01) and more marked inflammatory infiltrates on liver biopsy (P < 0.05). CONCLUSIONS: Autoantibodies to NPCs are more prevalent in PBC patients than in controls and are strongly associated with more active and severe disease.


Subject(s)
Autoantibodies/analysis , Liver Cirrhosis, Biliary/immunology , Liver Cirrhosis, Biliary/pathology , Nuclear Pore/immunology , Bilirubin/blood , Humans , Liver/pathology , Liver Cirrhosis, Biliary/blood , Liver Cirrhosis, Biliary/complications , Liver Diseases/immunology , Lupus Erythematosus, Systemic/immunology , Middle Aged , Models, Theoretical , Prognosis , Reference Values , Severity of Illness Index
9.
J Cell Sci ; 114(Pt 4): 663-70, 2001 Feb.
Article in English | MEDLINE | ID: mdl-11171371

ABSTRACT

We studied morphological changes of the nucleoli in HeLa cells treated with cisplatin and compared them with induction of markers of programmed cell death and TUNEL staining. We used different light microscopic nucleolar staining methods allowing us to visualize not only nucleolar proteins but also nucleolar RNA. Our results show predominantly compact, centrally localized nucleoli in intact control HeLa cells. In cisplatin-treated HeLa cells, we found an early onset of nucleolar segregation of proteins detected by argyrophilic nucleolar organizer regions and anti-nucleolar monoclonal antibody as well as an increased immunoreactivity for activated caspase-3 after 6 hours. Staining with Toluidine Blue and Methyl-green Pyronine revealed segregated nucleoli 12 hours after the treatment with cisplatin. TUNEL positivity in cisplatin-treated HeLa cells was accompanied by the aggregation of the argyrophilic proteins in the central portion of nucleus, disappearance of nucleolar RNA and shrinkage of the nucleus after 24 hours. Monitoring of the biochemical changes by immunoblotting revealed that activation of distinct caspases and degradation of their downstream protein substrates is executed in two phases. During an early apoptotic stage beginning 4.5 hours post treatment an activation of caspase-9 and caspase-3 was observed. This was accompanied by proteolytic cleavage of poly(ADP-ribose) polymerase-1 (PARP-1). The caspase-9 activation seems to be mediated by recruitment by the activating factor Apaf-1 because the increased accumulation of Apaf-1 and cytochrome C in cytosol preceded the generation of mature caspase-9 form. A second phase of apoptosis occurring between 10 and 15 hours post treatment was characterized by degradation of other nucleolar and nuclear proteins such as nuclear lamins, topoisomerase I and B23. In conclusion, remarkable segregation of nucleolar argyrophilic proteins, nucleolar RNA and a simultaneous activation of the cascade of caspases markedly preceded the TUNEL positivity in cisplatin-treated HeLa cells thereby substantiating the hypothesis that the nucleolus is a preferred target for caspase-3-dependent proteolysis in cisplatin-treated HeLa cells.


Subject(s)
Caspases/metabolism , Cell Nucleolus/ultrastructure , Cisplatin/pharmacology , Apoptosis/drug effects , Caspase 3 , Caspase 9 , Cell Fractionation , Cell Size/drug effects , Enzyme Activation , HeLa Cells , Humans , Hydrolysis , Immunohistochemistry , In Situ Nick-End Labeling , Poly(ADP-ribose) Polymerases/metabolism
10.
J Cell Biochem ; 80(1): 85-103, 2000 Sep 18.
Article in English | MEDLINE | ID: mdl-11029756

ABSTRACT

We have previously reported that in cells ectopically expressing temperature-sensitive p53(135val) mutant, p53 formed tight complexes with poly(ADP-ribose) polymerase (PARP). At elevated temperatures, p53(135val) protein, adopting the mutant phenotype, was localized in the cytoplasm and sequestered the endogenous PARP. To prove whether an excess of p53(135val) protein led to this unusual intracellular distribution of PARP, we have established cell lines overexpressing p53(135val) + c-Ha-ras alone or in combination with PARP. Interestingly, immunostaining revealed that PARP is sequestered in the cytoplasm by mutant p53 in cells overexpressing both proteins. Simultaneous overexpression of PARP had no effect on temperature-dependent cell proliferation and only negligibly affected the kinetics of p53-mediated G(1) arrest. However, if the cells were completely growth arrested at 32 degrees C and then shifted up to 37 degrees C, coexpressed PARP dramatically delayed the reentry of transformed cells into the cell cycle. Even after 72 h at 37 degrees C the proportion of S-phase cells was reduced to 20% compared to those expressing only p53(135val) + c-Ha-ras. The coexpressed PARP stabilized wt p53 protein and its enzymatic activity was necessary for stabilization.


Subject(s)
G1 Phase/genetics , Poly(ADP-ribose) Polymerases/metabolism , Tumor Suppressor Protein p53/physiology , Animals , Cell Division/genetics , Cell Line, Transformed , Fluorescent Antibody Technique, Indirect , Phenotype , Precipitin Tests , Rats , Tumor Suppressor Protein p53/genetics
11.
Oncogene ; 19(36): 4096-107, 2000 Aug 24.
Article in English | MEDLINE | ID: mdl-10962570

ABSTRACT

To dissect the p53-dependent apoptotic pathway, events following induction of temperature sensitive (ts) p53val138 were studied in a Ewing tumor cell line. Transcriptional deregulation of p53 targets first observable after 1 h at 32 degrees C preceded activation of caspases and the break-down of mitochondrial respiratory activity. Activation of caspases was first observed 4 h after p53 induction. Using peptide inhibitors we identified activation of caspase 8 upstream of caspases-9 and -3. Although the caspase 8 specific inhibitor z-IETD.fmk did not affect translocation of BAX to the mitochondrial membrane and cytochrome C release it almost completely blocked cleavage of the prototype caspase substrate PARP and DNA fragmentation while enforcing mitochondrial depolarization and production of reactive oxygene species (ROS). Activation of caspase 8 did not involve death-domain receptor signaling. Expression of BCL2 only partially suppressed caspase activation but blocked apoptosis. Replacement of the N-terminus of p53val138 by the related VP16 transactivation domain created a ts p53 with a tanscriptional activity indistinguishable from p53val138 until the time of caspase activation. However, the VP16 - p53 fusion failed to trigger caspases and subsequent induction of the ROS producing gene pig3 paralleled by complete loss of apoptotic activity. These results indicate that p53-dependent transcriptional deregulation, triggering of the caspase cascade and the mitochondrial break-down occur in a timely ordered sequence coordinated by the genuine p53 amino terminus and suggest caspase 8 and PIG3 as key regulatory elements in this process. Oncogene (2000) 19, 4096 - 4107


Subject(s)
Apoptosis , Caspases/genetics , Tumor Suppressor Protein p53/metabolism , Blotting, Northern , Caspase 8 , Caspase 9 , Caspase Inhibitors , Caspases/metabolism , Cysteine Proteinase Inhibitors/pharmacology , Cytochrome c Group/metabolism , DNA Fragmentation , Flow Cytometry , Herpes Simplex Virus Protein Vmw65/genetics , Humans , In Situ Nick-End Labeling , Intracellular Signaling Peptides and Proteins , Mitochondria/metabolism , Mutation , Oligopeptides/pharmacology , Poly(ADP-ribose) Polymerases/genetics , Poly(ADP-ribose) Polymerases/metabolism , Proteins/genetics , Proteins/metabolism , Proto-Oncogene Proteins/genetics , Proto-Oncogene Proteins/metabolism , Proto-Oncogene Proteins c-bcl-2/genetics , Proto-Oncogene Proteins c-bcl-2/metabolism , Reactive Oxygen Species/metabolism , Recombinant Fusion Proteins/genetics , Recombinant Fusion Proteins/metabolism , Sarcoma, Ewing , Transcriptional Activation , Tumor Cells, Cultured , Tumor Suppressor Protein p53/genetics , bcl-2-Associated X Protein
12.
Cancer Res ; 60(15): 4238-44, 2000 Aug 01.
Article in English | MEDLINE | ID: mdl-10945636

ABSTRACT

Mouse embryo fibroblasts lacking poly(ADP-ribose) polymerase (PARP)-1 express a barely detectable level of wild-type (wt) p53 protein. Doxorubicin at concentrations activating wt p53 in normal mouse embryo fibroblasts failed to induce it in mutant cells. wt p53 was only activated in response to a 10-fold higher doxorubicin dose. Treatment with higher doxorubicin concentrations was cytotoxic for normal but not for PARP-1 -/- cells. The latter was also resistant to other anticancer agents. The increased resistance of mutant cells to drugs resembled a unique phenomenon known as multidrug resistance (MDR). Interestingly, the MDR gene product P-glycoprotein was clearly up-regulated in PARP-1-deficient cells as compared with normal counterparts. Pretreatment with verapamil reversed the MDR phenotype.


Subject(s)
ATP Binding Cassette Transporter, Subfamily B, Member 1/metabolism , Drug Resistance, Multiple/physiology , Poly(ADP-ribose) Polymerases/deficiency , ATP Binding Cassette Transporter, Subfamily B, Member 1/antagonists & inhibitors , ATP Binding Cassette Transporter, Subfamily B, Member 1/genetics , Animals , Antineoplastic Agents/pharmacokinetics , Antineoplastic Agents/pharmacology , Breast Neoplasms/drug therapy , Breast Neoplasms/enzymology , Breast Neoplasms/metabolism , Clone Cells , Dose-Response Relationship, Drug , Doxorubicin/pharmacokinetics , Doxorubicin/pharmacology , Drug Resistance, Multiple/genetics , Drug Resistance, Neoplasm/genetics , Drug Resistance, Neoplasm/physiology , Embryo, Mammalian , Fibroblasts/drug effects , Fibroblasts/enzymology , Fibroblasts/metabolism , Gene Expression Regulation/drug effects , Gene Expression Regulation, Neoplastic , Humans , Mice , Mice, Knockout , Poly(ADP-ribose) Polymerases/genetics , Poly(ADP-ribose) Polymerases/metabolism , Probenecid/pharmacology , Tumor Cells, Cultured , Tumor Suppressor Protein p53/biosynthesis , Tumor Suppressor Protein p53/genetics , Up-Regulation , Verapamil/pharmacology
13.
J Cell Biochem ; 78(4): 681-96, 2000 Jun 12.
Article in English | MEDLINE | ID: mdl-10861865

ABSTRACT

Recently we found a clearly reduced basal level of wt p53 protein in PARP-deficient cells. Interestingly, PARP deficiency affected only regularly spliced (RS) wt p53. No significant difference of the p53 transcription rate was observed between wt and PARP-lacking cells. To clarify whether the reduction of RS p53 protein is due to a lower translation rate or rather to its instability in the absence of functional PARP, we investigated the effect of the inhibition of proteasome activity and nuclear export on the p53 level. The p53 half-life was approximately eight-fold decreased in PARP-lacking cells. Surprisingly, treatment with three proteasome inhibitors increased RS p53 in normal but not in PARP-deficient cells. However, the inhibition of nuclear export resulted in a considerable accumulation of RS p53 in the latter. Therefore, we decided to increase concentrations of the inhibitors. Their higher concentrations strongly affected viability of normal, but not of PARP-deficient cells, about 70% of MEFs died. Interestingly, higher concentrations of proteasome inhibitors resulted in the appearance of RS p53 in PARP-lacking fibroblasts. Reconstitution of PARP-deficient cells with PARP restored the normal susceptibility to proteasome inhibitors thereby unequivocally demonstrating that the enhanced cytotoxicity of proteasome inhibitors and their action on p53 level depends on the presence of functional PARP.


Subject(s)
Cysteine Endopeptidases/metabolism , Multienzyme Complexes/metabolism , Proteins/genetics , Proteins/physiology , Acetylcysteine/analogs & derivatives , Acetylcysteine/pharmacology , Animals , Cell Line , Cell Nucleus/drug effects , Cell Survival/drug effects , Cells, Cultured , Cysteine Proteinase Inhibitors/pharmacology , Dose-Response Relationship, Drug , Fatty Acids, Unsaturated/pharmacology , Fibroblasts/drug effects , Fluorescent Dyes/pharmacology , Humans , Immunoblotting , Indoles/pharmacology , Leupeptins/pharmacology , Mice , Mice, Knockout , Microscopy, Fluorescence , Oligopeptides/pharmacology , Phenotype , Poly (ADP-Ribose) Polymerase-1 , Poly(ADP-ribose) Polymerases , Proteasome Endopeptidase Complex , Time Factors , Tumor Suppressor Protein p53/biosynthesis
14.
Mol Carcinog ; 24(4): 263-75, 1999 Apr.
Article in English | MEDLINE | ID: mdl-10326863

ABSTRACT

The aim of this work was to study the role of the tumor suppressor p53 and of poly(ADP-ribose) transferase (pADPRT) in the control of hepatocyte apoptosis in two different in vivo models, i.e., during the process of tumor initiation by the genotoxin and cytotoxin N-nitrosomorpholine (NNM) and after withdrawal of the hepatomitogen cyproterone acetate (CPA). Treatment with NNM induces apoptosis followed by necrosis and regenerative DNA synthesis. At the first wave of apoptosis 12 h after NNM application, no p53 expression could be detected by immunohistochemical analysis and immunoblotting. However, 24 h after treatment, numerous p53-positive hepatocyte nuclei were detected, whereas hepatocytes in early and later stages of apoptosis were always negative. Simultaneously with the increased p53 levels, p21 protein was induced. This was accompanied by a block in replicative DNA synthesis, as detected by proliferating-cell nuclear antigen immunostaining. Concomitantly with the increase in apoptosis, dramatic degradation of the nuclear enzyme pADPRT was observed, as evidenced by immunoblotting and activity blotting. The decrease in pADPRT enzymatic activity observed 12 h after treatment coincided with the greatest extent of pADPRT cleavage. One prominent cleavage product was 64 kDa, suggesting that granzyme B was involved in pADPRT degradation. In the second in vivo model we used, i.e., withdrawal of treatment with the hepatomitogen CPA, apoptosis of excessive hepatocytes but no necrosis occurs. Again, no induction of p53 expression could be detected in the liver even at the maximum level of apoptosis, whereas a strong correlation between induction of apoptosis and cleavage of pADPRT to a 64-kDa fragment was observed. These results from whole-animal experiments strongly suggest that the induction of apoptosis in rat liver after genotoxic and cytotoxic damage and during regression of hyperplasia is driven by a p53-independent pathway but is accompanied by cleavage of pADPRT.


Subject(s)
Apoptosis/physiology , Carcinogens/toxicity , Cyproterone Acetate/pharmacology , Liver/drug effects , Mitogens/pharmacology , Nitrosamines/toxicity , Poly(ADP-ribose) Polymerases/metabolism , Tumor Suppressor Protein p53/physiology , Animals , Cyclin-Dependent Kinase Inhibitor p21 , Cyclins/biosynthesis , Cyclins/genetics , DNA Replication , Female , Gene Expression Regulation/drug effects , Genes, p53 , Glycoside Hydrolases/metabolism , Granzymes , Hyperplasia , Liver/pathology , Liver Neoplasms, Experimental/chemically induced , Liver Neoplasms, Experimental/pathology , Male , Necrosis , Rats , Rats, Wistar , Recombinant Fusion Proteins/biosynthesis , Serine Endopeptidases/physiology , Specific Pathogen-Free Organisms , Testis/drug effects , Testis/pathology , Tumor Suppressor Protein p53/biosynthesis
15.
Liver ; 19(2): 122-8, 1999 Apr.
Article in English | MEDLINE | ID: mdl-10220742

ABSTRACT

AIMS/BACKGROUND: Autoimmune cholangitis/cholangiopathy (AIC) is an enigmatic disease marked by chronic cholangitis, antinuclear antibodies (ANA) and sero-negativity for conventionally detected antimitochondrial antibodies (AMA). We examined whether AIC is a distinct entity, an AMA-negative variant of primary biliary cirrhosis (PBC), or a cholangiopathic variant of autoimmune hepatitis (AIH) by comparing the clinical, laboratory and autoantibody profiles of 21 cases of AIC, 37 cases PBC and 16 cases of AIH from selected Japanese patients. METHODS: The specificities of AMA and ANA were determined by immunofluorescence, immunoblotting and enzyme inhibition assays using various mitochondrial and nuclear autoantigens, and the frequencies for these groups were compared. RESULTS: By clinical, biochemical and histological data, AIC and PBC were similar and both were clearly distinct from AIH. Serologically, by immunofluorescence of AMA and ANA, there was polarisation. By immunoblotting, and notwithstanding the negative test for AMA, a proportion of the AIC sera reacted with the E2 subunits of the 2-oxo-acid dehydrogenase enzyme complexes, but more particularly with the lower molecular weight E2 subunits. The antinuclear reactivity in AIC was with centromere, Sp100 and nuclear pore complex proteins as in PBC, but preferentially with the nuclear pore complex. CONCLUSION: Our results demonstrate that AIC and PBC are similar diseases. However this duo is of interest because, usually, among sets of autoimmune syndromes, differences in serological targetting are matched by differences in clinical presentation: AIC and PBC are an exception to this.


Subject(s)
Autoimmune Diseases/immunology , Cholangitis/immunology , Hepatitis, Autoimmune/immunology , Liver Cirrhosis, Biliary/immunology , Adult , Aged , Aged, 80 and over , Antibodies, Antinuclear/analysis , Autoantibodies/analysis , Catalysis , Female , Fluorescent Antibody Technique , Humans , Immunoblotting , Male , Middle Aged , Mitochondria/immunology , Muscle, Smooth/immunology , Nuclear Proteins/immunology , Pyruvate Dehydrogenase Complex/antagonists & inhibitors
16.
Biochem Biophys Res Commun ; 255(2): 399-405, 1999 Feb 16.
Article in English | MEDLINE | ID: mdl-10049720

ABSTRACT

We have investigated the effect of PARP gene inactivation on the expression of wild-type (wt) p53 protein. Using immortalized fibroblasts from control and PARP knock-out mice we have found by immunoblotting with the PAb421 antibody a profound decrease of the p53 expression to a barely detectable level in PARP knock-out cells. Surprisingly, longer exposure of immunoblots revealed an immunoreactive band at about 75 kD which was stronger in PARP-deficient cells than in wt cells and was not affected upon doxorubicin treatment. The size of the PAb421 immunoreactive protein and the lack of its inducibility in response to DNA damage resembled those of p73, the first described p53 homologue. Therefore, we examined the reactivity of anti-p53 antibodies with in vitro translated p73 protein. Interestingly, p73 was efficiently immunoprecipitated with distinct antibodies recognizing the carboxy-terminus of p53. In Northern blots we observed p73 signals of comparable intensity in controls and PARP-deficient cells. We conclude that elevated expression of p73 may compensate the reduced level of p53 in PARP-deficient cells.


Subject(s)
DNA-Binding Proteins/biosynthesis , Fibroblasts/metabolism , Nuclear Proteins/biosynthesis , Poly(ADP-ribose) Polymerases/deficiency , Poly(ADP-ribose) Polymerases/genetics , RNA Splicing , Animals , Apoptosis/genetics , Blotting, Northern , Cell Line, Transformed , Culture Media, Serum-Free , DNA-Binding Proteins/genetics , DNA-Binding Proteins/metabolism , Doxorubicin/pharmacology , Embryo, Mammalian , Fibroblasts/drug effects , Fibroblasts/enzymology , Genes, Tumor Suppressor , Immunoblotting , Mice , Mice, Knockout , Nuclear Proteins/genetics , Nuclear Proteins/metabolism , Tumor Protein p73 , Tumor Suppressor Proteins
17.
Cancer Res ; 59(1): 28-34, 1999 Jan 01.
Article in English | MEDLINE | ID: mdl-9892179

ABSTRACT

The interaction between poly(ADP-ribose) polymerase (PARP) and the product of the tumor suppressor gene p53 has been described previously. Here, we have investigated whether PARP deficiency may affect the expression and regulation of wild-type (wt) p53. For this purpose, we have used immortalized cells derived from wt and PARP knockout mice. We have found a clearly reduced basal level of PAb421 immunoreactive wt p53 protein in PARP-deficient cells. The monoclonal antibody PAb421 is known to recognize an epitope in the COOH terminus of normally spliced p53 protein. Under indirect immunofluorescence, this antibody stained nuclei in normal but not in PARP-deficient cells. Despite marked reduction of wt p53 protein in PARP knockout cells, no significant difference of the p53 transcription rate was observed between wt and PARP-deficient cells. Interestingly, in both cell types, an additional p53 transcript representing the alternatively spliced (AS) p53 form was detected. Because of its reactivity with different specific anti-p53 antibodies, we have determined that the p53 protein present in PARP knockout mouse cells possesses characteristic features of AS p53. Our results clearly show that PARP-deficient cells constitutively express the AS form of wt p53 and indicate that the regularly spliced p53 is extremely unstable in the absence of PARP. Moreover, PARP-/- cells fail to transactivate p53-responsive genes. Treatment of PARP-/- cells with genotoxic agents primarily leads to the activation of AS p53 protein.


Subject(s)
Poly(ADP-ribose) Polymerases/genetics , RNA Splicing/genetics , Tumor Suppressor Protein p53/genetics , Animals , Cell Line, Transformed , Fibroblasts , Gene Expression Regulation , Mice , Mice, Knockout , Poly(ADP-ribose) Polymerases/deficiency
18.
Histochem Cell Biol ; 109(2): 111-8, 1998 Feb.
Article in English | MEDLINE | ID: mdl-9504771

ABSTRACT

We investigated how only three morphologically distinguished nucleolar components can integrate the many necessary tasks in ribosome biogenesis. For the mapping of ribosomal (r)DNA transcription loci, we combined non-autoradiographic in situ transcription assays with the immunological analysis of the ultrastructural distribution of transcription-associated proteins, i.e., polymerase I, the human polymerase I-specific upstream binding factor, and topoisomerase I. Furthermore, we visualized the nascent transcripts simultaneously with the rDNA in the nucleoli. All tested transcription proteins were found in both the fibrillar center and the dense fibrillar component (DF) of nucleoli in human cells. In the DF the nascent transcripts, detected by bromouridine incorporation, were found colocalized with the transcription proteins only within circumscribed regions. We did not observe colocalization of rDNA with nascent transcripts within the fibrillar centers, which corroborates the view that transcription proteins in this component are rather inactive. Our results suggests that only a minor portion of the DF is involved in transcriptional activity. Transcription appears to be confined to small foci, which exist close to or associated with the DF. Our results are in favor of the view that the DF has different functions which are localized in subcompartments of the DF.


Subject(s)
Cell Nucleolus/genetics , DNA, Ribosomal/genetics , Pol1 Transcription Initiation Complex Proteins , Bromodeoxyuridine/metabolism , Cell Nucleolus/metabolism , Cell Nucleolus/ultrastructure , DNA Polymerase I/metabolism , DNA Topoisomerases, Type I/metabolism , DNA, Ribosomal/metabolism , DNA-Binding Proteins/metabolism , HeLa Cells , Humans , In Vitro Techniques , Male , Microscopy, Immunoelectron , Sertoli Cells/metabolism , Sertoli Cells/ultrastructure , Transcription Factors/metabolism , Transcription, Genetic
19.
Gastroenterology ; 114(2): 329-35, 1998 Feb.
Article in English | MEDLINE | ID: mdl-9453494

ABSTRACT

BACKGROUND & AIMS: Autoimmmune hepatitis (AIH), a chronic liver disorder, can be classified into two subtypes on the basis of the specificities of circulating autoantibodies. Type I AIH is defined by antibodies to nuclear and/or smooth muscle antigens (SMA), and type II is characterized by antibodies to cytochrome P450IID6. There is an additional type of AIH characterized by antibodies to a cytosolic soluble liver antigen (SLA), which can occur alone or in combination with antinuclear antibodies and SMA. The aim of this study was to identify the reactive antigen in SLA, a heterogenous cytosolic fraction consisting of at least 100 extremely soluble proteins. METHODS: Sera from 31 patients with AIH reacting with SLA and from 30 disease controls were tested. The immunoreactive antigens were determined using immunoprecipitation and immunoblotting after one- and two-dimensional polyacrylamide gel electrophoresis. The antigens were identified by microsequencing of the corresponding protein spots. RESULTS: Twenty-five of 31 anti-SLA-positive sera (80, 7%) reacted with a set of proteins ranging from 25 to 27 kilodaltons that were identified as three subunits of glutathione S-transferases: Ya, Yb1, and Yc. CONCLUSIONS: Glutathione S-transferase subunit proteins represent the major autoantigen in anti-SLA-positive AIH. This new finding permits the establishment of standardized immunoassays for routine diagnosis.


Subject(s)
Autoantigens/analysis , Glutathione Transferase/immunology , Hepatitis, Autoimmune/immunology , Adolescent , Adult , Animals , Child , Cytosol/immunology , Female , Glutathione Transferase/genetics , Humans , Liver/immunology , Male , Middle Aged , Rats
20.
Biochem Biophys Res Commun ; 224(1): 96-102, 1996 Jul 05.
Article in English | MEDLINE | ID: mdl-8694840

ABSTRACT

We have recently reported that mutant but not wild-type (wt) p53 protein was ADP-ribosylated in primary rat cells overexpressing the temperature-sensitive murine p53val135 gene. To examine whether the lack of susceptibility to modification is a specific feature of p53val135 adopting wt conformation or rather a general property of this tumor suppressor protein, we have studied ADP-ribosylation of wt p53 of different origin in vitro using semi-purified poly(ADP-ribose) transferase (pADPRT). In vitro pADPRT modified human and mouse wt p53 and p53val135. Under limiting substrate concentration, the molar mass of ADP-ribosylated p53 was only slightly altered. Chase experiments with high NAD concentration resulted in the formation of poly(ADP-ribosyl)ated p53 protein shifted to 64 kD. However, preincubation of wt p53 proteins with a p53 consensus sequence resulting in complex formation abolished the modification of wt p53. This indicates that in the cellular environment the specific DNA binding of wt p53 prevents its covalent modification by poly(ADP-ribose).


Subject(s)
Poly(ADP-ribose) Polymerases/metabolism , Tumor Suppressor Protein p53/metabolism , Adenosine Diphosphate Ribose/metabolism , Animals , Cells, Cultured , Consensus Sequence , Embryo, Mammalian , Embryo, Nonmammalian , Humans , Mice , NAD/metabolism , Rats , Rats, Inbred F344 , Recombinant Proteins/biosynthesis , Recombinant Proteins/metabolism , Spodoptera , Substrate Specificity , Transfection , Tumor Suppressor Protein p53/biosynthesis
SELECTION OF CITATIONS
SEARCH DETAIL