首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
2.
3.
A polyomavirus mutant (315YF) blocked in binding phosphatidylinositol 3-kinase (PI 3-kinase) has previously been shown to be partially deficient in transformation and to induce fewer tumors and with a significant delay compared to wild-type virus. The role of polyomavirus middle T antigen-activated PI 3-kinase in apoptosis was investigated as a possible cause of this behavior. When grown in medium containing 1d-3-deoxy-3-fluoro-myo-inositol to block formation of 3′-phosphorylated phosphatidylinositols, F111 rat fibroblasts transformed by wild-type polyomavirus (PyF), but not normal F111 cells, showed a marked loss of viability with evidence of apoptosis. Similarly, treatment with wortmannin, an inhibitor of PI 3-kinase, stimulated apoptosis in PyF cells but not in normal cells. Activation of Akt, a serine/threonine kinase whose activity has been correlated with regulation of apoptosis, was roughly twofold higher in F111 cells transformed by either wild-type virus or mutant 250YS blocked in binding Shc compared to cells transformed by mutant 315YF. In the same cells, levels of apoptosis were inversely correlated with Akt activity. Apoptosis induced by serum withdrawal in Rat-1 cells expressing a temperature-sensitive p53 was shown to be at least partially p53 independent. Expression of either wild-type or 250YS middle T antigen inhibited apoptosis in serum-starved Rat-1 cells at both permissive and restrictive temperatures for p53. Mutant 315YF middle T antigen was partially defective for inhibition of apoptosis in these cells. The results indicate that unlike other DNA tumor viruses which block apoptosis by inactivation of p53, polyomavirus achieves protection from apoptotic death through a middle T antigen–PI 3-kinase–Akt pathway that is at least partially p53 independent.Programmed cell death occurs during normal development and under certain pathological conditions. In mammalian cells, apoptosis can be induced by a variety of stimuli, including DNA damage (45), virus infection (54, 57), oncogene activation (25), and serum withdrawal (34, 37). Apoptosis can also be blocked by a number of factors, including adenovirus E1B 55- or 19-kDa proteins (9, 16), baculovirus p35 and iap genes (10), Bcl-2 (36, 61), and survival factors (12, 21). DNA tumor viruses have evolved mechanisms that both trigger and inhibit apoptosis. These frequently involve binding and inactivation of tumor suppressor proteins. E7 in some papillomaviruses (22), E1A in adenovirus (31, 43, 64), and large T antigen in simian virus 40 (SV40) (17) bind Rb and/or p300 and lead to upregulation of p53, which is thought to trigger apoptosis in virus-infected cells. The same viruses also inhibit apoptosis by inactivating p53 by various mechanisms (44, 63, 67). In contrast, the mechanism by which polyomavirus interacts with apoptotic pathways in the cell is not known; no direct interaction with p53 by any of the proteins encoded by this virus has been demonstrated (19, 62).The principal oncoprotein of polyomavirus is the middle T antigen. Neoplastic transformation by polyomavirus middle T antigen has as a central feature its association with and activation of members of the Src family of tyrosine kinases p60c-src (13) and p62c-yes (42). The major known consequence of these interactions is phosphorylation of middle T antigen on specific tyrosine residues creating binding sites for other signaling proteins. Phosphorylation at tyrosines 250, 315, and 322 promotes binding to Shc (18), the p85 regulatory subunit of phosphatidylinositol 3-kinase (PI 3-kinase) (59), and phospholipase Cγ-1 (58), respectively. Recognition of multiple signaling pathways emanating from middle T antigen has led to a keen interest in identifying their downstream biochemical effects, which collectively lead to the emergence of neoplastic transformation and presumably underlie the dramatic ability of the virus to induce many kinds of tumors in the mouse.Previous work has shown that the binding of PI 3-kinase to middle T antigen is essential for full transformation of rat fibroblasts in culture (8) and for rapid development of a broad spectrum of tumors in mice (30), for translocation of the GLUT1 transporter (68), and activation of p70 S6 kinase (14). While the mutant 315YF (blocked in PI 3-kinase activation) was able to induce some tumors, it did so at reduced frequencies and with an average latency three times longer than that of either the wild-type virus or a mutant, 250YS, blocked in binding Shc (4, 30). Recent studies have indicated a role of PI 3-kinase in blocking apoptosis in nonviral systems. Growth factor receptors acting through protein tyrosine kinases may prevent apoptosis by activating PI 3-kinase in PC12 cells, T lymphocytes, hematopoietic progenitors, and rat fibroblasts (7, 48, 56, 65, 66). The failure of mutant 315YF to induce full transformation of cells in culture and to induce the rapid development of tumors in mice could therefore be related, at least in part, to a failure to block apoptosis. In this study, we focus on the question of whether middle T antigen–PI 3-kinase interaction is involved in blocking apoptosis in cells transformed by polyomavirus.  相似文献   

4.
Whereas it is now clear that human bone marrow stromal cells (BMSCs) can be immunosuppressive and escape cytotoxic lymphocytes (CTLs) in vitro and in vivo, the mechanisms of this phenomenon remain controversial. Here, we test the hypothesis that BMSCs suppress immune responses by Fas-mediated apoptosis of activated lymphocytes and find both Fas and FasL expression by primary BMSCs. Jurkat cells or activated lymphocytes were each killed by BMSCs after 72 h of co-incubation. In comparison, the cytotoxic effect of BMSCs on non-activated lymphocytes and on caspase-8(−/−) Jurkat cells was extremely low. Fas/Fc fusion protein strongly inhibited BMSC-induced lymphocyte apoptosis. Although we detected a high level of Fas expression in BMSCs, stimulation of Fas with anti-Fas antibody did not result in the expected BMSC apoptosis, regardless of concentration, suggesting a disruption of the Fas activation pathway. Thus BMSCs may have an endogenous mechanism to evade Fas-mediated apoptosis. Cumulatively, these data provide a parallel between adult stem/progenitor cells and cancer cells, consistent with the idea that stem/progenitor cells can use FasL to prevent lymphocyte attack by inducing lymphocyte apoptosis during the regeneration of injured tissues.Human bone marrow stromal cells (BMSCs)2 (also referred to as mesenchymal stem cells (MSCs)) (1) contain a subset of multipotent, non-hematopoietic stem/progenitor cells. BMSCs can differentiate into hematopoiesis-supporting stromal tissue, adipocytes, osteoblasts, and chondrocytes (2, 3). In addition, they may be able to transdifferentiate into hepatocytes, myocytes, neuroectodermal cells, and endothelial cells, (46) although proof of such differentiation is not definitive to date. BMSCs have immunosuppressive potential, as recently demonstrated in both in vitro (7) and in vivo (8, 9) systems, including clinical studies (10, 11). However, the mechanisms by which BMSCs suppress immune responses are unresolved. Soluble factor-mediated immunosuppressive effects are beginning to come to light, (10, 12), and in addition there are as yet unexplained effects of cell-to-cell contact.In the present study, we hypothesize that BMSC-mediated cytotoxicity of lymphocytes involves the FasL-activated apoptotic machinery. FasL is a type II transmembrane protein belonging to the tumor necrosis factor (TNF) family. FasL interacts with its receptor, Fas (CD95/APO-1) and triggers a cascade of subcellular events culminating in apoptotic cell death. FasL and Fas are key regulators of apoptosis in the immune system. In addition, FasL is expressed by cells in immune-privileged sites, such as cancer cells, neurons, eyes, cytotrophoblasts of the placenta, and reproductive organs (1317). In neurons, FasL expression specifically protects against T cell-mediated cytotoxicity (16).The discovery that FasL is also expressed by a variety of tumor cells raises the possibility that FasL may mediate immune privilege in human tumors (18). Activated T cells expressing Fas are sensitive to Fas-mediated apoptosis. Thus, up-regulation of FasL expression by tumor cells may enable tumorigenesis by targeting apoptosis in infiltrating lymphocytes. In the present work, we show that BMSCs can mediate immunosuppressive activity by FasL-induced killing of activated lymphocytes. Thus, BMSCs have properties of immune-privileged cells.  相似文献   

5.
6.
Drug resistance poses a major challenge to ovarian cancer treatment. Understanding mechanisms of drug resistance is important for finding new therapeutic targets. In the present work, a cisplatin-resistant ovarian cancer cell line A2780-DR was established with a resistance index of 6.64. The cellular accumulation of cisplatin was significantly reduced in A2780-DR cells as compared with A2780 cells consistent with the general character of drug resistance. Quantitative proteomic analysis identified 340 differentially expressed proteins between A2780 and A2780-DR cells, which involve in diverse cellular processes, including metabolic process, cellular component biogenesis, cellular processes, and stress responses. Expression levels of Ras-related proteins Rab 5C and Rab 11B in A2780-DR cells were lower than those in A2780 cells as confirmed by real-time quantitative PCR and Western blotting. The short hairpin (sh)RNA-mediated knockdown of Rab 5C in A2780 cells resulted in markedly increased resistance to cisplatin whereas overexpression of Rab 5C in A2780-DR cells increases sensitivity to cisplatin, demonstrating that Rab 5C-dependent endocytosis plays an important role in cisplatin resistance. Our results also showed that expressions of glycolytic enzymes pyruvate kinase, glucose-6-phosphate isomerase, fructose-bisphosphate aldolase, lactate dehydrogenase, and phosphoglycerate kinase 1 were down-regulated in drug resistant cells, indicating drug resistance in ovarian cancer is directly associated with a decrease in glycolysis. Furthermore, it was found that glutathione reductase were up-regulated in A2780-DR, whereas vimentin, HSP90, and Annexin A1 and A2 were down-regulated. Taken together, our results suggest that drug resistance in ovarian cancer cell line A2780 is caused by multifactorial traits, including the down-regulation of Rab 5C-dependent endocytosis of cisplatin, glycolytic enzymes, and vimentin, and up-regulation of antioxidant proteins, suggesting Rab 5C is a potential target for treatment of drug-resistant ovarian cancer. This constitutes a further step toward a comprehensive understanding of drug resistance in ovarian cancer.Ovarian cancer is the major cause of death in women with gynecological cancer. Early diagnosis of ovarian cancer is difficult, while its progression is fast. The standard treatment is surgical removal followed by platinum-taxane chemotherapy. However, the efficacy of the traditional surgery and chemotherapy is rather compromised and platinum resistant cancer recurs in ∼25% of patients within six months, and the overall five-year survival rate is about 31% (13). Virtually no efficient second line treatment is available. In order to increase survival rates from ovarian cancer and enhance patients'' quality of life, new therapeutic targets are urgently required, necessitating a deeper understanding of molecular mechanisms of drug resistance.Mechanisms of drug-resistance in ovarian cancer have been extensively studied over the last 30 years. Earlier studies have found that multiple factors are linked to drug resistance in human ovarian cancer including reduced intracellular drug accumulation, intracellular cisplatin inactivation, and increased DNA repair (4). Reduced cellular drug accumulation is mediated by the copper transporter-1 responsible for the influx of cisplatin (59) and MDR1, which encodes an integral membrane protein named P-glycoprotein for the active efflux of platinum drugs. Up-regulation of MDR1 has been observed in cisplatin-treated ovarian cancer cells although cisplatin is not a substrate of P-glycoprotein (1013). A fraction of intracellular cisplatin can be converted into cisplatin-thiol conjugates by glutathione-S-transferase (GST) π, leading to inactivation of cisplatin. Up-regulation of both GSTπ and γ-glutamylcysteine synthetase has been associated with cisplatin resistance in ovarian, cervical and lung cancer cell lines (1418). Binding of cisplatin to DNA leads to intrastrand or interstrand cross-links that alter the structure of the DNA molecule causing DNA damage. It has been amply documented that pathways for recognition and repair of damaged DNA are up-regulated in drug-resistant cancer cells (1926). Furthermore, the secondary mutations have been identified, which restore the wild-type BRCA2 reading frame enhancing the acquired resistance to platinum-based chemotherapy (24). Alternations in other signaling pathways have also been found in drug resistant ovarian cancer (2729). For example, DNA-PK phosphorylates RAC-alpha serine/threonine-protein kinase (AKT) and inhibits cisplatin-mediated apoptosis (28); and silencing of HDAC4 increases acetyl-STAT1 levels to prevent platinum-induced STAT1 activation and restore cisplatin sensitivity (29).Proteomics is playing an increasingly important role in identifying differentially expressed proteins between drug-resistant and drug sensitive ovarian cancer cells (3035). An earlier study has identified 57 differentially expressed proteins in human ovarian cancer cells and their platinum-resistant sublines, including annexin A3, destrin, cofilin 1, Glutathione-S-transferase omega 1, and cytosolic NADP+-dependent isocitrate dehydrogenase using 2D gel electrophoresis (30). Employing a similar 2D gel electrophoresis approach, changes in protein expressions of capsid glycoprotein, fructose-bisphosphate aldolase C, heterogeneous nuclear ribonucleoproteins A2/B1, putative RNA-binding protein 3, Ran-specific GTPase-activating protein, ubiquitin carboxyl-terminal hydrolase isozyme L1, stathmin, ATPSH protein, chromobox protein homolog3, and phosphoglycerate kinase 1 (PGK)1 were found in A2780 and drug-resistant A2780 cells (32). It is worth mentioning that ALDO and PGK are glycolytic enzymes, indicating that glycolysis plays a role in drug resistance. Studies have demonstrated that resistance to platinum drugs in ovarian cancer cells is linked to mitochondrial dysfunctions in oxidative phosphorylation and energy production (3640). Mitochondrial proteomic analysis of drug-resistant cells has shown that five mitochondrial proteins (ATP-a, PRDX3, PHB, ETF, and ALDH) that participate in the electron transport respiratory chain are down-regulated in drug-resistant cell lines (41). PRDX3 is involved in redox regulation of the cell to protect radical-sensitive enzymes from oxidative damage. However, it is not clear how down-regulation of PRDX3 is associated with drug-resistance. A more recent study showed that activated leukocyte cell adhesion molecule (ALCA) and A kinase anchoring protein 12 (AKAP12) are elevated in drug-resistant A2780-CP20 cells by quantifying the mitochondrial proteins (42). Despite these efforts, the drug-resistance mechanisms are not yet well understood.In this work, we established and characterized a drug-resistant cell line A2780-DR from A2780 cells. We employed a quantitative proteomic method to identify the differentially expressed proteins between A2780 and A2780-DR cells. Expression changes of selected proteins were confirmed by qPCR and Western blotting. We also used shRNA silencing to explore functions of Rab 5C and Rab 11B proteins in drug resistance. Our data indicate that the differentially expressed proteins participate in a variety of cellular processes and enhance our understanding of the mechanisms of drug resistance in ovarian cancer cells.  相似文献   

7.
8.
9.
10.
11.
12.
13.
Insulin plays a central role in the regulation of vertebrate metabolism. The hormone, the post-translational product of a single-chain precursor, is a globular protein containing two chains, A (21 residues) and B (30 residues). Recent advances in human genetics have identified dominant mutations in the insulin gene causing permanent neonatal-onset DM2 (14). The mutations are predicted to block folding of the precursor in the ER of pancreatic β-cells. Although expression of the wild-type allele would in other circumstances be sufficient to maintain homeostasis, studies of a corresponding mouse model (57) suggest that the misfolded variant perturbs wild-type biosynthesis (8, 9). Impaired β-cell secretion is associated with ER stress, distorted organelle architecture, and cell death (10). These findings have renewed interest in insulin biosynthesis (1113) and the structural basis of disulfide pairing (1419). Protein evolution is constrained not only by structure and function but also by susceptibility to toxic misfolding.Insulin plays a central role in the regulation of vertebrate metabolism. The hormone, the post-translational product of a single-chain precursor, is a globular protein containing two chains, A (21 residues) and B (30 residues). Recent advances in human genetics have identified dominant mutations in the insulin gene causing permanent neonatal-onset DM2 (14). The mutations are predicted to block folding of the precursor in the ER of pancreatic β-cells. Although expression of the wild-type allele would in other circumstances be sufficient to maintain homeostasis, studies of a corresponding mouse model (57) suggest that the misfolded variant perturbs wild-type biosynthesis (8, 9). Impaired β-cell secretion is associated with ER stress, distorted organelle architecture, and cell death (10). These findings have renewed interest in insulin biosynthesis (1113) and the structural basis of disulfide pairing (1419). Protein evolution is constrained not only by structure and function but also by susceptibility to toxic misfolding.  相似文献   

14.
Aldo-keto reductase family 1 member B10 (AKR1B10) is primarily expressed in the normal human colon and small intestine but overexpressed in liver and lung cancer. Our previous studies have shown that AKR1B10 mediates the ubiquitin-dependent degradation of acetyl-CoA carboxylase-α. In this study, we demonstrate that AKR1B10 is critical to cell survival. In human colon carcinoma cells (HCT-8) and lung carcinoma cells (NCI-H460), small-interfering RNA-induced AKR1B10 silencing resulted in caspase-3-mediated apoptosis. In these cells, the total and subspecies of cellular lipids, particularly of phospholipids, were decreased by more than 50%, concomitant with 2–3-fold increase in reactive oxygen species, mitochondrial cytochrome c efflux, and caspase-3 cleavage. AKR1B10 silencing also increased the levels of α,β-unsaturated carbonyls, leading to the 2–3-fold increase of cellular lipid peroxides. Supplementing the HCT-8 cells with palmitic acid (80 μm), the end product of fatty acid synthesis, partially rescued the apoptosis induced by AKR1B10 silencing, whereas exposing the HCT-8 cells to epalrestat, an AKR1B10 inhibitor, led to more than 2-fold elevation of the intracellular lipid peroxides, resulting in apoptosis. These data suggest that AKR1B10 affects cell survival through modulating lipid synthesis, mitochondrial function, and oxidative status, as well as carbonyl levels, being an important cell survival protein.Aldo-keto reductase family 1 member B10 (AKR1B10,2 also designated aldose reductase-like-1, ARL-1) is primarily expressed in the human colon, small intestine, and adrenal gland, with a low level in the liver (13). However, this protein is overexpressed in hepatocellular carcinoma, cervical cancer, lung squamous cell carcinoma, and lung adenocarcinoma in smokers, being a potential diagnostic and/or prognostic marker (1, 2, 46).The biological function of AKR1B10 in the intestine and adrenal gland, as well as its role in tumor development and progression, remains unclear. AKR1B10 is a monomeric enzyme that efficiently catalyzes the reduction to corresponding alcohols of a range of aromatic and aliphatic aldehydes and ketones, including highly electrophilic α,β-unsaturated carbonyls and antitumor drugs containing carbonyl groups, with NADPH as a co-enzyme (1, 712). The electrophilic carbonyls are constantly produced by lipid peroxidation, particularly in oxidative conditions, and are highly cytotoxic; through interaction with proteins, peptides, and DNA, the carbonyls cause protein dysfunction and DNA damage (breaks and mutations), resulting in mutagenesis, carcinogenesis, or apoptosis (10, 1319). AKR1B10 also shows strong enzymatic activity toward all-trans-retinal, 9-cis-retinal, and 13-cis-retinal, reducing them to the corresponding retinols, which may regulate the intracellular retinoic acid, a signaling molecule modulating cell proliferation and differentiation (6, 2023). In lung cancer, AKR1B10 expression is correlated with the patient smoking history and activates procarcinogens in cigarette smoke, such as polycyclic aromatic hydrocarbons, thus involved in lung tumorigenesis (2426).Recent studies have shown that in breast cancer cells, AKR1B10 associates with acetyl-CoA carboxylase-α (ACCA) and blocks its ubiquitination and proteasome degradation (27). ACCA is a rate-limiting enzyme of de novo synthesis of long chain fatty acids, catalyzing the ATP-dependent carboxylation of acetyl-CoA to form malonyl-CoA (28). Long chain fatty acids are the building blocks of biomembranes and the precursor of lipid second messengers, playing a critical role in cell growth and proliferation (29, 30). Therefore, ACCA activity is tightly regulated by both metabolite-mediated allosteric mechanisms and phosphorylation-dependent mechanisms; the latter are controlled by multiple hormones, such as insulin, glucagon, and growth factors (3133). ACCA activity is also regulated through physical protein-protein interaction. For instance, breast cancer 1 (BRCA1) protein associates with the ACCA and blocks its Ser79 residue from dephosphorylation (34, 35). The AKR1B10-mediated regulation on ACCA stability represents a novel regulatory mechanism, and this current study elucidated the biological significance of this regulation. The results show that AKR1B10 promotes cell survival via modulating lipid synthesis, mitochondrial function and oxidative stress, and carbonyl levels.  相似文献   

15.
A decoding algorithm is tested that mechanistically models the progressive alignments that arise as the mRNA moves past the rRNA tail during translation elongation. Each of these alignments provides an opportunity for hybridization between the single-stranded, -terminal nucleotides of the 16S rRNA and the spatially accessible window of mRNA sequence, from which a free energy value can be calculated. Using this algorithm we show that a periodic, energetic pattern of frequency 1/3 is revealed. This periodic signal exists in the majority of coding regions of eubacterial genes, but not in the non-coding regions encoding the 16S and 23S rRNAs. Signal analysis reveals that the population of coding regions of each bacterial species has a mean phase that is correlated in a statistically significant way with species () content. These results suggest that the periodic signal could function as a synchronization signal for the maintenance of reading frame and that codon usage provides a mechanism for manipulation of signal phase.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32]  相似文献   

16.
17.
18.
Ligation of major histocompatability complex class I (MHC-I) molecules expressed on T cells leads to both growth arrest and apoptosis. The aim of the current study was to investigate the intracellular signal pathways that mediate these effects.MHC-I ligation of human Jurkat T cells induced a morphologically distinct form of apoptosis within 6 h. A specific caspase inhibitor, which inhibited Fas-induced apoptosis, did not affect apoptosis induced by MHC-I ligation. Furthermore, MHC-I–induced apoptosis did not involve cleavage and activation of the poly(ADP- ribose) polymerase (PARP) endonuclease or degradation of genomic DNA into the typical fragmentation ladder, both prominent events of Fas-induced apoptosis. These results suggest that MHC-I ligation of Jurkat T cells induce apoptosis through a signal pathway distinct from the Fas molecule.In our search for other signal pathways leading to apoptosis, we found that the regulatory 85-kD subunit of the phosphoinositide-3 kinase (PI-3) kinase was tyrosine phosphorylated after ligation of MHC-I and the PI-3 kinase inhibitor wortmannin selectively blocked MHC-I–, but not Fas-induced, apoptosis. As the c-Jun NH2-terminal kinase (JNK) can be activated by PI-3 kinase activity, and has been shown to be involved in apoptosis of lymphocytes, we examined JNK activation after MHC-I ligation. Strong JNK activity was observed after MHC-I ligation and the activity was completely blocked by wortmannin. Inhibition of JNK activity, by transfecting cells with a dominant-negative JNKK– MKK4 construct, led to a strong reduction of apoptosis after MHC-I ligation. These results suggest a critical engagement of PI-3 kinase–induced JNK activity in apoptosis induced by MHC-I ligation.Apoptosis is an active form of cell death associated with certain characteristic morphological changes of the cell. These include cell shrinkage, condensation of chromatin, and usually, but not always, fragmentation of genomic DNA into specific oligonucleosomal fragments, also referred to as apoptotic DNA ladder (21). In addition, a morphologically distinct form of apoptosis has been described in germinal centers, thymocyte suspensions, and certain tumors with characteristic features of type B dark cells (7, 34). The condensed chromatin in these cells is not smoothly redistributed into the characteristic eye seen in the nucleus of classical apoptosis; the cytoplasm is darkened and the mitochondria and endoplasmatic reticulum tend to be swollen (7, 34).The mammalian interleukin-1β–convertase enzyme (ICE)1 protease family (caspases) are known to be critically involved in Fas- and tumor necrosis factor α–induced apoptosis (12). All caspases share two features: (a) they are synthesized as proenzymes and they are activated by cleavage at specific aspartate residues, and (b) both have the same consensus sequence as their own protease activity (14). Thus, the caspases are presumed to be regulated within a hierarchy of auto- and trans-cleavage. In Fas- mediated apoptosis, a sequential activation of caspase 1 (ICE) and caspase 3 (CPP32) has been demonstrated (13). Recent evidence suggests that one subfamily of caspases (inhibitable by YVAD pseudo-substrate) works proximally in Fas-induced apoptosis, leading to release of cytochrome C and other factors from the mitochondria. The initially activated caspases, together with the released factors from the mitochondria, activate the effector caspases (inhibitable by DEVD pseudo-substrate) by a mechanism that is not well defined (22). One of the targets of the effector caspases is the nuclear enzyme poly(ADP-ribose) polymerase (PARP) involved in DNA damage sensing. PARP activity is thought to be critical for apoptosis induced through caspase 3 (5, 43). The PARP enzyme is a 116-kD protein that is cleaved into a 85-kD fragment upon activation (43).Aurintricarboxylic acid (ATA), an inhibitor of Ca2+- dependent endonuclease activity (23, 29), has been shown to inhibit apoptosis that occurs without DNA fragmentation (25, 32). The mechanism by which ATA inhibits apoptosis is not fully understood. However, inhibition of endonuclease activity may not be the only function of ATA; rather, inhibition of topoisomerase II that induces chromatin condensation during apoptosis seems to be important (6). ATA has also been shown to inhibit the Ca2+-activated enzyme calpain, which may be involved in apoptosis (33).A new member of the mitogen-activated protein kinase (MAPK) superfamily designated c-Jun NH2-terminal kinase (JNK), has recently been identified (16). A signal pathway functionally independent from extracellular signal-regulated kinase (ERK), which involves JNKK–MKK4, activates JNK (11, 49, 52). JNK is activated by dual phosphorylation of a Thr-Pro-Tyr motif during apoptosis induced by UV light, heat shock, and ligation of the Fas antigen (8, 17, 45, 50). Costimulation of T cells with T cell antigen receptor complex (TCR–CD3) and CD28 ligation, or CD40 ligation of B cells also results in activation of JNK (36, 41). Thus, JNK activity is involved in processes leading to both cell death and differentiation/activation. It has been demonstrated that apoptosis can be regulated through a balance between members of the MAPK superfamily; e.g., high JNK and low ERK activities may lead to apoptosis, whereas high JNK and ERK activities prevent apoptosis (17, 51).Stimulation of T cells through the major histocompatability complex class I (MHC-I) molecule initiates a cascade of biochemical changes that can lead to either activation and growth or cell cycle arrest and apoptosis (1, 4, 31, 40, 44). One of the critical events that occurs in both cases is the activation of tyrosine kinases, resulting in tyrosine phosphorylation of a variety of proteins including phospholipase C-γ1 (PLC-γ1) and ZAP70 (38, 39). We have recently shown that tyrosine kinase activity appears to be critical for growth inhibition and apoptosis induced by ligation of MHC-I molecules (4, 38). The functional outcome of MHC-I ligation is tightly linked to regulation of peripheral T cell activity and tolerance induction (37, 42).In the present study we have investigated the intracellular signal pathway leading to apoptosis after MHC-I ligation of T cells, in an attempt to find a cause-and-effect relationship between the biochemical and functional consequences of MHC-I ligation. We present evidence that MHC-I induces apoptosis through a distinct pathway involving phosphoinositide-3 (PI-3) kinase–induced JNK activity.  相似文献   

19.
There is a mounting evidence of the existence of autoantibodies associated to cancer progression. Antibodies are the target of choice for serum screening because of their stability and suitability for sensitive immunoassays. By using commercial protein microarrays containing 8000 human proteins, we examined 20 sera from colorectal cancer (CRC) patients and healthy subjects to identify autoantibody patterns and associated antigens. Forty-three proteins were differentially recognized by tumoral and reference sera (p value <0.04) in the protein microarrays. Five immunoreactive antigens, PIM1, MAPKAPK3, STK4, SRC, and FGFR4, showed the highest prevalence in cancer samples, whereas ACVR2B was more abundant in normal sera. Three of them, PIM1, MAPKAPK3, and ACVR2B, were used for further validation. A significant increase in the expression level of these antigens on CRC cell lines and colonic mucosa was confirmed by immunoblotting and immunohistochemistry on tissue microarrays. A diagnostic ELISA based on the combination of MAPKAPK3 and ACVR2B proteins yielded specificity and sensitivity values of 73.9 and 83.3% (area under the curve, 0.85), respectively, for CRC discrimination after using an independent sample set containing 94 sera representative of different stages of progression and control subjects. In summary, these studies confirmed the presence of specific autoantibodies for CRC and revealed new individual markers of disease (PIM1, MAPKAPK3, and ACVR2B) with the potential to diagnose CRC with higher specificity and sensitivity than previously reported serum biomarkers.Colorectal cancer (CRC)1 is the second most prevalent cancer in the western world. The development of this disease takes decades and involves multiple genetic events. CRC remains a major cause of mortality in developed countries because most of the patients are diagnosed at advanced stages because of the reluctance to use highly invasive diagnostic tools like colonoscopy. Actually only a few proteins have been described as biomarkers in CRC (carcinoembryonic antigen (CEA), CA19.9, and CA125 (13)), although none of them is recommended for clinical screening (4). Proteomics analysis is actively used for the identification of new biomarkers. In previous studies, the use of two-dimensional DIGE and antibody microarrays allowed the identification of differentially expressed proteins in CRC tissue, including isoforms and post-translational modifications responsible for modifications in signaling pathways (58). Both approaches resulted in the identification of a collection of potential tumoral tissue biomarkers that is currently being investigated.However, the implementation of simpler, non-invasive methods for the early detection of CRC should be based on the identification of proteins or antibodies in serum or plasma (913). There is ample evidence of the existence of an immune response to cancer in humans as demonstrated by the presence of autoantibodies in cancer sera. Self-proteins (autoantigens) altered before or during tumor formation can elicit an immune response (1319). These tumor-specific autoantibodies can be detected at early cancer stages and prior to cancer diagnosis revealing a great potential as biomarkers (14, 15, 20). Tumor proteins can be affected by specific point mutations, misfolding, overexpression, aberrant glycosylation, truncation, or aberrant degradation (e.g. p53, HER2, NY-ESO1, or MUC1 (16, 2125)). In fact, a number of tumor-associated autoantigens (TAAs) were identified previously in different studies involving autoantibody screening in CRC (2628).Several approaches have been used to identify TAAs in cancer, including natural protein arrays prepared with fractions obtained from two-dimensional LC separations of tumoral samples (29, 30) or protein extracts from cancer cells or tissue (9, 31) probed by Western blot with patient sera, cancer tissue peptide libraries expressed as cDNA expression libraries for serological screening (serological analysis of recombinant cDNA expression libraries (SEREX)) (22, 32), or peptides expressed on the surface of phages in combination with microarrays (17, 18, 33, 34). However, these approaches suffer from several drawbacks. In some cases TAAs have to be isolated and identified from the reactive protein lysate by LC-MS techniques, or in the phage display approach, the reactive TAA could be a mimotope without a corresponding linear amino acid sequence. Moreover, cDNA libraries might not be representative of the protein expression levels in tumors as there is a poor correspondence between mRNA and protein levels.Protein arrays provide a novel platform for the identification of both autoantibodies and their respective TAAs for diagnostic purposes in cancer serum patients. They present some advantages. Proteins printed on the microarray are known “a priori,” avoiding the need for later identifications and the discovery of mimotopes. There is no bias in protein selection as the proteins are printed at a similar concentration. This should result in a higher sensitivity for biomarker identification (13, 35, 36).In this study, we used commercially available high density protein microarrays for the identification of autoantibody signatures and tumor-associated antigens in colorectal cancer. We screened 20 CRC patient and control sera with protein microarrays containing 8000 human proteins to identify the CRC-associated autoantibody repertoire and the corresponding TAAs. Autoantibody profiles that discriminated the different types of CRC metastasis were identified. Moreover a set of TAAs showing increased or decreased expression in cancer cell lines and paired tumoral tissues was found. Finally an ELISA was set up to test the ability of the most immunoreactive proteins to detect colorectal adenocarcinoma. On the basis of the antibody response, combinations of three antigens, PIM1, MAPKAPK3, and ACVR2B, showed a great potential for diagnosis.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号