Sample records for g2-chromatid breaks induced

  1. Dose--response of initial G2-chromatid breaks induced in normal human fibroblasts by heavy ions

    NASA Technical Reports Server (NTRS)

    Kawata, T.; Durante, M.; Furusawa, Y.; George, K.; Takai, N.; Wu, H.; Cucinotta, F. A.; Dicello, J. F. (Principal Investigator)

    2001-01-01

    PURPOSE: To investigate initial chromatid breaks in prematurely condensed G2 chromosomes following exposure to heavy ions of different LET. MATERIAL AND METHODS: Exponentially growing human fibroblast cells AG1522 were irradiated with gamma-rays, energetic carbon (13 keV/ microm, 80 keV/microm), silicon (55 keV/microm) and iron (140 keV/microm, 185keV/microm, 440keV/microm) ions. Chromosomes were prematurely condensed using calyculin-A. Initial chromatid-type and isochromatid breaks in G2 cells were scored. RESULTS: The dose response curves for total chromatid breaks were linear regardless of radiation type. The relative biological effectiveness (RBE) showed a LET-dependent increase, peaking around 2.7 at 55-80keV/microm and decreasing at higher LET. The dose response curves for isochromatid-type breaks were linear for high-LET radiations, but linear-quadratic for gamma-rays and 13 keV/microm carbon ions. The RBE for the induction of isochromatid breaks obtained from linear components increased rapidly between 13keV/microm (about 7) and 80keV/microm carbon (about 71), and decreased gradually until 440 keV/microm iron ions (about 66). CONCLUSIONS: High-LET radiations are more effective at inducing isochromatid breaks, while low-LET radiations are more effective at inducing chromatid-type breaks. The densely ionizing track structures of heavy ions and the proximity of sister chromatids in G2 cells result in an increase in isochromatid breaks.

  2. A comparison of G2 phase radiation-induced chromatid break kinetics using calyculin-PCC with those obtained using colcemid block.

    PubMed

    Bryant, Peter E; Mozdarani, Hossein

    2007-09-01

    To study the possible influence of cell-cycle delay on cells reaching mitosis during conventional radiation-induced chromatid break experiments using colcemid as a blocking agent, we have compared the chromatid break kinetics following a single dose of gamma rays (0.75 Gy) in metaphase CHO cells using calyculin-induced premature chromosome condensation (PCC), with those using colcemid block. Calyculin-induced PCC causes very rapid condensation of G2 cell chromosomes without the need for a cell to progress to mitosis, hence eliminating any effect of cell-cycle checkpoint on chromatid break frequency. We found that the kinetics of the exponential first-order decrease in chromatid breaks with time after irradiation was similar (not significantly different) between the two methods of chromosome condensation. However, use of the calyculin-PCC technique resulted in a slightly increased rate of disappearance of chromatid breaks and thus higher frequencies of breaks at 1.5 and 2.5 h following irradiation. We also report on the effect of the nucleoside analogue ara A on chromatid break kinetics using the two chromosome condensation techniques. Ara A treatment of cells abrogated the decrease in chromatid breaks with time, both using the calyculin-PCC and colcemid methods. We conclude that cell-cycle delay may be a factor determining the absolute frequency of chromatid breaks at various times following irradiation of cells in G2 phase but that the first-order disappearance of chromatid breaks with time and its abrogation by ara A are not significantly influenced by the G2 checkpoint.

  3. Induction and disappearance of G2 chromatid breaks in lymphocytes after low doses of low-LET gamma-rays and high-LET fast neutrons.

    PubMed

    Vral, A; Thierens, H; Baeyens, A; De Ridder, L

    2002-04-01

    To determine by means of the G2 assay the number of chromatid breaks induced by low-LET gamma-rays and high-LET neutrons, and to compare the kinetics of chromatid break rejoining for radiations of different quality. The G2 assay was performed on blood samples of four healthy donors who were irradiated with low-LET gamma-rays and high-LET neutrons. In a first set of experiments a dose-response curve for the formation of chromatid breaks was carried out for gamma-rays and neutrons with doses ranging between 0.1 and 0.5 Gy. In a second set of experiments, the kinetics of chromatid break formation and disappearance were investigated after a dose of 0.5 Gy using post-irradiation times ranging between 0.5 and 3.5 h. For the highest dose of 0.5 Gy, the number of isochromatid breaks was also scored. No significant differences in the number of chromatid breaks were observed between low-LET gamma-rays and high-LET neutrons for the four donors at any of the doses given. The dose-response curves for the formation of chromatid breaks are linear for both radiation qualities and RBEs = 1 were obtained. Scoring of isochromatid breaks at the highest dose of 0.5 Gy revealed that high-LET neutrons were, however, more effective at inducing isochromatid breaks (RBE = 6.2). The rejoining experiments further showed that the kinetics of disappearance of chromatid breaks following irradiation with low-LET gamma-rays or high-LET neutrons were not significantly different. Half-times of 0.92 h for gamma-rays and 0.84 h for neutrons were obtained. Applying the G2 assay, the results demonstrate that at low doses of irradiation, the induction as well as the disappearance of chromatid breaks is independent of the LET of the radiation qualities used (0.24 keV x microm(-1) 60Co gamma-rays and 20 keV x microm(-1) fast neutrons). As these radiation qualities produce the same initial number of double-strand breaks, the results support the signal model that proposes that chromatid breaks are the result

  4. Kinetics of chromatid break repair in G2-human fibroblasts exposed to low- and high-LET radiations

    NASA Technical Reports Server (NTRS)

    Kawata, T.; Durante, M.; George, K.; Furusawa, Y.; Gotoh, E.; Takai, N.; Wu, H.; Cucinotta, F. A.

    2001-01-01

    The purpose of this study is to determine the kinetics of chromatid break rejoining following exposure to radiations of different quality. Exponentially growing human fibroblast cells AG1522 were irradiated with gamma-rays, energetic carbon (290 MeV/u), silicon (490 MeV/u) and iron (200 MeV/u, 600 MeV/u). Chromosomes were prematurely condensed using calyculin A. Prematurely condensed chromosomes were collected after several post-irradiation incubation times, ranging from 5 to 600 minutes, and the number of chromatid breaks and exchanges in G2 cells were scored. The relative biological effectiveness (RBE) for initial chromatid breaks per unit dose showed LET dependency having a peak at 55 keV/micrometers silicon (2.4) or 80 keV/micrometers carbon particles (2.4) and then decreased with increasing LET. The kinetics of chromatid break rejoining following low- or high-LET irradiation consisted of two exponential components. Chromatid breaks decreased rapidly after exposure, and then continued to decrease at a slower rate. The rejoining kinetics was similar for exposure to each type of radiation, although the rate of unrejoined breaks was higher for high-LET radiation. Chromatid exchanges were also formed quickly.

  5. Rejoining of isochromatid breaks induced by heavy ions in G2-phase normal human fibroblasts

    NASA Technical Reports Server (NTRS)

    Kawata, T.; Durante, M.; Furusawa, Y.; George, K.; Ito, H.; Wu, H.; Cucinotta, F. A.

    2001-01-01

    We reported previously that exposure of normal human fibroblasts in G2 phase of the cell cycle to high-LET radiation produces a much higher frequency of isochromatid breaks than exposure to gamma rays. We concluded that an increase in the production of isochromatid breaks is a signature of initial high-LET radiation-induced G2-phase damage. In this paper, we report the repair kinetics of isochromatid breaks induced by high-LET radiation in normal G2-phase human fibroblasts. Exponentially growing human fibroblasts (AG1522) were irradiated with gamma rays or energetic carbon (290 MeV/nucleon), silicon (490 MeV/nucleon), or iron (200 MeV/nucleon) ions. Prematurely condensed chromosomes were induced by calyculin A after different postirradiation incubation times ranging from 0 to 600 min. Chromosomes were stained with Giemsa, and aberrations were scored in cells at G2 phase. G2-phase fragments, the result of the induction of isochromatid breaks, decreased quickly with incubation time. The curve for the kinetics of the rejoining of chromatid-type breaks showed a slight upward curvature with time after exposure to 440 keV/microm iron particles, probably due to isochromatid-isochromatid break rejoining. The formation of chromatid exchanges after exposure to high-LET radiation therefore appears to be underestimated, because isochromatid-isochromatid exchanges cannot be detected. Increased induction of isochromatid breaks and rejoining of isochromatid breaks affect the overall kinetics of chromatid-type break rejoining after exposure to high-LET radiation.

  6. G2 Chromatid Damage and Repair Kinetics in Normal Human Fibroblast Cells Exposed to Low-or High-LET Radiation

    NASA Technical Reports Server (NTRS)

    Kawata, T.; Ito, H.; Uno, T.; Saito, M.; Yamamoto, S.; Furusawa, Y.; Durante, M.; George, K.; Wu, H.; Cucinotta, F. A.

    2004-01-01

    Radiation-induced chromosome damage can be measured in interphase using the Premature Chromosome Condensation (PCC) technique. With the introduction of a new PCC technique using the potent phosphatase inhibitor calyculin-A, chromosomes can be condensed within five minutes, and it is now possible to examine the early damage induced by radiation. Using this method, it has been shown that high-LET radiation induces a higher frequency of chromatid breaks and a much higher frequency of isochromatid breaks than low-LET radiation. The kinetics of chromatid break rejoining consists of two exponential components representing a rapid and a slow time constant, which appears to be similar for low- and high- LET radiations. However, after high-LET radiation exposures, the rejoining process for isochromatid breaks influences the repair kinetics of chromatid-type breaks, and this plays an important role in the assessment of chromatid break rejoining in the G2 phase of the cell cycle.

  7. Induction by alkylating agents of sister chromatid exchanges and chromatid breaks in Fanconi's anemia.

    PubMed

    Latt, S A; Stetten, G; Juergens, L A; Buchanan, G R; Gerald, P S

    1975-10-01

    Sister chromatid exchanges, which may reflect chromosome repair in response to certain types of DNA damage, provide a means of investigating the increased chromosome fragility characteristic of Fanconi's anemia. By a recently developed technique using 33258 Hoechst and 5-bromodeoxyuridine, it was observed that the baseline frequency of sister chromatid exchanges in phytohemagglutinin-stimulated lymphocytes from four males with Fanconi's anemia differed little from that of normal lymphocytes. However, addition of the bifunctional alkylating agent mitomycin C (0.01 or 0.03 mug/ml) to the Fanconi's anemia cells during culture induces less than half of the increase in exchanges found in identically treated normal lymphocytes. This reduced increment in exchanges in accompanied by a partial suppression of mitosis and a marked increase in chromatid breaks and rearrangements. Many of these events occur at sites of incomplete chromatid interchange. The increase in sister chromatid exchanges induced in Fanconi's anemia lymphocytes by the monofunctional alkylating agent ethylmethane sulfonate (0.25 mg/ml) was slightly less than that in normal cells. Lymphocytes from two sets of parents of the patients with Fanconi's anemia exhibited a normal response to alkylating agents, while dermal fibroblasts from two different patients with Fanconi's anemia reacted to mitomycin C with an increase in chromatid breaks, but a nearly normal increment of sister chromatid exchanges. The results suggest that chromosomal breaks and rearrangements in Fanconi's anemia lymphocytes may result from a defect in a form of repair of DNA damage.

  8. High-LET radiation-induced aberrations in prematurely condensed G2 chromosomes of human fibroblasts

    NASA Technical Reports Server (NTRS)

    Kawata, T.; Gotoh, E.; Durante, M.; Wu, H.; George, K.; Furusawa, Y.; Cucinotta, F. A.; Dicello, J. F. (Principal Investigator)

    2000-01-01

    PURPOSE: To determine the number of initial chromatid breaks induced by low- or high-LET irradiations, and to compare the kinetics of chromatid break rejoining for radiations of different quality. MATERIAL AND METHODS: Exponentially growing human fibroblast cells AG1522 were irradiated with gamma-rays, energetic carbon (290MeV/u), silicon (490MeV/u) and iron (200 and 600 MeV/u). Chromosomes were prematurely condensed using calyculin A. Chromatid breaks and exchanges in G2 cells were scored. PCC were collected after several post-irradiation incubation times, ranging from 5 to 600 min. RESULTS: The kinetics of chromatid break rejoining following low- or high-LET irradiation consisted of two exponential components representing a rapid and a slow time constant. Chromatid breaks decreased rapidly during the first 10min after exposure, then continued to decrease at a slower rate. The rejoining kinetics were similar for exposure to each type of radiation. Chromatid exchanges were also formed quickly. Compared to low-LET radiation, isochromatid breaks were produced more frequently and the proportion of unrejoined breaks was higher for high-LET radiation. CONCLUSIONS: Compared with gamma-rays, isochromatid breaks were observed more frequently in high-LET irradiated samples, suggesting that an increase in isochromatid breaks is a signature of high-LET radiation exposure.

  9. The Rejoining Time of Chromatid Breaks Induced by Gamma Radiation in Vicia faba Root Tips at 3 °C

    PubMed Central

    Savage, J. R. K.; Neary, G. J.; Evans, H. J.

    1960-01-01

    The observation was made previously that the reduction in radiosensitivity in Vicia faba (as measured by postirradiation root growth) by prolonging the exposure time from about 10 minutes to 24 hours is much less marked at 3°C. than at 19°C. If chromosome damage is mainly responsible for the reduced root growth, this observation might be explained by a smaller drop in the "two-hit" aberration component, resulting from an increased time for which breaks are available for rejoining at 3°C. This hypothesis was tested by comparing chromatid aberration frequencies in root meristem cells produced by 105 rads of 60Co γ rays, given at dose rates of 19.4 and 0.073 rads per minute. Beans were maintained in aerated water at 2°C. prior to and during irradiation, and at this temperature the rate of development of cells was such that the two different exposure times both occupied a period during which the cell sensitivity was approximately constant. Immediately subsequent to irradiation, the roots were returned to 19°C. and examined cytologically. All chromatid aberrations were less frequent after low dose rate treatment, but only the chromatid interchange reduction was significant. The average time for which breaks are available for reunion, calculated from Lea's G function, was found to be 12 hours (95 per cent C.L. 6 to 24 hours). PMID:14442001

  10. Progress towards understanding the nature of chromatid breakage.

    PubMed

    Bryant, P E; Gray, L J; Peresse, N

    2004-01-01

    The wide range of sensitivities of stimulated T-cells from different individuals to radiation-induced chromatid breakage indicates the involvement of several low penetrance genes that appear to link elevated chromatid breakage to cancer susceptibility. The mechanisms of chromatid breakage are not yet fully understood. However, evidence is accumulating that suggests chromatid breaks are not simply expanded DNA double-strand breaks (DSB). Three models of chromatid breakage are considered. The classical breakage-first and the Revell "exchange" models do not accord with current evidence. Therefore a derivative of Revell's model has been proposed whereby both spontaneous and radiation-induced chromatid breaks result from DSB signaling and rearrangement processes from within large looped chromatin domains. Examples of such rearrangements can be observed by harlequin staining whereby an exchange of strands occurs immediately adjacent to the break site. However, these interchromatid rearrangements comprise less than 20% of the total breaks. The rest are thought to result from intrachromatid rearrangements, including a very small proportion involving complete excision of a looped domain. Work is in progress with the aim of revealing these rearrangements, which may involve the formation of inversions adjacent to the break sites. It is postulated that the disappearance of chromatid breaks with time results from the completion of such rearrangements, rather than from the rejoining of DSB. Elevated frequencies of chromatid breaks occur in irradiated cells with defects in both nonhomologous end-joining (NHEJ) and homologous recombination (HR) pathways, however there is little evidence of a correlation between reduced DSB rejoining and disappearance of chromatid breaks. Moreover, at least one treatment which abrogates the disappearance of chromatid breaks with time leaves DSB rejoining unaffected. The I-SceI DSB system holds considerable promise for the elucidation of these

  11. RPA Mediates Recruitment of MRX to Forks and Double-Strand Breaks to Hold Sister Chromatids Together.

    PubMed

    Seeber, Andrew; Hegnauer, Anna Maria; Hustedt, Nicole; Deshpande, Ishan; Poli, Jérôme; Eglinger, Jan; Pasero, Philippe; Gut, Heinz; Shinohara, Miki; Hopfner, Karl-Peter; Shimada, Kenji; Gasser, Susan M

    2016-12-01

    The Mre11-Rad50-Xrs2 (MRX) complex is related to SMC complexes that form rings capable of holding two distinct DNA strands together. MRX functions at stalled replication forks and double-strand breaks (DSBs). A mutation in the N-terminal OB fold of the 70 kDa subunit of yeast replication protein A, rfa1-t11, abrogates MRX recruitment to both types of DNA damage. The rfa1 mutation is functionally epistatic with loss of any of the MRX subunits for survival of replication fork stress or DSB recovery, although it does not compromise end-resection. High-resolution imaging shows that either the rfa1-t11 or the rad50Δ mutation lets stalled replication forks collapse and allows the separation not only of opposing ends but of sister chromatids at breaks. Given that cohesin loss does not provoke visible sister separation as long as the RPA-MRX contacts are intact, we conclude that MRX also serves as a structural linchpin holding sister chromatids together at breaks. Copyright © 2016 Elsevier Inc. All rights reserved.

  12. Replication-Dependent Sister Chromatid Recombination in Rad1 Mutants of Saccharomyces Cerevisiae

    PubMed Central

    Kadyk, L. C.; Hartwell, L. H.

    1993-01-01

    Homolog recombination and unequal sister chromatid recombination were monitored in rad1-1/rad1-1 diploid yeast cells deficient for excision repair, and in control cells, RAD1/rad1-1, after exposure to UV irradiation. In a rad1-1/rad1-1 diploid, UV irradiation stimulated much more sister chromatid recombination relative to homolog recombination when cells were irradiated in the G(1) or the G(2) phases of the cell cycle than was observed in RAD1/rad1-1 cells. Since sister chromatids are not present during G(1), this result suggested that unexcised lesions can stimulate sister chromatid recombination events during or subsequent to DNA replication. The results of mating rescue experiments suggest that unexcised UV dimers do not stimulate sister chromatid recombination during the G(2) phase, but only when they are present during DNA replication. We propose that there are two types of sister chromatid recombination in yeast. In the first type, unexcised UV dimers and other bulky lesions induce sister chromatid recombination during DNA replication as a mechanism to bypass lesions obstructing the passage of DNA polymerase, and this type is analogous to the type of sister chromatid exchange commonly observed cytologically in mammalian cells. In the second type, strand scissions created by X-irradiation or the excision of damaged bases create recombinogenic sites that result in sister chromatid recombination directly in G(2). Further support for the existence of two types of sister chromatid recombination is the fact that events induced in rad1-1/rad1-1 were due almost entirely to gene conversion, whereas those in RAD1/rad1-1 cells were due to a mixture of gene conversion and reciprocal recombination. PMID:8454200

  13. Radiation induces premature chromatid separation via the miR-142-3p/Bod1 pathway in carcinoma cells.

    PubMed

    Pan, Dong; Du, Yarong; Ren, Zhenxin; Chen, Yaxiong; Li, Xiaoman; Wang, Jufang; Hu, Burong

    2016-09-13

    Radiation-induced genomic instability plays a vital role in carcinogenesis. Bod1 is required for proper chromosome biorientation, and Bod1 depletion increases premature chromatid separation. MiR-142-3p influences cell cycle progression and inhibits proliferation and invasion in cervical carcinoma cells. We found that radiation induced premature chromatid separation and altered miR-142-3p and Bod1 expression in 786-O and A549 cells. Overexpression of miR-142-3p increased premature chromatid separation and G2/M cell cycle arrest in 786-O cells by suppressing Bod1 expression. We also found that either overexpression of miR-142-3p or knockdown of Bod1 sensitized 786-O and A549 cells to X-ray radiation. Overexpression of Bod1 inhibited radiation- and miR-142-3p-induced premature chromatid separation and increased resistance to radiation in 786-O and A549 cells. Taken together, these results suggest that radiation alters miR-142-3p and Bod1 expression in carcinoma cells, and thus contributes to early stages of radiation-induced genomic instability. Combining ionizing radiation with epigenetic regulation may help improve cancer therapies.

  14. Enhanced G2 chromatid radiosensitivity, an early stage in the neoplastic transformation of human epidermal keratinocytes in culture

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Gantt, R.; Sanford, K.K.; Parshad, R.

    1987-03-01

    A deficiency in DNA repair, manifest as enhanced chromatid radiosensitivity during the G2 phase of the cell cycle, together with a proliferative stimulus such as that provided by active oncogenes may be necessary and sufficient for the malignant neoplastic transformation of human keratinocytes in culture. Normal epidermal keratinocytes established as continuous cell lines by transfection with pSV3-neo or infection with adeno 12-SV40 hybrid virus developed enhanced G2 chromatid radiosensitivity after 18 passages in culture. In contrast to cells from primary or secondary culture, these cells could be transformed to malignant neoplastic cells by infection with Kirsten murine sarcoma virus containingmore » the Ki-ras oncogene or in one line by the chemical carcinogen, N-methyl-N'-nitro-N-nitrosoguanidine; both of these agents produced a marked proliferative response. Cytological heterogeneity and karyotypic instability characterized the cells during their progression to neoplasia. These results are interpreted in terms of a mechanism for neoplastic transformation.« less

  15. Induction of chromatin damage and distribution of isochromatid breaks in human fibroblast cells exposed to heavy ions

    NASA Technical Reports Server (NTRS)

    Kawata, Tetsuya; Ito, Hisao; Motoori, Ken; Ueda, Takuya; Shigematsu, Naoyuki; Furusawa, Yoshiya; Durante, Marco; George, Kerry; Wu, Honglu; Cucinotta, Francis A.

    2002-01-01

    The frequency of chromatid breaks and the distribution of isochromatid breaks were measured in G2-phase normal human fibroblasts prematurely condensed a short time after exposure to low- or high-LET radiations. The average number of isochromatid breaks from a single particle traversal increased with increasing LET values, while the average number of chromatid-type breaks appeared to reach a plateau. The distribution of isochromatid breaks after high-LET iron particles exposure was overdispersed compared to gamma-rays, indicating that a single iron particle traversal through a cell nucleus can produce multiple isochromatid breaks.

  16. Induction of chromatin damage and distribution of isochromatid breaks in human fibroblast cells exposed to heavy ions.

    PubMed

    Kawata, Tetsuya; Ito, Hisao; Motoori, Ken; Ueda, Takuya; Shigematsu, Naoyuki; Furusawa, Yoshiya; Durante, Marco; George, Kerry; Wu, Honglu; Cucinotta, Francis A

    2002-12-01

    The frequency of chromatid breaks and the distribution of isochromatid breaks were measured in G2-phase normal human fibroblasts prematurely condensed a short time after exposure to low- or high-LET radiations. The average number of isochromatid breaks from a single particle traversal increased with increasing LET values, while the average number of chromatid-type breaks appeared to reach a plateau. The distribution of isochromatid breaks after high-LET iron particles exposure was overdispersed compared to gamma-rays, indicating that a single iron particle traversal through a cell nucleus can produce multiple isochromatid breaks.

  17. Faithful anaphase is ensured by Mis4, a sister chromatid cohesion molecule required in S phase and not destroyed in G1 phase

    PubMed Central

    Furuya, Kanji; Takahashi, Kohta; Yanagida, Mitsuhiro

    1998-01-01

    The loss of sister chromatid cohesion triggers anaphase spindle movement. The budding yeast Mcd1/Scc1 protein, called cohesin, is required for associating chromatids, and proteins homologous to it exist in a variety of eukaryotes. Mcd1/Scc1 is removed from chromosomes in anaphase and degrades in G1. We show that the fission yeast protein, Mis4, which is required for equal sister chromatid separation in anaphase is a different chromatid cohesion molecule that behaves independent of cohesin and is conserved from yeast to human. Its inactivation in G1 results in cell lethality in S phase and subsequent premature sister chromatid separation. Inactivation in G2 leads to cell death in subsequent metaphase–anaphase progression but missegregation occurs only in the next round of mitosis. Mis4 is not essential for condensation, nor does it degrade in G1. Rather, it associates with chromosomes in a punctate fashion throughout the cell cycle. mis4 mutants are hypersensitive to hydroxyurea (HU) and UV irradiation but retain the ability to restrain cell cycle progression when damaged or sustaining a block to replication. The mis4 mutation results in synthetic lethality with a DNA ligase mutant. Mis4 may form a stable link between chromatids in S phase that is split rather than removed in anaphase. PMID:9808627

  18. Histone H3 K79 methylation states play distinct roles in UV-induced sister chromatid exchange and cell cycle checkpoint arrest in Saccharomyces cerevisiae

    PubMed Central

    Rossodivita, Alyssa A.; Boudoures, Anna L.; Mecoli, Jonathan P.; Steenkiste, Elizabeth M.; Karl, Andrea L.; Vines, Eudora M.; Cole, Arron M.; Ansbro, Megan R.; Thompson, Jeffrey S.

    2014-01-01

    Histone post-translational modifications have been shown to contribute to DNA damage repair. Prior studies have suggested that specific H3K79 methylation states play distinct roles in the response to UV-induced DNA damage. To evaluate these observations, we examined the effect of altered H3K79 methylation patterns on UV-induced G1/S checkpoint response and sister chromatid exchange (SCE). We found that the di- and trimethylated states both contribute to activation of the G1/S checkpoint to varying degrees, depending on the synchronization method, although methylation is not required for checkpoint in response to high levels of UV damage. In contrast, UV-induced SCE is largely a product of the trimethylated state, which influences the usage of gene conversion versus popout mechanisms. Regulation of H3K79 methylation by H2BK123 ubiquitylation is important for both checkpoint function and SCE. H3K79 methylation is not required for the repair of double-stranded breaks caused by transient HO endonuclease expression, but does play a modest role in survival from continuous exposure. The overall results provide evidence for the participation of H3K79 methylation in UV-induced recombination repair and checkpoint activation, and further indicate that the di- and trimethylation states play distinct roles in these DNA damage response pathways. PMID:24748660

  19. Complex chromatid-isochromatid exchanges following irradiation with heavy ions?

    PubMed

    Loucas, B D; Eberle, R L; Durante, M; Cornforth, M N

    2004-01-01

    We describe a peculiar and relatively rare type of chromosomal rearrangement induced in human peripheral lymphocytes that were ostensibly irradiated in G(0) phase of the cell cycle by accelerated heavy ions, and which, to the best of our knowledge, have not been previously described. The novel rearrangements which were detected using mFISH following exposure to 500 MeV/nucleon and 5 GeV/n 56Fe particles, but were not induced by either 137Cs gamma rays or 238Pu alpha particles, can alternatively be described as either complex chromatid-isochromatid or complex chromatid-chromosome exchanges. Different mechanisms potentially responsible for their formation are discussed. Copyright 2003 S. Karger AG, Basel

  20. Cut2 proteolysis required for sister-chromatid seperation in fission yeast.

    PubMed

    Funabiki, H; Yamano, H; Kumada, K; Nagao, K; Hunt, T; Yanagida, M

    1996-05-30

    Although mitotic cyclins are well-known substrates for ubiquitin-mediated proteolysis at the metaphase-anaphase transition, their degradation is not essential for separation of sister chromatids; several lines of evidence suggest that proteolysis of other protein(s) is required, however. Here we report the anaphase-specific proteolysis of the Schizosaccharomyces pombe Cut2 protein, which is essential for sister-chromatid separation. Cut2 is located in the nucleus, where it is concentrated along the short metaphase spindle. The rapid degradation of Cut2 at anaphase requires its amino-terminal region and the activity of Cut9 (ref. 14), a component of the 20S cyclosome/anaphase-promoting complex (APC), which is necessary for cyclin destruction. Expression of non-degradable Cut2 blocks sister-chromatid separation but not cell-cycle progression. This defect can be overcome by grafting the N terminus of cyclin B onto the truncated Cut2, demonstrating that the regulated proteolysis of Cut2 is essential for sister-chromatid separation.

  1. Ultraviolet-induced sister chromatid exchanges in V-79 cells with normal and BrdUrd-substituted DNA and the influence of intercalating substances and cysteine

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Speit, G.; Mehnert, K.; Wolf, M.

    1982-06-01

    The influence of intercalating substances (proflavine, ethidium bromide) and of an SH compound (L-cysteine) on uv-induced sister chromatid exchanges (SCEs) was investigated in V-79 cells with normal and BrdUrd-substituted DNA. The results are discussed in relation to the primary damages leading to SCE induction produced by uv irradiation. The data indicate that neither the pyrimidine dimers nor DNA single-strand breaks are the primary cause of SCE induction, and that the damages leading to SCEs by uv irradiation differ from those which cause chromosome aberrations.

  2. Double strand breaks and cell-cycle arrest induced by the cyanobacterial toxin cylindrospermopsin in HepG2 cells.

    PubMed

    Alja, Štraser; Filipič, Metka; Novak, Matjaž; Žegura, Bojana

    2013-08-21

    The newly emerging cyanobacterial cytotoxin cylindrospermopsin (CYN) is increasingly found in surface freshwaters, worldwide. It poses a potential threat to humans after chronic exposure as it was shown to be genotoxic in a range of test systems and is potentially carcinogenic. However, the mechanisms of CYN toxicity and genotoxicity are not well understood. In the present study CYN induced formation of DNA double strand breaks (DSBs), after prolonged exposure (72 h), in human hepatoma cells, HepG2. CYN (0.1-0.5 µg/mL, 24-96 h) induced morphological changes and reduced cell viability in a dose and time dependent manner. No significant increase in lactate dehydrogenase (LDH) leakage could be observed after CYN exposure, indicating that the reduction in cell number was due to decreased cell proliferation and not due to cytotoxicity. This was confirmed by imunocytochemical analysis of the cell-proliferation marker Ki67. Analysis of the cell-cycle using flow-cytometry showed that CYN has an impact on the cell cycle, indicating G0/G1 arrest after 24 h and S-phase arrest after longer exposure (72 and 96 h). Our results provide new evidence that CYN is a direct acting genotoxin, causing DSBs, and these facts need to be considered in the human health risk assessment.

  3. Chromosomal Integrity after UV Irradiation Requires FANCD2-Mediated Repair of Double Strand Breaks.

    PubMed

    Federico, María Belén; Vallerga, María Belén; Radl, Analía; Paviolo, Natalia Soledad; Bocco, José Luis; Di Giorgio, Marina; Soria, Gastón; Gottifredi, Vanesa

    2016-01-01

    Fanconi Anemia (FA) is a rare autosomal recessive disorder characterized by hypersensitivity to inter-strand crosslinks (ICLs). FANCD2, a central factor of the FA pathway, is essential for the repair of double strand breaks (DSBs) generated during fork collapse at ICLs. While lesions different from ICLs can also trigger fork collapse, the contribution of FANCD2 to the resolution of replication-coupled DSBs generated independently from ICLs is unknown. Intriguingly, FANCD2 is readily activated after UV irradiation, a DNA-damaging agent that generates predominantly intra-strand crosslinks but not ICLs. Hence, UV irradiation is an ideal tool to explore the contribution of FANCD2 to the DNA damage response triggered by DNA lesions other than ICL repair. Here we show that, in contrast to ICL-causing agents, UV radiation compromises cell survival independently from FANCD2. In agreement, FANCD2 depletion does not increase the amount of DSBs generated during the replication of UV-damaged DNA and is dispensable for UV-induced checkpoint activation. Remarkably however, FANCD2 protects UV-dependent, replication-coupled DSBs from aberrant processing by non-homologous end joining, preventing the accumulation of micronuclei and chromatid aberrations including non-homologous chromatid exchanges. Hence, while dispensable for cell survival, FANCD2 selectively safeguards chromosomal stability after UV-triggered replication stress.

  4. Genes on chromosomes 1 and 4 in the mouse are associated with repair of radiation-induced chromatin damage.

    PubMed

    Potter, M; Sanford, K K; Parshad, R; Tarone, R E; Price, F M; Mock, B; Huppi, K

    1988-04-01

    Early-passage skin fibroblasts from different inbred and congenic strains of mice were X-irradiated (1 Gy), and the number of chromatid breaks was determined at 2.0 h after irradiation. The cells from DBA/2N, C3H/HeN, STS/A, C57BL/6N, BALB/cJ, and AKR/N had 25 to 42 chromatid breaks per 100 metaphase cells (efficient repair phenotype). NZB/NJ had greater than 78 and BALB/cAn had 87 to 110 chromatid breaks per 100 cells (inefficient repair phenotype). Differences between BALB/cAn and BALB/c. DBA/2 congenic strains which carry less than 1% of the DBA/2 genome indicate that two genes, one on chromosome 1 linked to bcl-2-Pep-3 and the other on chromosome 4 closely linked to Fv-1, affect the efficiency with which the cells repair radiation-induced chromatin damage.

  5. Mitomycin C-induced pairing of heterochromatin reflects initiation of DNA repair and chromatid exchange formation.

    PubMed

    Abdel-Halim, H I; Natarajan, A T; Mullenders, L H F; Boei, J J W A

    2005-04-15

    Chromatid interchanges induced by the DNA cross-linking agent mitomycin C (MMC) are over-represented in human chromosomes containing large heterochromatic regions. We found that nearly all exchange breakpoints of chromosome 9 are located within the paracentromeric heterochromatin and over 70% of exchanges involving chromosome 9 are between its homologues. We provide evidence that the required pairing of chromosome 9 heterochromatic regions occurs in G(0)/G(1) and S-phase cells as a result of an active cellular process initiated upon MMC treatment. By contrast, no pairing was observed for a euchromatic paracentromeric region of the equal-sized chromosome 8. The MMC-induced pairing of chromosome 9 heterochromatin is observed in a subset of cells; its percentage closely mimics the frequency of homologous interchanges found at metaphase. Moreover, the absence of pairing in cells derived from XPF patients correlates with an altered spectrum of MMC-induced exchanges. Together, the data suggest that the heterochromatin-specific pairing following MMC treatment reflects the initiation of DNA cross-link repair and the formation of exchanges.

  6. INDUCTION OF DNA STRAND BREAKS BY TRIHALOMETHANES IN PRIMARY HUMAN LUNG EPITHELIAL CELLS

    EPA Science Inventory


    Abstract

    Trihalomethanes (TEMs) are disinfection by-products and suspected human carcinogens present in chlorinated drinking water. Previous studies have shown that many THMs induce sister chromatid exchanges and DNA strand breaks in human peripheral blood lymphocyte...

  7. Genes on chromosomes 1 and 4 in the mouse are associated with repair of radiation-induced chromatin damage

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Potter, M.; Sanford, K.K.; Parshad, R.

    Early-passage skin fibroblasts from different inbred and congenic strains of mice were X-irradiated (1 Gy), and the number of chromatid breaks was determined at 2.0 h after irradiation. The cells from DBA/2N, C3H/HeN, STS/A, C57BL/6N, BALB/cJ, and AKR/N had 25 to 42 chromatid breaks per 100 metaphase cells (efficient repair phenotype). NZB/NJ had greater than 78 and BALB/cAn had 87 to 110 chromatid breaks per 100 cells (inefficient repair phenotype). Differences between BALB/cAn and BALB/c. DBA/2 congenic strains which carry less than 1% of the DBA/2 genome indicate that two genes, one on chromosome 1 linked to bcl-2-Pep-3 and themore » other on chromosome 4 closely linked to Fv-1, affect the efficiency with which the cells repair radiation-induced chromatin damage.« less

  8. "Breaking up is hard to do": the formation and resolution of sister chromatid intertwines.

    PubMed

    Baxter, Jonathan

    2015-02-13

    The absolute necessity to resolve every intertwine between the two strands of the DNA double helix provides a massive challenge to the cellular processes that duplicate and segregate chromosomes. Although the overwhelming majority of intertwines between the parental DNA strands are resolved during DNA replication, there are numerous chromosomal contexts where some intertwining is maintained into mitosis. These mitotic sister chromatid intertwines (SCIs) can be found as; short regions of unreplicated DNA, fully replicated and intertwined sister chromatids--commonly referred to as DNA catenation--and as sister chromatid linkages generated by homologous recombination-associated processes. Several overlapping mechanisms, including intra-chromosomal compaction, topoisomerase action and Holliday junction resolvases, ensure that all SCIs are removed before they can prevent normal chromosome segregation. Here, I discuss why some DNA intertwines persist into mitosis and review our current knowledge of the SCI resolution mechanisms that are employed in both prokaryotes and eukaryotes, including how deregulating SCI formation during DNA replication or disrupting the resolution processes may contribute to aneuploidy in cancer. Copyright © 2014 Elsevier Ltd. All rights reserved.

  9. Chromosomal Integrity after UV Irradiation Requires FANCD2-Mediated Repair of Double Strand Breaks

    PubMed Central

    Federico, María Belén; Vallerga, María Belén; Radl, Analía; Paviolo, Natalia Soledad; Bocco, José Luis; Di Giorgio, Marina; Soria, Gastón; Gottifredi, Vanesa

    2016-01-01

    Fanconi Anemia (FA) is a rare autosomal recessive disorder characterized by hypersensitivity to inter-strand crosslinks (ICLs). FANCD2, a central factor of the FA pathway, is essential for the repair of double strand breaks (DSBs) generated during fork collapse at ICLs. While lesions different from ICLs can also trigger fork collapse, the contribution of FANCD2 to the resolution of replication-coupled DSBs generated independently from ICLs is unknown. Intriguingly, FANCD2 is readily activated after UV irradiation, a DNA-damaging agent that generates predominantly intra-strand crosslinks but not ICLs. Hence, UV irradiation is an ideal tool to explore the contribution of FANCD2 to the DNA damage response triggered by DNA lesions other than ICL repair. Here we show that, in contrast to ICL-causing agents, UV radiation compromises cell survival independently from FANCD2. In agreement, FANCD2 depletion does not increase the amount of DSBs generated during the replication of UV-damaged DNA and is dispensable for UV-induced checkpoint activation. Remarkably however, FANCD2 protects UV-dependent, replication-coupled DSBs from aberrant processing by non-homologous end joining, preventing the accumulation of micronuclei and chromatid aberrations including non-homologous chromatid exchanges. Hence, while dispensable for cell survival, FANCD2 selectively safeguards chromosomal stability after UV-triggered replication stress. PMID:26765540

  10. Effect of borax on immune cell proliferation and sister chromatid exchange in human chromosomes

    PubMed Central

    Pongsavee, Malinee

    2009-01-01

    Background Borax is used as a food additive. It becomes toxic when accumulated in the body. It causes vomiting, fatigue and renal failure. Methods The heparinized blood samples from 40 healthy men were studied for the impact of borax toxicity on immune cell proliferation (lymphocyte proliferation) and sister chromatid exchange in human chromosomes. The MTT assay and Sister Chromatid Exchange (SCE) technic were used in this experiment with the borax concentrations of 0.1, 0.15, 0.2, 0.3 and 0.6 mg/ml. Results It showed that the immune cell proliferation (lymphocyte proliferation) was decreased when the concentrations of borax increased. The borax concentration of 0.6 mg/ml had the most effectiveness to the lymphocyte proliferation and had the highest cytotoxicity index (CI). The borax concentrations of 0.15, 0.2, 0.3 and 0.6 mg/ml significantly induced sister chromatid exchange in human chromosomes (P < 0.05). Conclusion Borax had effects on immune cell proliferation (lymphocyte proliferation) and induced sister chromatid exchange in human chromosomes. Toxicity of borax may lead to cellular toxicity and genetic defect in human. PMID:19878537

  11. Sister chromatid exchanges induced by inhaled anesthetics

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    White,A.E.; Takehisa, S.; Eger II, E.I.

    1970-05-01

    There is sufficient evidence that anesthetics may cause cancer to justify a test of their carcinogenic potential. Baden et al., using the Ames test, a rapid and inexpensive genetic indicator of carcinogenicity, have shown that among currently used anesthetics fluorxene alone caused bacterial mutations. The authors used the sister chromatid exchange (SCE) technique, another rapid assay of mutagenic-carcinogenic potential. The frequency of sister chromatid exchanges in Chinese hamster ovary cells increases when the cell cultures are exposed to mutagen-carcinogens, particulary in the presence of a metabolic activating system. With this test system a one-hour exposure to 1 MAC nitrous oxide,more » diethyl ether, trichloroethylene, halothane, enflurane, isoflurane, methoxyflurane, or chloroform did not increase SCE values. Divinyl ether, fluroxene and ethyl vinyl ether increased SCE values in the same circumstances. Results of this study of mammalian cells suggest that no currently used anesthetic is a mutagen-carcinogen. The results also suggest that anesthetics containing a vinyl moiety may be mutagen-carcinogens.« less

  12. Zinc Chromate Induces Chromosome Instability and DNA Double Strand Breaks in Human Lung Cells

    PubMed Central

    Xie, Hong; Holmes, Amie L.; Young, Jamie L.; Qin, Qin; Joyce, Kellie; Pelsue, Stephen C.; Peng, Cheng; Wise, Sandra S.; Jeevarajan, Antony S.; Wallace, William T.; Hammond, Dianne; Wise, John Pierce

    2014-01-01

    Hexavalent chromium Cr(VI) is a respiratory toxicant and carcinogen, with solubility playing an important role in its carcinogenic potential. Zinc chromate, a water insoluble or ‘particulate’ Cr(VI) compound, has been shown to be carcinogenic in epidemiology studies and to induce tumors in experimental animals, but its genotoxicity is poorly understood. Our study shows that zinc chromate induced concentration-dependent increases in cytotoxicity, chromosome damage and DNA double strand breaks in human lung cells. In response to zinc chromate-induced breaks, MRE11 expression was increased and ATM and ATR were phosphorylated, indicating that the DNA double strand break repair system was initiated in the cells. In addition, our data show that zinc chromate-induced double strand breaks were only observed in the G2/M phase population, with no significant amount of double strand breaks observed in G1 and S phase cells. These data will aid in understanding the mechanisms of zinc chromate toxicity and carcinogenesis. PMID:19027772

  13. Genotoxic assessment in peripheral blood lymphocytes of post-polio individuals using sister chromatid exchange analysis and micronucleus assay.

    PubMed

    Bhattacharya, Saurabh Kumar; Saraswathy, Radha; Sivakumar, E

    2011-07-01

    Environmental pollution is a complex issue because of the diversity of anthropogenic agents, both chemical and physical, that have been detected and catalogued. The consequences to biota from exposure to genotoxic agents present an additional problem because of the potential for these agents to produce adverse change at the cellular and organism levels. Past studies in virus have focused on structural damage to the DNA of environmental species that may occur after exposure to genotoxic agents and the use of this information to document exposure and to monitor remediation. In an effort to predict effects at the population, community and ecosystem levels, in the present study, we attempt to characterize damage occurring through genotoxic agents like 5-bromo-2-deoxyuridine, BrdU, using sister chromatid exchange technique and the formation of micronuclei (MN) in the peripheral lymphocytes of the post-polio syndrome sequelae affected by poliovirus. Analysis of structural chromosomal aberrations (CAs) and involvement of the specific chromosome break were pursued in this study. They revealed a significantly higher incidence of CAs (chromatid and chromosome breaks) in patients compared with controls, where the specific chromosome break has emerged as specific. Also, the maximum numbers of breaks were found to be in chromosome 1 at the position 1p36.1. The results also suggest a correlation between CAs and content of MN.

  14. 6-gingerol prevents patulin-induced genotoxicity in HepG2 cells.

    PubMed

    Yang, Guang; Zhong, Laifu; Jiang, Liping; Geng, Chengyan; Cao, Jun; Sun, Xiance; Liu, Xiaofang; Chen, Min; Ma, Yufang

    2011-10-01

    Patulin (PAT) is a mycotoxin produced by several Penicillium, Aspergillus and Byssochlamys species. Since PAT is a potent genotoxic compound, and PAT contamination is common in fruits and fruit products, the search for newer, better agents for protection against genotoxicity of PAT is required. In this study, the chemoprotective effect of 6-gingerol against PAT-induced genotoxicity in HepG2 cells was investigated. The comet assay and micronucleus test (MNT) were used to monitor genotoxic effects. To further elucidate the underlying mechanisms, the intracellular generation of reactive oxygen species (ROS) and level of reduced glutathione (GSH) were tested. In addition, the level of oxidative DNA damage was evaluated by immunocytochemical analysis of 8-hydroxydeoxyguanosine (8-OHdG). The results showed that 6-gingerol significantly reduced the DNA strand breaks and micronuclei formation caused by PAT. Moreover, 6-gingerol effectively suppressed PAT-induced intracellular ROS formation and 8-OHdG level. The GSH depletion induced by PAT in HepG2 cells was also attenuated by 6-gingerol pretreatment. These findings suggest that 6-gingerol has a strong protective ability against the genotoxicity caused by PAT, and the antioxidant activity of 6-gingerol may play an important part in attenuating the genotoxicity of PAT. Copyright © 2011 John Wiley & Sons, Ltd.

  15. Defects in the Fanconi Anemia Pathway and Chromatid Cohesion in Head and Neck Cancer.

    PubMed

    Stoepker, Chantal; Ameziane, Najim; van der Lelij, Petra; Kooi, Irsan E; Oostra, Anneke B; Rooimans, Martin A; van Mil, Saskia E; Brink, Arjen; Dietrich, Ralf; Balk, Jesper A; Ylstra, Bauke; Joenje, Hans; Feller, Stephan M; Brakenhoff, Ruud H

    2015-09-01

    Failure to repair DNA damage or defective sister chromatid cohesion, a process essential for correct chromosome segregation, can be causative of chromosomal instability (CIN), which is a hallmark of many types of cancers. We investigated how frequent this occurs in head and neck squamous cell carcinoma (HNSCC) and whether specific mechanisms or genes could be linked to these phenotypes. The genomic instability syndrome Fanconi anemia is caused by mutations in any of at least 16 genes regulating DNA interstrand crosslink (ICL) repair. Since patients with Fanconi anemia have a high risk to develop HNSCC, we investigated whether and to which extent Fanconi anemia pathway inactivation underlies CIN in HNSCC of non-Fanconi anemia individuals. We observed ICL-induced chromosomal breakage in 9 of 17 (53%) HNSCC cell lines derived from patients without Fanconi anemia. In addition, defective sister chromatid cohesion was observed in five HNSCC cell lines. Inactivation of FANCM was responsible for chromosomal breakage in one cell line, whereas in two other cell lines, somatic mutations in PDS5A or STAG2 resulted in inadequate sister chromatid cohesion. In addition, FANCF methylation was found in one cell line by screening an additional panel of 39 HNSCC cell lines. Our data demonstrate that CIN in terms of ICL-induced chromosomal breakage and defective chromatid cohesion is frequently observed in HNSCC. Inactivation of known Fanconi anemia and chromatid cohesion genes does explain CIN in the minority of cases. These findings point to phenotypes that may be highly relevant in treatment response of HNSCC. ©2015 American Association for Cancer Research.

  16. Frequency of sister chromatid exchange and chromosomal aberrations in asbestos cement workers.

    PubMed

    Fatma, N; Jain, A K; Rahman, Q

    1991-02-01

    Exposure to asbestos minerals has been associated with a wide variety of adverse health effects including lung cancer, pleural mesothelioma, and cancer of other organs. It was shown previously that asbestos samples collected from a local asbestos factory enhanced sister chromatid exchanges (SCEs) and chromosomal aberrations in vitro using human lymphocytes. In the present study, 22 workers from the same factory and 12 controls were further investigated. Controls were matched for age, sex, and socioeconomic state. The peripheral blood lymphocytes were cultured and harvested at 48 hours for studies of chromosomal aberrations and at 72 hours for SCE frequency determinations. Asbestos workers had a raised mean SCE rate and increased numbers of chromosomal aberrations compared with a control population. Most of the chromosomal aberrations were chromatid gap and break types.

  17. Histone hyperacetylation during meiosis interferes with large-scale chromatin remodeling, axial chromatid condensation and sister chromatid separation in the mammalian oocyte.

    PubMed

    Yang, Feikun; Baumann, Claudia; Viveiros, Maria M; De La Fuente, Rabindranath

    2012-01-01

    Histone acetylation regulates higher-order chromatin structure and function and is critical for the control of gene expression. Histone deacetylase inhibitors (HDACi) are currently under investigation as novel cancer therapeutic drugs. Here, we show that female germ cells are extremely susceptible to chromatin changes induced by HDACi. Our results indicate that exposure to trichostatin A (TSA) at nanomolar levels interferes with major chromatin remodeling events in the mammalian oocyte leading to chromosome instability. High resolution analysis of chromatin structure and live-cell imaging revealed a striking euchromatin decondensation associated with histone H4 hyperacetylation following exposure to 15 nM TSA in >90% of pre-ovulatory oocytes. Dynamic changes in large-scale chromatin structure were detected after 2 h of exposure and result in the formation of misaligned chromosomes in >75% (P<0.05) of in vitro matured oocytes showing chromosome lagging as well as abnormal sister chromatid separation at anaphase I. Abnormal axial chromatid condensation during meiosis results in the formation of elongated chromosomes exhibiting hyperacetylation of histone H4 at lysine 5 and lysine 16 at interstitial chromosome segments, but not pericentric heterochromatin, while highly decondensed bivalents exhibit prominent histone H3 phosphorylation at centromeric domains. Notably, no changes were observed in the chromosomal localization of the condensin protein SMC4. These results indicate that HDAC activity is required for proper chromosome condensation in the mammalian oocyte and that HDACi may induce abnormal chromosome segregation by interfering with both chromosome-microtubule interactions, as well as sister chromatid separation. Thus, HDACi, proposed for cancer therapy, may disrupt the epigenetic status of female germ cells, predisposing oocytes to aneuploidy at previously unrecognized low doses.

  18. Colchicine promotes a change in chromosome structure without loss of sister chromatid cohesion in prometaphase I-arrested bivalents.

    PubMed

    Rodríguez, E M; Parra, M T; Rufas, J S; Suja, J A

    2001-12-01

    In somatic cells colchicine promotes the arrest of cell division at prometaphase, and chromosomes show a sequential loss of sister chromatid arm and centromere cohesion. In this study we used colchicine to analyse possible changes in chromosome structure and sister chromatid cohesion in prometaphase I-arrested bivalents of the katydid Pycnogaster cucullata. After silver staining we observed that in colchicine-arrested prometaphase I bivalents, and in contrast to what was found in control bivalents, sister kinetochores appeared individualised and sister chromatid axes were completely separated all along their length. However, this change in chromosome structure occurred without loss of sister chromatid arm cohesion. We also employed the MPM-2 monoclonal antibody against mitotic phosphoproteins on control and colchicine-treated spermatocytes. In control metaphase I bivalents this antibody labelled the tightly associated sister kinetochores and the interchromatid domain. By contrast, in colchicine-treated prometaphase I bivalents individualised sister kinetochores appeared labelled, but the interchromatid domain did not show labelling. These results support the notion that MPM-2 phosphoproteins, probably DNA topoisomerase IIalpha, located in the interchromatid domain act as "chromosomal staples" associating sister chromatid axes in metaphase I bivalents. The disappearance of these chromosomal staples would induce a change in chromosome structure, as reflected by the separation of sister kinetochores and sister axes, but without a concomitant loss of sister chromatid cohesion.

  19. Cyclophosphamide induced in vivo sister chromatid exchanges (SCE) in Mus musculus. I: Strain differences and empirical association with relative chromosome size.

    PubMed

    Reimer, D L; Singh, S M

    1982-01-01

    The inducibility of sister chromatid exchanges (SCEs) by cyclophosphamide (CP) in bone marrow cells was evaluated in vivo in the three genetic strains of mice (C3H/s, C57BL/6J, and Balb/c). Female mice (10 to 12 wks old, mean = 22.9g, SD = 3.2g) were administered with nine hourly injections of 214.19 mg/kg 5-Bromo-2' deoxyuridine (BrdU) followed by 0, 0.048, 0.449, 4.585 or 46.93 mg/kg CP and 4 mg/kg colcemid. SCEs were evaluated following differential staining procedures of Perry and Wolff (1974). The base-line SCEs were similar in all strains with about ten SCEs/cell. Increasing CP concentrations yielded an increased level of SCEs. Most cells showed extensive damage in CP doses exceeding 4.55 mg/kg. No SCE evaluation was possible beyond this concentration. Strain differences were evident at every dose of CP, and Balb/c was the least susceptible strain to SCE induction. F1 hybrids involving C3H/s female and Balb/c male showed SCE values closer to Balb/c. Data on the association between chromosome length and frequency of SCEs are provided. They empirically establish a positive correlation (r = 0.90) between the two features. Most induced SCEs were interstitially located rather than terminally positioned on the chromosome.

  20. Cell killing and chromatid damage in primary human bronchial epithelial cells irradiated with accelerated 56Fe ions

    NASA Technical Reports Server (NTRS)

    Suzuki, M.; Piao, C.; Hall, E. J.; Hei, T. K.

    2001-01-01

    We examined cell killing and chromatid damage in primary human bronchial epithelial cells irradiated with high-energy 56Fe ions. Cells were irradiated with graded doses of 56Fe ions (1 GeV/nucleon) accelerated with the Alternating Gradient Synchrotron at Brookhaven National Laboratory. The survival curves for cells plated 1 h after irradiation (immediate plating) showed little or no shoulder. However, the survival curves for cells plated 24 h after irradiation (delayed plating) had a small initial shoulder. The RBE for 56Fe ions compared to 137Cs gamma rays was 1.99 for immediate plating and 2.73 for delayed plating at the D10. The repair ratio (delayed plating/immediate plating) was 1.67 for 137Cs gamma rays and 1.22 for 56Fe ions. The dose-response curves for initially measured and residual chromatid fragments detected by the Calyculin A-mediated premature chromosome condensation technique showed a linear response. The results indicated that the induction frequency for initially measured fragments was the same for 137Cs gamma rays and 56Fe ions. On the other hand, approximately 85% of the fragments induced by 137Cs gamma rays had rejoined after 24 h of postirradiation incubation; the corresponding amount for 56Fe ions was 37%. Furthermore, the frequency of chromatid exchanges induced by gamma rays measured 24 h after irradiation was higher than that induced by 56Fe ions. No difference in the amount of chromatid damage induced by the two types of radiations was detected when assayed 1 h after irradiation. The results suggest that high-energy 56Fe ions induce a higher frequency of complex, unrepairable damage at both the cellular and chromosomal levels than 137Cs gamma rays in the target cells for radiation-induced lung cancers.

  1. No significant level of inheritable interchromosomal aberrations in the progeny of bystander primary human fibroblasts after alpha particle irradiation

    NASA Astrophysics Data System (ADS)

    Hu, Burong; Zhu, Jiayun; Zhou, Hongning; Hei, Tom K.

    2013-02-01

    A major concern for bystander effects is the probability that normal healthy cells adjacent to the irradiated cells become genomically unstable and undergo further carcinogenesis after therapeutic irradiation or space mission where astronauts are exposed to low dose of heavy ions. Genomic instability is a hallmark of cancer cells. In the present study, two irradiation protocols were performed in order to ensure pure populations of bystander cells and the genomic instability in their progeny were investigated. After irradiation, chromosomal aberrations of cells were analyzed at designated time points using G2 phase premature chromosome condensation (G2-PCC) coupled with Giemsa staining and with multiplex fluorescent in situ hybridization (mFISH). Our Giemsa staining assay demonstrated that elevated yields of chromatid breaks were induced in the progeny of pure bystander primary fibroblasts up to 20 days after irradiation. mFISH assay showed no significant level of inheritable interchromosomal aberrations were induced in the progeny of the bystander cell groups, while the fractions of gross aberrations (chromatid breaks or chromosomal breaks) significantly increased in some bystander cell groups. These results suggest that genomic instability occurred in the progeny of the irradiation associated bystander normal fibroblasts exclude the inheritable interchromosomal aberration.

  2. No significant level of inheritable interchromosomal aberrations in the progeny of bystander primary human fibroblast after alpha particle irradiation.

    PubMed

    Hu, Burong; Zhu, Jiayun; Zhou, Hongning; Hei, Tom K

    2013-02-01

    A major concern for bystander effects is the probability that normal healthy cells adjacent to the irradiated cells become genomically unstable and undergo further carcinogenesis after therapeutic irradiation or space mission where astronauts are exposed to low dose of heavy ions. Genomic instability is a hallmark of cancer cells. In the present study, two irradiation protocols were performed in order to ensure pure populations of bystander cells and the genomic instability in their progeny were investigated. After irradiation, chromosomal aberrations of cells were analyzed at designated time points using G 2 phase premature chromosome condensation (G 2 -PCC) coupled with Giemsa staining and with multiplex fluorescent in situ hybridization (mFISH). Our Giemsa staining assay demonstrated that elevated yields of chromatid breaks were induced in the progeny of pure bystander primary fibroblasts up to 20 days after irradiation. MFISH assay showed no significant level of inheritable interchromosomal aberrations were induced in the progeny of the bystander cell groups, while the fractions of gross aberrations (chromatid breaks or chromosomal breaks) significantly increased in some bystander cell groups. These results suggest that genomic instability occurred in the progeny of the irradiation associated bystander normal fibroblasts exclude the inheritable interchromosomal aberration.

  3. Low-Energy Electron-Induced Strand Breaks in Telomere-Derived DNA Sequences-Influence of DNA Sequence and Topology.

    PubMed

    Rackwitz, Jenny; Bald, Ilko

    2018-03-26

    During cancer radiation therapy high-energy radiation is used to reduce tumour tissue. The irradiation produces a shower of secondary low-energy (<20 eV) electrons, which are able to damage DNA very efficiently by dissociative electron attachment. Recently, it was suggested that low-energy electron-induced DNA strand breaks strongly depend on the specific DNA sequence with a high sensitivity of G-rich sequences. Here, we use DNA origami platforms to expose G-rich telomere sequences to low-energy (8.8 eV) electrons to determine absolute cross sections for strand breakage and to study the influence of sequence modifications and topology of telomeric DNA on the strand breakage. We find that the telomeric DNA 5'-(TTA GGG) 2 is more sensitive to low-energy electrons than an intermixed sequence 5'-(TGT GTG A) 2 confirming the unique electronic properties resulting from G-stacking. With increasing length of the oligonucleotide (i.e., going from 5'-(GGG ATT) 2 to 5'-(GGG ATT) 4 ), both the variety of topology and the electron-induced strand break cross sections increase. Addition of K + ions decreases the strand break cross section for all sequences that are able to fold G-quadruplexes or G-intermediates, whereas the strand break cross section for the intermixed sequence remains unchanged. These results indicate that telomeric DNA is rather sensitive towards low-energy electron-induced strand breakage suggesting significant telomere shortening that can also occur during cancer radiation therapy. © 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

  4. A Benzothiazole Derivative (5g) Induces DNA Damage And Potent G2/M Arrest In Cancer Cells.

    PubMed

    Hegde, Mahesh; Vartak, Supriya V; Kavitha, Chandagirikoppal V; Ananda, Hanumappa; Prasanna, Doddakunche S; Gopalakrishnan, Vidya; Choudhary, Bibha; Rangappa, Kanchugarakoppal S; Raghavan, Sathees C

    2017-05-31

    Chemically synthesized small molecules play important role in anticancer therapy. Several chemical compounds have been reported to damage the DNA, either directly or indirectly slowing down the cancer cell progression by causing a cell cycle arrest. Direct or indirect reactive oxygen species formation causes DNA damage leading to cell cycle arrest and subsequent cell death. Therefore, identification of chemically synthesized compounds with anticancer potential is important. Here we investigate the effect of benzothiazole derivative (5g) for its ability to inhibit cell proliferation in different cancer models. Interestingly, 5g interfered with cell proliferation in both, cell lines and tumor cells leading to significant G2/M arrest. 5g treatment resulted in elevated levels of ROS and subsequently, DNA double-strand breaks (DSBs) explaining observed G2/M arrest. Consistently, we observed deregulation of many cell cycle associated proteins such as CDK1, BCL2 and their phosphorylated form, CyclinB1, CDC25c etc. Besides, 5g treatment led to decreased levels of mitochondrial membrane potential and activation of apoptosis. Interestingly, 5g administration inhibited tumor growth in mice without significant side effects. Thus, our study identifies 5g as a potent biochemical inhibitor to induce G2/M phase arrest of the cell cycle, and demonstrates its anticancer properties both ex vivo and in vivo.

  5. Visualizing Vpr-Induced G2 Arrest and Apoptosis

    PubMed Central

    Murakami, Tomoyuki; Aida, Yoko

    2014-01-01

    Vpr is an accessory protein of human immunodeficiency virus type 1 (HIV-1) with multiple functions. The induction of G2 arrest by Vpr plays a particularly important role in efficient viral replication because the transcriptional activity of the HIV-1 long terminal repeat is most active in G2 phase. The regulation of apoptosis by Vpr is also important for immune suppression and pathogenesis during HIV infection. However, it is not known whether Vpr-induced apoptosis depends on the ability of Vpr to induce G2 arrest, and the dynamics of Vpr-induced G2 arrest and apoptosis have not been visualized. We performed time-lapse imaging to examine the temporal relationship between Vpr-induced G2 arrest and apoptosis using HeLa cells containing the fluorescent ubiquitination-based cell cycle indicator2 (Fucci2). The dynamics of G2 arrest and subsequent long-term mitotic cell rounding in cells transfected with the Vpr-expression vector were visualized. These cells underwent nuclear mis-segregation after prolonged mitotic processes and then entered G1 phase. Some cells subsequently displayed evidence of apoptosis after prolonged mitotic processes and nuclear mis-segregation. Interestingly, Vpr-induced apoptosis was seldom observed in S or G2 phase. Likewise, visualization of synchronized HeLa/Fucci2 cells infected with an adenoviral vector expressing Vpr clearly showed that Vpr arrests the cell cycle at G2 phase, but does not induce apoptosis at S or G2 phase. Furthermore, time-lapse imaging of HeLa/Fucci2 cells expressing SCAT3.1, a caspase-3-sensitive fusion protein, clearly demonstrated that Vpr induces caspase-3-dependent apoptosis. Finally, to examine whether the effects of Vpr on G2 arrest and apoptosis were reversible, we performed live-cell imaging of a destabilizing domain fusion Vpr, which enabled rapid stabilization and destabilization by Shield1. The effects of Vpr on G2 arrest and subsequent apoptosis were reversible. This study is the first to characterize the

  6. Chromosomal radiosensitivity in head and neck cancer patients: evidence for genetic predisposition?

    PubMed Central

    De Ruyck, K; de Gelder, V; Van Eijkeren, M; Boterberg, T; De Neve, W; Vral, A; Thierens, H

    2008-01-01

    The association between chromosomal radiosensitivity and genetic predisposition to head and neck cancer was investigated in this study. In all, 101 head and neck cancer patients and 75 healthy control individuals were included in the study. The G2 assay was used to measure chromosomal radiosensitivity. The results demonstrated that head and neck cancer patients had a statistically higher number of radiation-induced chromatid breaks than controls, with mean values of 1.23 and 1.10 breaks per cell, respectively (P<0.001). Using the 90th percentile of the G2 scores of the healthy individuals as a cutoff value for chromosomal radiosensitivity, 26% of the cancer patients were radiosensitive compared with 9% of the healthy controls (P=0.008). The mean number of radiation-induced chromatid breaks and the proportion of radiosensitive individuals were highest for oral cavity cancer patients (1.26 breaks per cell, 38%) and pharynx cancer patients (1.27 breaks per cell, 35%). The difference between patients and controls was most pronounced in the lower age group (⩽50 years, 1.32 breaks per cell, 38%) and in the non- and light smoking patient group (⩽10 pack-years, 1.28 breaks per cell, 46%). In conclusion, enhanced chromosomal radiosensitivity is a marker of genetic predisposition to head and neck cancer, and the genetic contribution is highest for oral cavity and pharynx cancer patients and for early onset and non- and light smoking patients. PMID:18414410

  7. Chromosomal radiosensitivity in head and neck cancer patients: evidence for genetic predisposition?

    PubMed

    De Ruyck, K; de Gelder, V; Van Eijkeren, M; Boterberg, T; De Neve, W; Vral, A; Thierens, H

    2008-05-20

    The association between chromosomal radiosensitivity and genetic predisposition to head and neck cancer was investigated in this study. In all, 101 head and neck cancer patients and 75 healthy control individuals were included in the study. The G(2) assay was used to measure chromosomal radiosensitivity. The results demonstrated that head and neck cancer patients had a statistically higher number of radiation-induced chromatid breaks than controls, with mean values of 1.23 and 1.10 breaks per cell, respectively (P<0.001). Using the 90th percentile of the G(2) scores of the healthy individuals as a cutoff value for chromosomal radiosensitivity, 26% of the cancer patients were radiosensitive compared with 9% of the healthy controls (P=0.008). The mean number of radiation-induced chromatid breaks and the proportion of radiosensitive individuals were highest for oral cavity cancer patients (1.26 breaks per cell, 38%) and pharynx cancer patients (1.27 breaks per cell, 35%). The difference between patients and controls was most pronounced in the lower age group (breaks per cell, 38%) and in the non- and light smoking patient group (breaks per cell, 46%). In conclusion, enhanced chromosomal radiosensitivity is a marker of genetic predisposition to head and neck cancer, and the genetic contribution is highest for oral cavity and pharynx cancer patients and for early onset and non- and light smoking patients.

  8. Inactivation of the budding yeast cohesin loader Scc2 alters gene expression both globally and in response to a single DNA double strand break

    PubMed Central

    Lindgren, Emma; Hägg, Sara; Giordano, Fosco; Björkegren, Johan; Ström, Lena

    2014-01-01

    Genome integrity is fundamental for cell survival and cell cycle progression. Important mechanisms for keeping the genome intact are proper sister chromatid segregation, correct gene regulation and efficient repair of damaged DNA. Cohesin and its DNA loader, the Scc2/4 complex have been implicated in all these cellular actions. The gene regulation role has been described in several organisms. In yeast it has been suggested that the proteins in the cohesin network would effect transcription based on its role as insulator. More recently, data are emerging indicating direct roles for gene regulation also in yeast. Here we extend these studies by investigating whether the cohesin loader Scc2 is involved in regulation of gene expression. We performed global gene expression profiling in the absence and presence of DNA damage, in wild type and Scc2 deficient G2/M arrested cells, when it is known that Scc2 is important for DNA double strand break repair and formation of damage induced cohesion. We found that not only the DNA damage specific transcriptional response is distorted after inactivation of Scc2 but also the overall transcription profile. Interestingly, these alterations did not correlate with changes in cohesin binding. PMID:25483075

  9. Topoisomerase IIα maintains genomic stability through decatenation G2 checkpoint signaling

    PubMed Central

    Bower, Jacquelyn J.; Karaca, Gamze F.; Zhou, Yingchun; Simpson, Dennis A.; Cordeiro-Stone, Marila; Kaufmann, William K.

    2010-01-01

    Topoisomerase IIα (topoIIα) is an essential mammalian enzyme that topologically modifies DNA and is required for chromosome segregation during mitosis. Previous research suggests that inhibition of topoII decatenatory activity triggers a G2 checkpoint response, which delays mitotic entry due to insufficient decatenation of daughter chromatids. Here we examine the effects of both topoIIα and topoIIβ on decatenatory activity in cell extracts, DNA damage and decatenation G2 checkpoint function, and the frequencies of p16INK4A allele loss and gain. In diploid human fibroblast lines, depletion of topoIIα by siRNA was associated with severely reduced decatenatory activity, delayed progression from G2 into mitosis, and insensitivity to G2 arrest induced by the topoII catalytic inhibitor ICRF-193. Furthermore, interphase nuclei of topoIIα-depleted cells displayed increased frequencies of losses and gains of the tumor suppressor genetic locus p16INK4A. This study demonstrates that the topoIIα protein is required for decatenation G2 checkpoint function, and inactivation of decatenation and the decatenation G2 checkpoint leads to abnormal chromosome segregation and genomic instability. PMID:20562910

  10. Cleavage of cohesin rings coordinates the separation of centrioles and chromatids.

    PubMed

    Schöckel, Laura; Möckel, Martin; Mayer, Bernd; Boos, Dominik; Stemmann, Olaf

    2011-07-10

    Cohesin pairs sister chromatids by forming a tripartite Scc1-Smc1-Smc3 ring around them. In mitosis, cohesin is removed from chromosome arms by the phosphorylation-dependent prophase pathway. Centromeric cohesin is protected by shugoshin 1 and protein phosphatase 2A (Sgo1-PP2A) and opened only in anaphase by separase-dependent cleavage of Scc1 (refs 4-6). Following chromosome segregation, centrioles loosen their tight orthogonal arrangement, which licenses later centrosome duplication in S phase. Although a role of separase in centriole disengagement has been reported, the molecular details of this process remain enigmatic. Here, we identify cohesin as a centriole-engagement factor. Both premature sister-chromatid separation and centriole disengagement are induced by ectopic activation of separase or depletion of Sgo1. These unscheduled events are suppressed by expression of non-cleavable Scc1 or inhibition of the prophase pathway. When endogenous Scc1 is replaced by artificially cleavable Scc1, the corresponding site-specific protease triggers centriole disengagement. Separation of centrioles can alternatively be induced by ectopic cleavage of an engineered Smc3. Thus, the chromosome and centrosome cycles exhibit extensive parallels and are coordinated with each other by dual use of the cohesin ring complex.

  11. Titanium Dioxide Nanoparticles are not Cytotoxic or Clastogenic in Human Skin Cells

    PubMed Central

    Browning, Cynthia L; The, Therry; Mason, Michael D; Wise, John Pierce

    2015-01-01

    The application of nanoparticle technology is rapidly expanding. The reduced dimensionality of nanoparticles can give rise to changes in chemical and physical properties, often resulting in altered toxicity. People are exposed dermally to titanium dioxide (TiO2) nanoparticles in industrial and residential settings. The general public is increasingly exposed to these nanoparticles as their use in cosmetics, sunscreens and lotions expands. The toxicity of TiO2 nanoparticles towards human skin cells is unclear and understudied. We used a human skin fibroblast cell line to investigate the cytotoxicity and clastogenicity of TiO2 nanoparticles after 24 h exposure. In a clonogenic survival assay, treatments of 10, 50 and 100 μg/cm2 induced 97.8, 88.8 and 84.7% relative survival, respectively. Clastogenicity was assessed using a chromosomal aberration assay in order to determine whether TiO2 nanoparticles induced serious forms of DNA damage such as chromatid breaks, isochromatid lesions or chromatid exchanges. Treatments of 0, 10, 50 and 100 μg/cm2 induced 3.3, 3.0, 3.0 and 2.7% metaphases with damage, respectively. No isochromatid lesions or chromatid exchanges were detected. These data show that TiO2 nanoparticles are not cytotoxic or clastogenic to human skin cells. PMID:26568896

  12. Titanium Dioxide Nanoparticles are not Cytotoxic or Clastogenic in Human Skin Cells.

    PubMed

    Browning, Cynthia L; The, Therry; Mason, Michael D; Wise, John Pierce

    2014-11-01

    The application of nanoparticle technology is rapidly expanding. The reduced dimensionality of nanoparticles can give rise to changes in chemical and physical properties, often resulting in altered toxicity. People are exposed dermally to titanium dioxide (TiO 2 ) nanoparticles in industrial and residential settings. The general public is increasingly exposed to these nanoparticles as their use in cosmetics, sunscreens and lotions expands. The toxicity of TiO 2 nanoparticles towards human skin cells is unclear and understudied. We used a human skin fibroblast cell line to investigate the cytotoxicity and clastogenicity of TiO 2 nanoparticles after 24 h exposure. In a clonogenic survival assay, treatments of 10, 50 and 100 μg/cm 2 induced 97.8, 88.8 and 84.7% relative survival, respectively. Clastogenicity was assessed using a chromosomal aberration assay in order to determine whether TiO 2 nanoparticles induced serious forms of DNA damage such as chromatid breaks, isochromatid lesions or chromatid exchanges. Treatments of 0, 10, 50 and 100 μg/cm 2 induced 3.3, 3.0, 3.0 and 2.7% metaphases with damage, respectively. No isochromatid lesions or chromatid exchanges were detected. These data show that TiO 2 nanoparticles are not cytotoxic or clastogenic to human skin cells.

  13. 750 GeV diphoton resonance, 125 GeV Higgs and muon g - 2 anomaly in deflected anomaly mediation SUSY breaking scenarios

    NASA Astrophysics Data System (ADS)

    Wang, Fei; Wu, Lei; Yang, Jin Min; Zhang, Mengchao

    2016-08-01

    We propose to interpret the 750 GeV diphoton excess in deflected anomaly mediation supersymmetry breaking scenarios, which can naturally predict couplings between a singlet field and vector-like messengers. The CP-even scalar component (S) of the singlet field can serve as the 750 GeV resonance. The messenger scale, which is of order the gravitino scale, can be as light as Fϕ ∼ O (10) TeV when the messenger species NF and the deflection parameter d are moderately large. Such messengers can induce the large loop decay process S → γγ. Our results show that such a scenario can successfully accommodate the 125 GeV Higgs boson, the 750 GeV diphoton excess and the muon g - 2 without conflicting with the LHC constraints. We also comment on the possible explanations in the gauge mediation supersymmetry breaking scenario.

  14. How-to-Do-It: Demonstrating Sister Chromatid Exchanges.

    ERIC Educational Resources Information Center

    Dye, Frank J.

    1988-01-01

    Outlines procedures for demonstrating and preparing a permanent slide of sister chromatid exchanges and recombination events between the two chromatids of a single chromosome. Provides the name of an additional resource for making preparations of exchanges. (RT)

  15. Mechanics of Sister Chromatids studied with a Polymer Model English</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhang, Yang; Isbaner, Sebastian; Heermann, Dieter</p> <p>2013-10-01</p> <p>Sister <span class="hlt">chromatid</span> cohesion denotes the phenomenon that sister <span class="hlt">chromatids</span> are initially attached to each other in mitosis to guarantee the error-free distribution into the daughter cells. Cohesion is mediated by binding proteins and only resolved after mitotic chromosome condensation is completed. However, the amount of attachement points required to maintain sister <span class="hlt">chromatid</span> cohesion while still allowing proper chromosome condensation is not known yet. Additionally the impact of cohesion on the mechanical properties of chromosomes also poses an interesting problem. In this work we study the conformational and mechanical properties of sister <span class="hlt">chromatids</span> by means of computer simulations. We model both protein-mediated cohesion between sister <span class="hlt">chromatids</span> and chromosome condensation with a dynamic binding mechanisms. We show in a phase diagram that only specific link concentrations lead to connected and fully condensed <span class="hlt">chromatids</span> that do not intermingle with each other nor separate due to entropic forces. Furthermore we show that dynamic bonding between <span class="hlt">chromatids</span> decrease the Young's modulus compared to non-bonded <span class="hlt">chromatids</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5384111','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5384111"><span><span class="hlt">Break-induced</span> telomere synthesis underlies alternative telomere maintenance</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Dilley, Robert L.; Verma, Priyanka; Cho, Nam Woo; Winters, Harrison D.; Wondisford, Anne R.; Greenberg, Roger A.</p> <p>2017-01-01</p> <p>Homology-directed DNA repair is essential for genome maintenance through templated DNA synthesis. Alternative lengthening of telomeres (ALT) necessitates homology-directed DNA repair to maintain telomeres in about 10–15% of human cancers. How DNA damage <span class="hlt">induces</span> assembly and execution of a DNA replication complex (<span class="hlt">break-induced</span> replisome) at telomeres or elsewhere in the mammalian genome is poorly understood. Here we define <span class="hlt">break-induced</span> telomere synthesis and demonstrate that it utilizes a specialized replisome, which underlies ALT telomere maintenance. DNA double-strand <span class="hlt">breaks</span> enact nascent telomere synthesis by long-tract unidirectional replication. Proliferating cell nuclear antigen (PCNA) loading by replication factor C (RFC) acts as the initial sensor of telomere damage to establish predominance of DNA polymerase δ (Pol δ) through its POLD3 subunit. <span class="hlt">Break-induced</span> telomere synthesis requires the RFC–PCNA–Pol δ axis, but is independent of other canonical replisome components, ATM and ATR, or the homologous recombination protein Rad51. Thus, the inception of telomere damage recognition by the <span class="hlt">break-induced</span> replisome orchestrates homology-directed telomere maintenance. PMID:27760120</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3674063','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3674063"><span>Sister <span class="hlt">chromatid</span> segregation in meiosis II</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Wassmann, Katja</p> <p>2013-01-01</p> <p>Meiotic divisions (meiosis I and II) are specialized cell divisions to generate haploid gametes. The first meiotic division with the separation of chromosomes is named reductional division. The second division, which takes place immediately after meiosis I without intervening S-phase, is equational, with the separation of sister <span class="hlt">chromatids</span>, similar to mitosis. This meiotic segregation pattern requires the two-step removal of the cohesin complex holding sister <span class="hlt">chromatids</span> together: cohesin is removed from chromosome arms that have been subjected to homologous recombination in meiosis I and from the centromere region in meiosis II. Cohesin in the centromere region is protected from removal in meiosis I, but this protection has to be removed—deprotected”—for sister <span class="hlt">chromatid</span> segregation in meiosis II. Whereas the mechanisms of cohesin protection are quite well understood, the mechanisms of deprotection have been largely unknown until recently. In this review I summarize our current knowledge on cohesin deprotection. PMID:23574717</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22285267-zinc-protects-hepg2-cells-against-oxidative-damage-dna-damage-induced-ochratoxin','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22285267-zinc-protects-hepg2-cells-against-oxidative-damage-dna-damage-induced-ochratoxin"><span>Zinc protects Hep<span class="hlt">G</span><span class="hlt">2</span> cells against the oxidative damage and DNA damage <span class="hlt">induced</span> by ochratoxin A</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Zheng, Juanjuan; Zhang, Yu; Xu, Wentao, E-mail: xuwentaoboy@sina.com</p> <p></p> <p>Oxidative stress and DNA damage are the most studied mechanisms by which ochratoxin A (OTA) <span class="hlt">induces</span> its toxic effects, which include nephrotoxicity, hepatotoxicity, immunotoxicity and genotoxicity. Zinc, which is an essential trace element, is considered a potential antioxidant. The aim of this paper was to investigate whether zinc supplement could inhibit OTA-<span class="hlt">induced</span> oxidative damage and DNA damage in Hep<span class="hlt">G</span><span class="hlt">2</span> cells and the mechanism of inhibition. The results indicated that that exposure of OTA decreased the intracellular zinc concentration; zinc supplement significantly reduced the OTA-<span class="hlt">induced</span> production of reactive oxygen species (ROS) and decrease in superoxide dismutase (SOD) activity but did notmore » affect the OTA-<span class="hlt">induced</span> decrease in the mitochondrial membrane potential (Δψ{sub m}). Meanwhile, the addition of the zinc chelator N,N,N′,N′-tetrakis(<span class="hlt">2</span>-pyridylmethyl)ethylenediamine (TPEN) strongly aggravated the OTA-<span class="hlt">induced</span> oxidative damage. This study also demonstrated that zinc helped to maintain the integrity of DNA through the reduction of OTA-<span class="hlt">induced</span> DNA strand <span class="hlt">breaks</span>, 8-hydroxy-<span class="hlt">2</span>′-deoxyguanosine (8-OHd<span class="hlt">G</span>) formation and DNA hypomethylation. OTA increased the mRNA expression of metallothionein1-A (MT1A), metallothionein<span class="hlt">2</span>-A (MT<span class="hlt">2</span>A) and Cu/Zn superoxide dismutase (SOD1). Zinc supplement further enhanced the mRNA expression of MT1A and MT<span class="hlt">2</span>A, but it had no effect on the mRNA expression of SOD1 and catalase (CAT). Zinc was for the first time proven to reduce the cytotoxicity of OTA through inhibiting the oxidative damage and DNA damage, and regulating the expression of zinc-associated genes. Thus, the addition of zinc can potentially be used to reduce the OTA toxicity of contaminated feeds. - Highlights: ► OTA decreased the intracellular zinc concentration. ► OTA <span class="hlt">induced</span> the formation of 8-OHd<span class="hlt">G</span> in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. ► It was testified for the first time that OTA <span class="hlt">induced</span> DNA hypomethylation. ► Zinc protects against the oxidative damage and DNA damage</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24709322','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24709322"><span>Organic extracts of coke oven emissions can <span class="hlt">induce</span> genetic damage in metabolically competent Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Xin, Lili; Wang, Jianshu; Guo, Sifan; Wu, Yanhu; Li, Xiaohai; Deng, Huaxin; Kuang, Dan; Xiao, Wei; Wu, Tangchun; Guo, Huan</p> <p>2014-05-01</p> <p>Coke oven emissions (COEs) containing various carcinogenic polycyclic aromatic hydrocarbons (PAHs) represent the coal-burning pollution in the air. Organic pollutants in the aerosol and particulate matter of COEs were collected from the bottom, side, and top of a coke oven. The Comet assay and cytokinesis-block micronucleus cytome assay were conducted to analyze the genetic damage of extractable organic matter (EOM) of COEs on Hep<span class="hlt">G</span><span class="hlt">2</span> cells. All the three EOMs could <span class="hlt">induce</span> significant dose-dependent increases in Olive tail moment, tail DNA, and tail length, micronuclei, nucleoplasmic bridges, and nuclear buds frequencies, which were mostly positively correlated with the total PAHs concentration in each EOM. In conclusion, EOMs of COEs in the three typical working places of coke oven can <span class="hlt">induce</span> DNA strand <span class="hlt">breaks</span> and genomic instability in the metabolically competent Hep<span class="hlt">G</span><span class="hlt">2</span> cells. The PAHs in EOMs may be important causative agents for the genotoxic effects of COEs. Copyright © 2014 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3488256','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3488256"><span>UVA-<span class="hlt">induced</span> DNA double-strand <span class="hlt">breaks</span> result from the repair of clustered oxidative DNA damages</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Greinert, R.; Volkmer, B.; Henning, S.; Breitbart, E. W.; Greulich, K. O.; Cardoso, M. C.; Rapp, Alexander</p> <p>2012-01-01</p> <p>UVA (320–400 nm) represents the main spectral component of solar UV radiation, <span class="hlt">induces</span> pre-mutagenic DNA lesions and is classified as Class I carcinogen. Recently, discussion arose whether UVA <span class="hlt">induces</span> DNA double-strand <span class="hlt">breaks</span> (dsbs). Only few reports link the induction of dsbs to UVA exposure and the underlying mechanisms are poorly understood. Using the Comet-assay and γH<span class="hlt">2</span>AX as markers for dsb formation, we demonstrate the dose-dependent dsb induction by UVA in <span class="hlt">G</span>1-synchronized human keratinocytes (HaCaT) and primary human skin fibroblasts. The number of γH<span class="hlt">2</span>AX foci increases when a UVA dose is applied in fractions (split dose), with a <span class="hlt">2</span>-h recovery period between fractions. The presence of the anti-oxidant Naringin reduces dsb formation significantly. Using an FPG-modified Comet-assay as well as warm and cold repair incubation, we show that dsbs arise partially during repair of bi-stranded, oxidative, clustered DNA lesions. We also demonstrate that on stretched chromatin fibres, 8-oxo-<span class="hlt">G</span> and abasic sites occur in clusters. This suggests a replication-independent formation of UVA-<span class="hlt">induced</span> dsbs through clustered single-strand <span class="hlt">breaks</span> via locally generated reactive oxygen species. Since UVA is the main component of solar UV exposure and is used for artificial UV exposure, our results shine new light on the aetiology of skin cancer. PMID:22941639</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_1");'>1</a></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li class="active"><span>3</span></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_3 --> <div id="page_4" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li class="active"><span>4</span></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="61"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11444040','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11444040"><span>Splitting the chromosome: cutting the ties that bind sister <span class="hlt">chromatids</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nasmyth, K; Peters, J M; Uhlmann, F</p> <p>2001-01-01</p> <p>In eukaryotic cells, replicated DNA molecules remain physically connected from their synthesis in S phase until they are separated during anaphase. This phenomenon, called sister <span class="hlt">chromatid</span> cohesion, is essential for the temporal separation of DNA replication and mitosis and for the equal separation of the duplicated genome. Recent work has identified a number of chromosomal proteins required for cohesion. In this review we discuss how these proteins may connect sister <span class="hlt">chromatids</span> and how they are removed from chromosomes to allow sister <span class="hlt">chromatid</span> separation at the onset of anaphase.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvD..96g5025W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvD..96g5025W"><span>Solving the muon <span class="hlt">g</span> -<span class="hlt">2</span> anomaly in deflected anomaly mediated SUSY <span class="hlt">breaking</span> with messenger-matter interactions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Fei; Wang, Wenyu; Yang, Jin Min</p> <p>2017-10-01</p> <p>We propose to introduce general messenger-matter interactions in the deflected anomaly mediated supersymmetry (SUSY) <span class="hlt">breaking</span> (AMSB) scenario to explain the <span class="hlt">g</span>μ-<span class="hlt">2</span> anomaly. Scenarios with complete or incomplete grand unified theory (GUT) multiplet messengers are discussed, respectively. The introduction of incomplete GUT mulitiplets can be advantageous in various aspects. We found that the <span class="hlt">g</span>μ-<span class="hlt">2</span> anomaly can be solved in both scenarios under current constraints including the gluino mass bounds, while the scenarios with incomplete GUT representation messengers are more favored by the <span class="hlt">g</span>μ-<span class="hlt">2</span> data. We also found that the gluino is upper bounded by about <span class="hlt">2</span>.5 TeV (<span class="hlt">2</span>.0 TeV) in scenario A and 3.0 TeV (<span class="hlt">2</span>.7 TeV) in scenario B if the generalized deflected AMSB scenarios are used to fully account for the <span class="hlt">g</span>μ-<span class="hlt">2</span> anomaly at 3 σ (<span class="hlt">2</span> σ ) level. Such a gluino should be accessible in the future LHC searches. Dark matter (DM) constraints, including DM relic density and direct detection bounds, favor scenario B with incomplete GUT multiplets. Much of the allowed parameter space for scenario B could be covered by the future DM direct detection experiments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4214429','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4214429"><span>Alternative meiotic <span class="hlt">chromatid</span> segregation in the holocentric plant Luzula elegans</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Heckmann, Stefan; Jankowska, Maja; Schubert, Veit; Kumke, Katrin; Ma, Wei; Houben, Andreas</p> <p>2014-01-01</p> <p>Holocentric chromosomes occur in a number of independent eukaryotic lineages. They form holokinetic kinetochores along the entire poleward <span class="hlt">chromatid</span> surfaces, and owing to this alternative chromosome structure, species with holocentric chromosomes cannot use the two-step loss of cohesion during meiosis typical for monocentric chromosomes. Here we show that the plant Luzula elegans maintains a holocentric chromosome architecture and behaviour throughout meiosis, and in contrast to monopolar sister centromere orientation, the unfused holokinetic sister centromeres behave as two distinct functional units during meiosis I, resulting in sister <span class="hlt">chromatid</span> separation. Homologous non-sister <span class="hlt">chromatids</span> remain terminally linked after metaphase I, by satellite DNA-enriched chromatin threads, until metaphase II. They then separate at anaphase II. Thus, an inverted sequence of meiotic sister <span class="hlt">chromatid</span> segregation occurs. This alternative meiotic process is most likely one possible adaptation to handle a holocentric chromosome architecture and behaviour during meiosis. PMID:25296379</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3546001','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3546001"><span>Middle Infrared Radiation <span class="hlt">Induces</span> <span class="hlt">G</span><span class="hlt">2</span>/M Cell Cycle Arrest in A549 Lung Cancer Cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Huang, Hsuan-Cheng; Tsai, Shang-Ru; Juan, Hsueh-Fen; Lee, Si-Chen</p> <p>2013-01-01</p> <p>There were studies investigating the effects of broadband infrared radiation (IR) on cancer cell, while the influences of middle-infrared radiation (MIR) are still unknown. In this study, a MIR emitter with emission wavelength band in the 3–5 µm region was developed to irradiate A549 lung adenocarcinoma cells. It was found that MIR exposure inhibited cell proliferation and <span class="hlt">induced</span> morphological changes by altering the cellular distribution of cytoskeletal components. Using quantitative PCR, we found that MIR promoted the expression levels of ATM (ataxia telangiectasia mutated), ATR (ataxia-telangiectasia and Rad3-related and Rad3-related), TP53 (tumor protein p53), p21 (CDKN1A, cyclin-dependent kinase inhibitor 1A) and GADD45 (growth arrest and DNA-damage <span class="hlt">inducible</span>), but decreased the expression levels of cyclin B coding genes, CCNB1 and CCNB<span class="hlt">2</span>, as well as CDK1 (Cyclin-dependent kinase 1). The reduction of protein expression levels of CDC25C, cyclin B1 and the phosphorylation of CDK1 at Thr-161 altogether suggest <span class="hlt">G</span><span class="hlt">2</span>/M arrest occurred in A549 cells by MIR. DNA repair foci formation of DNA double-strand <span class="hlt">breaks</span> (DSB) marker γ-H<span class="hlt">2</span>AX and sensor 53BP1 was <span class="hlt">induced</span> by MIR treatment, it implies the MIR <span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span>/M cell cycle arrest resulted from DSB. This study illustrates a potential role for the use of MIR in lung cancer therapy by initiating DSB and blocking cell cycle progression. PMID:23335992</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/8248278','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/8248278"><span>Enhancement of antineoplastic effect and attenuation of sister <span class="hlt">chromatid</span> exchanges by prostaglandin E<span class="hlt">2</span> in Ehrlich ascites tumour cells treated with cyclophosphamide in vivo.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mourelatos, D; Kritsi, Z; Mioglou, E; Dozi-Vassiliades, J</p> <p>1993-09-01</p> <p>Reduced sister <span class="hlt">chromatid</span> exchanges (SCE) frequency in response to cyclophosphamide (CP) was observed when Ehrlich ascites tumour (EAT) cells were exposed in vivo to <span class="hlt">2</span> micrograms/<span class="hlt">g</span> body weight of prostaglandin E<span class="hlt">2</span> (PGE<span class="hlt">2</span>). 1 h before i.p. injection of 5-bromodeoxyuridine (BrdUrd) adsorbed to activated charcoal, EAT-bearing mice treated i.p. with CP appeared to have increased SCE rates and cell division delays. PGE<span class="hlt">2</span> had no effect on survival and in inhibiting tumour growth. CP had only a slight non-significant effect on survival and in inhibiting tumour growth. In mice treated with the combined CP (5 micrograms/<span class="hlt">g</span> bd wt) plus PGE<span class="hlt">2</span> (<span class="hlt">2</span> micrograms/<span class="hlt">g</span> bd wt) a significant enhancement (P < 0.01) of survival time was accompanied by inhibition of tumour growth (P < 0.01) in comparison with the untreated controls. These data imply that SCEs might result from errors in a repair process which might involve a PGE<span class="hlt">2</span> sensitive step.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12132876','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12132876"><span>Comparative study of sister <span class="hlt">chromatid</span> exchange induction and antitumor effects by homo-aza-steroidal esters of [p-[bis(<span class="hlt">2</span>-chloroethyl)amino]phenyl]butyric acid.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Camoutsis, C; Catsoulacos, D; Karayiann, V; Papageorgiou, A; Mourelatos, D; Mioglou, E; Kritsi, Z; Nikolaropoulos, S</p> <p>2001-01-01</p> <p>The present work was undertaken in order to test the hypothesis that the Sister <span class="hlt">Chromatid</span> Exchange (SCE) assay in vitro can be used for the prediction of in vivo tumor response to newly synthesized potential chemotherapeutics. The effect of three homo-aza-steroidal esters containing the -CONH- in the steroidal nucleus, 1, <span class="hlt">2</span>, and 3 on SCE rates and on cell kinetics in cultured human lymphocytes was studied. The antitumor activity of these compounds was tested on leukemia P388- and leukemia L1210-bearing mice. The three substances <span class="hlt">induced</span> statistically significant enhancement of SCEs and of cell division delays. Compounds 1 and 3 were identified, on a molar basis, as more effective <span class="hlt">inducers</span> of SCEs and of cell division delays compared with compound <span class="hlt">2</span>. Compounds 1 and 3 had upon both experimental tumors better therapeutic effects compared with compound <span class="hlt">2</span> at equitoxic doses. Therefore, the order of the antitumor effectiveness of the three compounds coincided with the order of the cytogenetic effects they <span class="hlt">induced</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvB..96n4113A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvB..96n4113A"><span>Shear-<span class="hlt">induced</span> mechanical failure of β -<span class="hlt">G</span> a<span class="hlt">2</span>O3 from quantum mechanics simulations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>An, Qi; Li, Guodong</p> <p>2017-10-01</p> <p>Monoclinic gallium oxide (β -<span class="hlt">G</span> a<span class="hlt">2</span>O3 ) has important applications in power devices and deep UV optoelectronic devices because of such novel properties as a wide band gap, high breakdown electric field, and a wide range of n -type doping conductivity. However, the intrinsic failure mechanisms of β -<span class="hlt">G</span> a<span class="hlt">2</span>O3 remain unknown, which limits the fabrication and packaging of β -<span class="hlt">G</span> a<span class="hlt">2</span>O3 -based electronic devices. Here we used density-functional theory at the Perdew-Burke-Ernzerhof level to examine the shear-<span class="hlt">induced</span> failure mechanisms of β -<span class="hlt">G</span> a<span class="hlt">2</span>O3 along various plausible slip systems. We found that the (001 )/〈010 〉 slip system has the lowest ideal shear strength of 3.8 GPa among five plausible slip systems, suggesting that (001 )/〈010 〉 is the most plausible activated slip system. This slip leads to an intrinsic failure mechanism arising from <span class="hlt">breaking</span> the longest Ga-O bond between octahedral Ga and fourfold-coordinated O. Then we identified the same failure mechanism of β -<span class="hlt">G</span> a<span class="hlt">2</span>O3 under biaxial shear deformation that mimics indentation stress conditions. Finally, the general stacking fault energy (SFE) surface is calculated for the (001) surface from which we concluded that there is no intrinsic stacking fault structure for β -<span class="hlt">G</span> a<span class="hlt">2</span>O3 . The deformation modes and SFE calculations are essential to understand the intrinsic mechanical processes of this semiconductor material, which provides insightful guidance for designing high-performance semiconductor devices.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007CRPhy...8.1013I','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007CRPhy...8.1013I"><span>Supersymmetric Higgs and radiative electroweak <span class="hlt">breaking</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ibáñez, Luis E.; Ross, Graham G.</p> <p>2007-11-01</p> <p>We review the mechanism of radiative electroweak symmetry <span class="hlt">breaking</span> taking place in SUSY versions of the Standard Model. We further discuss different proposals for the origin of SUSY-<span class="hlt">breaking</span> and the corresponding <span class="hlt">induced</span> SUSY-<span class="hlt">breaking</span> soft terms. Several proposals for the understanding of the little hierarchy problem are critically discussed. To cite this article: L.E. Ibáñez, <span class="hlt">G.G</span>. Ross, C. R. Physique 8 (2007).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/10189155','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/10189155"><span>Sister <span class="hlt">chromatid</span> exchange analysis in workers exposed to noise and vibration.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Silva, M J; Carothers, A; Castelo Branco, N A; Dias, A; Boavida, M G</p> <p>1999-03-01</p> <p>There has been a growing interest in the combined effects of noise and vibration. In a population of aeronautical workers diagnosed with vibroacoustic disease (VAD), a large incidence of malignancy was detected. These workers were exposed to large pressure amplitude (LPA) (> or = 90 dB SPL) noise, with energy content concentrated within the low frequency (LF) bands (< or = 500 Hz) and whole-body vibration (WBV). To our knowledge, there are no studies conducted in humans or animals that address the issue of the potential genotoxic effects of vibration combined with noise. In the present study, the levels of sister <span class="hlt">chromatid</span> exchanges (SCE) and of cells with high frequencies of SCE (HFC) were analyzed in peripheral blood lymphocytes of workers employed in various occupations within the aeronautical industry. SCE and HFC were analyzed in lymphocytes of 50 workers occupationally exposed to noise and vibration and of 34 office-worker controls (<span class="hlt">G</span>0). The exposed group included: 10 hand-vibrating tool operators (<span class="hlt">G</span>1), 15 engine test cell technicians (<span class="hlt">G</span><span class="hlt">2</span>), 12 aircraft run-up technicians (<span class="hlt">G</span>3) and 13 Portuguese Air Force helicopter pilots (<span class="hlt">G</span>4). Groups <span class="hlt">2</span>-4 were exposed to WBV and LPALF noise; group 1 was exposed to LPA high frequency noise and local vibration. Statistical analysis of the mean SCE count per cell was carried out by multiple regression analysis comparing various predictor variables: type of exposure, duration of exposure, age, and cigarette consumption. Only cigarette consumption and type of exposure were found to be significantly correlated with the mean SCE frequency. After allowing for the effects of smoking, the analysis indicates that: 1) there was no significant difference between <span class="hlt">G</span>1 and <span class="hlt">G</span>0 (p > 0.05); <span class="hlt">2</span>) the differences between <span class="hlt">G</span><span class="hlt">2</span> and <span class="hlt">G</span>0, <span class="hlt">G</span>3 and <span class="hlt">G</span>0, <span class="hlt">G</span>4 and <span class="hlt">G</span>0 were all highly significant (p < 0.001); 3) there was no significant difference between <span class="hlt">G</span><span class="hlt">2</span> and <span class="hlt">G</span>3 (p > 0.05), nor between <span class="hlt">G</span><span class="hlt">2</span> and <span class="hlt">G</span>3 combined and <span class="hlt">G</span>4 (p > 0.05); and 4) <span class="hlt">G</span><span class="hlt">2</span> and <span class="hlt">G</span>4 combined had a</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21144119-low-concentration-arsenite-exacerbates-uvr-induced-dna-strand-breaks-inhibiting-parp-activity','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21144119-low-concentration-arsenite-exacerbates-uvr-induced-dna-strand-breaks-inhibiting-parp-activity"><span>Low concentration of arsenite exacerbates UVR-<span class="hlt">induced</span> DNA strand <span class="hlt">breaks</span> by inhibiting PARP-1 activity</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Qin Xujun; Department of Toxicology, Fourth Military Medical University, Xi'an, Shaanxi, 710032; Hudson, Laurie G.</p> <p>2008-10-01</p> <p>Epidemiological studies have associated arsenic exposure with many types of human cancers. Arsenic has also been shown to act as a co-carcinogen even at low concentrations. However, the precise mechanism of its co-carcinogenic action is unknown. Recent studies indicate that arsenic can interfere with DNA-repair processes. Poly(ADP-ribose) polymerase (PARP)-1 is a zinc-finger DNA-repair protein, which can promptly sense DNA strand <span class="hlt">breaks</span> and initiate DNA-repair pathways. In the present study, we tested the hypothesis that low concentrations of arsenic could inhibit PAPR-1 activity and so exacerbate levels of ultraviolet radiation (UVR)-<span class="hlt">induced</span> DNA strand <span class="hlt">breaks</span>. HaCat cells were treated with arsenite and/ormore » UVR, and then DNA strand <span class="hlt">breaks</span> were assessed by comet assay. Low concentrations of arsenite ({<=} <span class="hlt">2</span> {mu}M) alone did not <span class="hlt">induce</span> significant DNA strand <span class="hlt">breaks</span>, but greatly enhanced the DNA strand <span class="hlt">breaks</span> <span class="hlt">induced</span> by UVR. Further studies showed that <span class="hlt">2</span> {mu}M arsenite effectively inhibited PARP-1 activity. Zinc supplementation of arsenite-treated cells restored PARP-1 activity and significantly diminished the exacerbating effect of arsenite on UVR-<span class="hlt">induced</span> DNA strand <span class="hlt">breaks</span>. Importantly, neither arsenite treatment, nor zinc supplementation changed UVR-triggered reactive oxygen species (ROS) formation, suggesting that their effects upon UVR-<span class="hlt">induced</span> DNA strand <span class="hlt">breaks</span> are not through a direct free radical mechanism. Combination treatments of arsenite with PARP-1 inhibitor 3-aminobenzamide or PARP-1 siRNA demonstrate that PARP-1 is the target of arsenite. Together, these findings show that arsenite at low concentration exacerbates UVR-<span class="hlt">induced</span> DNA strand <span class="hlt">breaks</span> by inhibiting PARP-1 activity, which may represent an important mechanism underlying the co-carcinogenicity of arsenic.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3514784','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3514784"><span>Cell elongation is an adaptive response for clearing long <span class="hlt">chromatid</span> arms from the cleavage plane</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Kotadia, Shaila; Montembault, Emilie; Sullivan, William</p> <p>2012-01-01</p> <p>Chromosome segregation must be coordinated with cell cleavage to ensure correct transmission of the genome to daughter cells. Here we identify a novel mechanism by which Drosophila melanogaster neuronal stem cells coordinate sister <span class="hlt">chromatid</span> segregation with cleavage furrow ingression. Cells adapted to a dramatic increase in <span class="hlt">chromatid</span> arm length by transiently elongating during anaphase/telophase. The degree of cell elongation correlated with the length of the trailing <span class="hlt">chromatid</span> arms and was concomitant with a slight increase in spindle length and an enlargement of the zone of cortical myosin distribution. Rho guanine-nucleotide exchange factor (Pebble)–depleted cells failed to elongate during segregation of long <span class="hlt">chromatids</span>. As a result, Pebble-depleted adult flies exhibited morphological defects likely caused by cell death during development. These studies reveal a novel pathway linking trailing <span class="hlt">chromatid</span> arms and cortical myosin that ensures the clearance of <span class="hlt">chromatids</span> from the cleavage plane at the appropriate time during cytokinesis, thus preserving genome integrity. PMID:23185030</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28489060','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28489060"><span>The Effect of VPA on Increasing Radiosensitivity in Osteosarcoma Cells and Primary-Culture Cells from Chemical Carcinogen-<span class="hlt">Induced</span> Breast Cancer in Rats.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liu, Guochao; Wang, Hui; Zhang, Fengmei; Tian, Youjia; Tian, Zhujun; Cai, Zuchao; Lim, David; Feng, Zhihui</p> <p>2017-05-10</p> <p>This study explored whether valproic acid (VPA, a histone deacetylase inhibitor) could radiosensitize osteosarcoma and primary-culture tumor cells, and determined the mechanism of VPA-<span class="hlt">induced</span> radiosensitization. The working system included osteosarcoma cells (U<span class="hlt">2</span>OS) and primary-culture cells from chemical carcinogen (DMBA)-<span class="hlt">induced</span> breast cancer in rats; and clonogenic survival, immunofluorescence, fluorescent in situ hybridization (FISH) for chromosome aberrations, and comet assays were used in this study. It was found that VPA at the safe or critical safe concentration of 0.5 or 1.0 mM VPA could result in the accumulation of more ionizing radiation (IR)-<span class="hlt">induced</span> DNA double strand <span class="hlt">breaks</span>, and increase the cell radiosensitivity. VPA-<span class="hlt">induced</span> radiosensitivity was associated with the inhibition of DNA repair activity in the working systems. In addition, the chromosome aberrations including chromosome <span class="hlt">breaks</span>, <span class="hlt">chromatid</span> <span class="hlt">breaks</span>, and radial structures significantly increased after the combination treatment of VPA and IR. Importantly, the results obtained by primary-culture cells from the tissue of chemical carcinogen-<span class="hlt">induced</span> breast cancer in rats further confirmed our findings. The data in this study demonstrated that VPA at a safe dose was a radiosensitizer for osteosarcoma and primary-culture tumor cells through suppressing DNA-double strand <span class="hlt">breaks</span> repair function.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5454939','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5454939"><span>The Effect of VPA on Increasing Radiosensitivity in Osteosarcoma Cells and Primary-Culture Cells from Chemical Carcinogen-<span class="hlt">Induced</span> Breast Cancer in Rats</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Liu, Guochao; Wang, Hui; Zhang, Fengmei; Tian, Youjia; Tian, Zhujun; Cai, Zuchao; Lim, David; Feng, Zhihui</p> <p>2017-01-01</p> <p>This study explored whether valproic acid (VPA, a histone deacetylase inhibitor) could radiosensitize osteosarcoma and primary-culture tumor cells, and determined the mechanism of VPA-<span class="hlt">induced</span> radiosensitization. The working system included osteosarcoma cells (U<span class="hlt">2</span>OS) and primary-culture cells from chemical carcinogen (DMBA)-<span class="hlt">induced</span> breast cancer in rats; and clonogenic survival, immunofluorescence, fluorescent in situ hybridization (FISH) for chromosome aberrations, and comet assays were used in this study. It was found that VPA at the safe or critical safe concentration of 0.5 or 1.0 mM VPA could result in the accumulation of more ionizing radiation (IR)-<span class="hlt">induced</span> DNA double strand <span class="hlt">breaks</span>, and increase the cell radiosensitivity. VPA-<span class="hlt">induced</span> radiosensitivity was associated with the inhibition of DNA repair activity in the working systems. In addition, the chromosome aberrations including chromosome <span class="hlt">breaks</span>, <span class="hlt">chromatid</span> <span class="hlt">breaks</span>, and radial structures significantly increased after the combination treatment of VPA and IR. Importantly, the results obtained by primary-culture cells from the tissue of chemical carcinogen-<span class="hlt">induced</span> breast cancer in rats further confirmed our findings. The data in this study demonstrated that VPA at a safe dose was a radiosensitizer for osteosarcoma and primary-culture tumor cells through suppressing DNA-double strand <span class="hlt">breaks</span> repair function. PMID:28489060</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3096717','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3096717"><span>Revised genetic requirements for the decatenation <span class="hlt">G</span><span class="hlt">2</span> checkpoint: the role of ATM</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Bower, Jacquelyn J.; Zhou, Yingchun; Zhou, Tong; Simpson, Dennis A.; Arlander, Sonnet J.; Paules, Richard S.; Cordeiro-Stone, Marila; Kaufmann, William K.</p> <p>2010-01-01</p> <p>The decatenation <span class="hlt">G</span><span class="hlt">2</span> checkpoint is proposed to delay cellular progression from <span class="hlt">G</span><span class="hlt">2</span> into mitosis when intertwined daughter <span class="hlt">chromatids</span> are insufficiently decatenated. Previous studies indicated that the ATM- and Rad3-related (ATR) checkpoint kinase, but not the ataxia telangiectasia-mutated (ATM) kinase, was required for decatenation <span class="hlt">G</span><span class="hlt">2</span> checkpoint function. Here, we show that the method used to quantify decatenation <span class="hlt">G</span><span class="hlt">2</span> checkpoint function can influence the identification of genetic requirements for the checkpoint. Normal human diploid fibroblast (NHDF) lines responded to the topoisomerase II (topo II) catalytic inhibitor ICRF-193 with a stringent <span class="hlt">G</span><span class="hlt">2</span> arrest and a reduction in the mitotic index. While siRNA-mediated depletion of ATR and CHEK1 increased the mitotic index in ICRF-193 treated NHDF lines, depletion of these proteins did not affect the mitotic entry rate, indicating that the decatenation <span class="hlt">G</span><span class="hlt">2</span> checkpoint was functional. These results suggest that ATR and CHEK1 are not required for the decatenation <span class="hlt">G</span><span class="hlt">2</span> checkpoint, but may influence mitotic exit after inhibition of topo II. A re-evaluation of ataxia telangiectasia (AT) cell lines using the mitotic entry assay indicated that ATM was required for the decatenation <span class="hlt">G</span><span class="hlt">2</span> checkpoint. Three NHDF cell lines responded to ICRF-193 with a mean 98% inhibition of the mitotic entry rate. Examination of the mitotic entry rates in AT fibroblasts upon treatment with ICRF-193 revealed a significantly attenuated decatenation <span class="hlt">G</span><span class="hlt">2</span> checkpoint response, with a mean 59% inhibition of the mitotic entry rate. In addition, a normal lymphoblastoid line exhibited a 95% inhibition of the mitotic entry rate after incubation with ICRF-193, whereas two AT lymphoblastoid lines displayed only 36% and 20% inhibition of the mitotic entry rate. Stable depletion of ATM in normal human fibroblasts with short hairpin RNA also attenuated decatenation <span class="hlt">G</span><span class="hlt">2</span> checkpoint function by an average of 40%. Western immunoblot analysis demonstrated that treatment with ICRF</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2910038','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2910038"><span>Evidence that MEK1 positively promotes interhomologue double-strand <span class="hlt">break</span> repair</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Terentyev, Yaroslav; Johnson, Rebecca; Neale, Matthew J.; Khisroon, Muhammad; Bishop-Bailey, Anna; Goldman, Alastair S. H.</p> <p>2010-01-01</p> <p>During meiosis there is an imperative to create sufficient crossovers for homologue segregation. This can be achieved during repair of programmed DNA double-strand <span class="hlt">breaks</span> (DSBs), which are biased towards using a homologue rather than sister <span class="hlt">chromatid</span> as a repair template. Various proteins contribute to this bias, one of which is a meiosis specific kinase Mek1. It has been proposed that Mek1 establishes the bias by creating a barrier to sister <span class="hlt">chromatid</span> repair, as distinct from enforcing strand invasion with the homologue. We looked for evidence that Mek1 positively stimulates strand invasion of the homologue. This was done by analysing repair of DSBs <span class="hlt">induced</span> by the VMA1-derived endonuclease (VDE) and flanked by directly repeated sequences that can be used for intrachromatid single-strand annealing (SSA). SSA competes with interhomologue strand invasion significantly more successfully when Mek1 function is lost. We suggest the increase in intrachromosomal SSA reflects an opportunistic default repair pathway due to loss of a MEK1 stimulated bias for strand invasion of the homologous chromosome. Making use of an inhibitor sensitive mek1-as1 allele, we found that Mek1 function influences the repair pathway throughout the first4–5 h of meiosis. Perhaps reflecting a particular need to create bias for successful interhomologue events before chromosome pairing is complete. PMID:20223769</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2938206','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2938206"><span>Electron attachment-<span class="hlt">induced</span> DNA single-strand <span class="hlt">breaks</span> at the pyrimidine sites</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Gu, Jiande; Wang, Jing; Leszczynski, Jerzy</p> <p>2010-01-01</p> <p>To elucidate the contribution of pyrimidine in DNA strand <span class="hlt">breaks</span> caused by low-energy electrons (LEEs), theoretical investigations of the LEE attachment-<span class="hlt">induced</span> C3′–O3′, and C5′–O5′ σ bond as well as N-glycosidic bond <span class="hlt">breaking</span> of <span class="hlt">2</span>′-deoxycytidine-3′,5′-diphosphate and <span class="hlt">2</span>′-deoxythymidine-3′,5′-diphosphate were performed using the B3LYP/DZP++ approach. The base-centered radical anions are electronically stable enough to assure that either the C–O or glycosidic bond <span class="hlt">breaking</span> processes might compete with the electron detachment and yield corresponding radical fragments and anions. In the gas phase, the computed glycosidic bond <span class="hlt">breaking</span> activation energy (24.1 kcal/mol) excludes the base release pathway. The low-energy barrier for the C3′–O3′ σ bond cleavage process (∼6.0 kcal/mol for both cytidine and thymidine) suggests that this reaction pathway is the most favorable one as compared to other possible pathways. On the other hand, the relatively low activation energy barrier (∼14 kcal/mol) for the C5′–O5′ σ bond cleavage process indicates that this bond <span class="hlt">breaking</span> pathway could be possible, especially when the incident electrons have relatively high energy (a few electronvolts). The presence of the polarizable medium greatly increases the activation energies of either C–O σ bond cleavage processes or the N-glycosidic bond <span class="hlt">breaking</span> process. The only possible pathway that dominates the LEE-<span class="hlt">induced</span> DNA single strands in the presence of the polarizable surroundings (such as in an aqueous solution) is the C3′–O3′ σ bond cleavage (the relatively low activation energy barrier, ∼13.4 kcal/mol, has been predicted through a polarizable continuum model investigation). The qualitative agreement between the ratio for the bond <span class="hlt">breaks</span> of C5′–O5′, C3′–O3′ and N-glycosidic bonds observed in the experiment of oligonucleotide tetramer CGAT and the theoretical sequence of the bond <span class="hlt">breaking</span> reaction pathways have been found</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22207822-xanthorrhizol-induced-dna-fragmentation-hepg2-cells-involving-bcl-family-proteins','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22207822-xanthorrhizol-induced-dna-fragmentation-hepg2-cells-involving-bcl-family-proteins"><span>Xanthorrhizol <span class="hlt">induced</span> DNA fragmentation in Hep<span class="hlt">G</span><span class="hlt">2</span> cells involving Bcl-<span class="hlt">2</span> family proteins</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Tee, Thiam-Tsui, E-mail: thiamtsu@yahoo.com; Cheah, Yew-Hoong; Bioassay Unit, Herbal Medicine Research Center, Institute for Medical Research, Jalan Pahang, Kuala Lumpur</p> <p></p> <p>Highlights: Black-Right-Pointing-Pointer We isolated xanthorrhizol, a sesquiterpenoid compound from Curcuma xanthorrhiza. Black-Right-Pointing-Pointer Xanthorrhizol <span class="hlt">induced</span> apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells as observed using SEM. Black-Right-Pointing-Pointer Apoptosis in xanthorrhizol-treated Hep<span class="hlt">G</span><span class="hlt">2</span> cells involved Bcl-<span class="hlt">2</span> family proteins. Black-Right-Pointing-Pointer DNA fragmentation was observed in xanthorrhizol-treated Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Black-Right-Pointing-Pointer DNA fragmentation maybe due to cleavage of PARP and DFF45/ICAD proteins. -- Abstract: Xanthorrhizol is a plant-derived pharmacologically active sesquiterpenoid compound isolated from Curcuma xanthorrhiza. Previously, we have reported that xanthorrhizol inhibited the proliferation of Hep<span class="hlt">G</span><span class="hlt">2</span> human hepatoma cells by <span class="hlt">inducing</span> apoptotic cell death via caspase activation. Here, we attempt to further elucidate the mode of action ofmore » xanthorrhizol. Apoptosis in xanthorrhizol-treated Hep<span class="hlt">G</span><span class="hlt">2</span> cells as observed by scanning electron microscopy was accompanied by truncation of BID; reduction of both anti-apoptotic Bcl-<span class="hlt">2</span> and Bcl-X{sub L} expression; cleavage of PARP and DFF45/ICAD proteins and DNA fragmentation. Taken together, these results suggest xanthorrhizol as a potent antiproliferative agent on Hep<span class="hlt">G</span><span class="hlt">2</span> cells by <span class="hlt">inducing</span> apoptosis via Bcl-<span class="hlt">2</span> family members. Hence we proposed that xanthorrhizol could be used as an anti-liver cancer drug for future studies.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=86942','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=86942"><span>Saccharomyces cerevisiae CTF18 and CTF4 Are Required for Sister <span class="hlt">Chromatid</span> Cohesion</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Hanna, Joseph S.; Kroll, Evgueny S.; Lundblad, Victoria; Spencer, Forrest A.</p> <p>2001-01-01</p> <p>CTF4 and CTF18 are required for high-fidelity chromosome segregation. Both exhibit genetic and physical ties to replication fork constituents. We find that absence of either CTF4 or CTF18 causes sister <span class="hlt">chromatid</span> cohesion failure and leads to a preanaphase accumulation of cells that depends on the spindle assembly checkpoint. The physical and genetic interactions between CTF4, CTF18, and core components of replication fork complexes observed in this study and others suggest that both gene products act in association with the replication fork to facilitate sister <span class="hlt">chromatid</span> cohesion. We find that Ctf18p, an RFC1-like protein, directly interacts with Rfc<span class="hlt">2</span>p, Rfc3p, Rfc4p, and Rfc5p. However, Ctf18p is not a component of biochemically purified proliferating cell nuclear antigen loading RF-C, suggesting the presence of a discrete complex containing Ctf18p, Rfc<span class="hlt">2</span>p, Rfc3p, Rfc4p, and Rfc5p. Recent identification and characterization of the budding yeast polymerase κ, encoded by TRF4, strongly supports a hypothesis that the DNA replication machinery is required for proper sister <span class="hlt">chromatid</span> cohesion. Analogous to the polymerase switching role of the bacterial and human RF-C complexes, we propose that budding yeast RF-CCTF18 may be involved in a polymerase switch event that facilities sister <span class="hlt">chromatid</span> cohesion. The requirement for CTF4 and CTF18 in robust cohesion identifies novel roles for replication accessory proteins in this process. PMID:11287619</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2230545','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2230545"><span>Deficient Repair of Particulate Hexavalent chromium-<span class="hlt">Induced</span> DNA Double Strand <span class="hlt">Breaks</span> Leads To Neoplastic Transformation</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Xie, Hong; Wise, Sandra S.; Wise, John. P.</p> <p>2008-01-01</p> <p>Hexavalent chromium (Cr(VI)) is a potent respiratory toxicant and carcinogen. The most carcinogenic forms of Cr(VI) are the particulate salts such as lead chromate, which deposit and persist in the respiratory tract after inhalation. We demonstrate here that particulate chromate <span class="hlt">induces</span> DNA double strand <span class="hlt">breaks</span> in human lung cells with 0.1, 0.5, and 1 ug/cm<span class="hlt">2</span> lead chromate <span class="hlt">inducing</span> 1.5, <span class="hlt">2</span> and 5 relative increases in the percent of DNA in the comet tail, respectively. These lesions are repaired within 24 h and require Mre11 expression for their repair. Particulate chromate also caused Mre11 to co-localize with gamma-H<span class="hlt">2</span>A.X and ATM. Failure to repair these <span class="hlt">breaks</span> with Mre11 <span class="hlt">induced</span> neoplastic transformation including loss of cell contact inhibition and anchorage independent growth. A 5-day exposure to lead chromate <span class="hlt">induced</span> loss of cell contact inhibition in a concentration-dependent manner with 0, 0.1, 0.5 and 1 ug/cm<span class="hlt">2</span> lead chromate <span class="hlt">inducing</span> 1, 78 and 103 foci in 20 dishes, respectively. These data indicate that Mre11 is critical to repairing particulate Cr(VI)-<span class="hlt">induced</span> double strand <span class="hlt">breaks</span> and preventing Cr(VI)-<span class="hlt">induced</span> neoplastic transformation. PMID:18023605</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25553380','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25553380"><span>Selective <span class="hlt">chromatid</span> segregation mechanism for Bruchus wings piebald color.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Klar, Amar J S</p> <p>2015-01-01</p> <p>The mechanisms of asymmetric organ development have been under intensive investigation for years, yet the proposed mechanisms remain controversial (1-3). The female Bruchus quadrimaculatus beetle insect develops two black-colored spots bilaterally located on each upper elytra wing by an unknown mechanism. Fifty percent of the P (for piebald, two colors) gene homozygous mutant insects, described in 1925, had a normal left elytrum (with two black spots) and an abnormal right elytrum (with two red spots) and the balance supported the converse lateralized pigment arrangement (4). Rather than supporting the conventional morphogen model for the wings pigmentation development, their biological origin is explained here with the somatic strand-specific epigenetic imprinting and selective sister <span class="hlt">chromatid</span> segregation (SSIS) mechanism (5). We propose that the P gene product performs the selective sister <span class="hlt">chromatid</span> segregation function to produce symmetric cell division of a specific cell during embryogenesis to result in the bilateral symmetric development of elytra black color spots and that the altered <span class="hlt">chromatid</span> segregation pattern of the mutant causes asymmetric cell division to confer the piebald phenotype. </p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li class="active"><span>4</span></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_4 --> <div id="page_5" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li class="active"><span>5</span></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="81"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5587752','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5587752"><span>FANCJ helicase controls the balance between short- and long-tract gene conversions between sister <span class="hlt">chromatids</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Nath, Sarmi; Somyajit, Kumar; Mishra, Anup; Scully, Ralph</p> <p>2017-01-01</p> <p>Abstract The FANCJ DNA helicase is linked to hereditary breast and ovarian cancers as well as bone marrow failure disorder Fanconi anemia (FA). Although FANCJ has been implicated in the repair of DNA double-strand <span class="hlt">breaks</span> (DSBs) by homologous recombination (HR), the molecular mechanism underlying the tumor suppressor functions of FANCJ remains obscure. Here, we demonstrate that FANCJ deficient human and hamster cells exhibit reduction in the overall gene conversions in response to a site-specific chromosomal DSB <span class="hlt">induced</span> by I-SceI endonuclease. Strikingly, the gene conversion events were biased in favour of long-tract gene conversions in FANCJ depleted cells. The fine regulation of short- (STGC) and long-tract gene conversions (LTGC) by FANCJ was dependent on its interaction with BRCA1 tumor suppressor. Notably, helicase activity of FANCJ was essential for controlling the overall HR and in terminating the extended repair synthesis during sister <span class="hlt">chromatid</span> recombination (SCR). Moreover, cells expressing FANCJ pathological mutants exhibited defective SCR with an increased frequency of LTGC. These data unravel the novel function of FANCJ helicase in regulating SCR and SCR associated gene amplification/duplications and imply that these functions of FANCJ are crucial for the genome maintenance and tumor suppression. PMID:28911102</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/6865499','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/6865499"><span>Cyclophosphamide-<span class="hlt">induced</span> in vivo sister <span class="hlt">chromatid</span> exchange in Mus Musculus. II: Effect of age and genotype on sister <span class="hlt">chromatid</span> exchange, micronuclei and metaphase index.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Reimer, D L; Singh, S M</p> <p>1983-01-01</p> <p>In vivo cyclophosphamide-<span class="hlt">induced</span> sister <span class="hlt">chromatid</span> exchanges (SCEs) micronuclei, and metaphase indices were assessed in two age groups (10.8 +/- 0.9 weeks' an 33.1 +/- 1.3 weeks' old) of female mice from three genetic strains (C3H/S, C57BL/6J, and Balb/c). In general, older animals showed diminished SCE induction over their younger counterparts. The relative difference between individuals of the two ages is strain-dependent. Unlike C57BL/6J and Balb/c, strain C3H/S showed significantly lower SCE values in the older animals at every cyclophosphamide treatment. It may reflect on the possible involvement of genetic determinant(s) for the component(s) of SCE formation during aging. Frequencies of micronuclei, however, were consistently higher in older animals than in their younger counterparts. Furthermore, cytotoxicity of cyclophosphamide, as reflected in metaphase indices, was also higher in older animals. Lower metaphase indices associated with higher micronuclei levels in older individuals may suggest a decline in the rate of cellular replication in these animals. Furthermore, the lower metaphase indices associated with lower SCE values, and increasing micronuclei levels accompanied by decreasing SCE frequencies in older animals, may reflect reduced DNA repair ability during aging. These results support the hypothesis of genotype-dependent decline in the rate of DNA repair and replication during aging, particularly under stressed conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1216365','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1216365"><span>Lack of Spontaneous Sister <span class="hlt">Chromatid</span> Exchanges in Somatic Cells of DROSOPHILA MELANOGASTER</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Gatti, M.; Santini, G.; Pimpinelli, S.; Olivieri, G.</p> <p>1979-01-01</p> <p>Neural ganglia of wild type third-instar larvae of Drosophila melanogaster were incubated for 13 hours at various concentrations of BUdR (1, 3, 9, 27 µg/ml). Metaphases were collected with colchicine, stained with Hoechst 33258, and scored under a fluorescence microscope. Metaphases in which the sister <span class="hlt">chromatids</span> were clearly differentiated were scored for the presence of sister-<span class="hlt">chromatid</span> exchanges (SCEs). At the lowest concentration of BUdR (1 µg/ml), no SCEs were observed in either male or female neuroblasts. The SCEs were found at the higher concentrations of BUdR (3, 9 and 27 µg/ml) and with a greater frequency in females than in males. Therefore SCEs are not a spontaneous phenomenon in D. melanogaster, but are <span class="hlt">induced</span> by BUdR incorporated in the DNA. A striking nonrandomness was found in the distribution of SCEs along the chromosomes. More than a third of the SCEs were clustered in the junctions between euchromatin and heterochromatin. The remaining SCEs were preferentially localized within the heterochromatic regions of the X chromosome and the autosomes and primarily on the entirely heterochromatic Y chromosome.—In order to find an alternative way of measuring the frequency of SCEs in Drosophila neuroblasts, the occurrence of double dicentric rings was studied in two stocks carrying monocentric ring-X chromosomes. One ring chromosome, C(1)TR 94–<span class="hlt">2</span>, shows a rate of dicentric ring formation corresponding to the frequency of SCEs observed in the BUdR-labelled rod chromosomes. The other ring studied, R(1)<span class="hlt">2</span>, exhibits a frequency of SCEs higher than that observed with both C(1)TR 94–<span class="hlt">2</span> and rod chromosomes. PMID:109350</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24852491','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24852491"><span>Health assessment of gasoline and fuel oxygenate vapors: micronucleus and sister <span class="hlt">chromatid</span> exchange evaluations.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Schreiner, Ceinwen A; Hoffman, Gary M; Gudi, Ramadevi; Clark, Charles R</p> <p>2014-11-01</p> <p>Micronucleus and sister <span class="hlt">chromatid</span> exchange (SCE) tests were performed for vapor condensate of baseline gasoline (BGVC), or gasoline with oxygenates, methyl tert-butyl ether (<span class="hlt">G</span>/MTBE), ethyl tert butyl ether (<span class="hlt">G</span>/ETBE), t-amyl methyl ether (<span class="hlt">G</span>/TAME), diisopropyl ether (<span class="hlt">G</span>/DIPE), t-butyl alcohol (TBA), or ethanol (<span class="hlt">G</span>/EtOH). Sprague Dawley rats (the same 5/sex/group for both endpoints) were exposed to 0, 2000, 10,000, or 20,000mg/m(3) of each condensate, 6h/day, 5days/week over 4weeks. Positive controls (5/sex/test) were given cyclophosphamide IP, 24h prior to sacrifice at 5mg/kg (SCE test) and 40mg/kg (micronucleus test). Blood was collected from the abdominal aorta for the SCE test and femurs removed for the micronucleus test. Blood cell cultures were treated with 5μ<span class="hlt">g</span>/ml bromodeoxyuridine (BrdU) for SCE evaluation. No significant increases in micronucleated immature erythrocytes were observed for any test material. Statistically significant increases in SCE were observed in rats given BGVC alone or in female rats given <span class="hlt">G</span>/MTBE. <span class="hlt">G</span>/TAME <span class="hlt">induced</span> increased SCE in both sexes at the highest dose only. Although DNA perturbation was observed for several samples, DNA damage was not expressed as increased micronuclei in bone marrow cells. Inclusion of oxygenates in gasoline did not increase the effects of gasoline alone or produce a cytogenetic hazard. Copyright © 2014 Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25294323','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25294323"><span>Sterigmatocystin <span class="hlt">induces</span> <span class="hlt">G</span>1 arrest in primary human esophageal epithelial cells but <span class="hlt">induces</span> <span class="hlt">G</span><span class="hlt">2</span> arrest in immortalized cells: key mechanistic differences in these two models.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Juan; Huang, Shujuan; Xing, Lingxiao; Cui, Jinfeng; Tian, Ziqiang; Shen, Haitao; Jiang, Xiujuan; Yan, Xia; Wang, Junling; Zhang, Xianghong</p> <p>2015-11-01</p> <p>Sterigmatocystin (ST), a mycotoxin commonly found in food and feed commodities, has been classified as a "possible human carcinogen." Our previous studies suggested that ST exposure might be a risk factor for esophageal cancer and that ST may <span class="hlt">induce</span> DNA damage and <span class="hlt">G</span><span class="hlt">2</span> phase arrest in immortalized human esophageal epithelial cells (Het-1A). To further confirm and explore the cellular responses of ST in human esophageal epithelia, we comparatively evaluated DNA damage, cell cycle distribution and the relative mechanisms in primary cultured human esophageal epithelial cells (EPC), which represent a more representative model of the in vivo state, and Het-1A cells. In this study, we found that ST could <span class="hlt">induce</span> DNA damage in both EPC and Het-1A cells but led to <span class="hlt">G</span>1 phase arrest in EPC cells and <span class="hlt">G</span><span class="hlt">2</span> phase arrest in Het-1A cells. Furthermore, our results indicated that the activation of the ATM-Chk<span class="hlt">2</span> pathway was involved in ST-<span class="hlt">induced</span> <span class="hlt">G</span>1 phase arrest in EPC cells, whereas the p53-p21 pathway activation in ST-<span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span> phase arrest in Het-1A cells. Studies have demonstrated that SV40 large T-antigen (SV40LT) may disturb cell cycle progression by inactivating some of the proteins involved in the <span class="hlt">G</span>1/S checkpoint. Het-1A is a non-cancerous epithelial cell line immortalized by SV40LT. To evaluate the possible perturbation effect of SV40LT on ST-<span class="hlt">induced</span> cell cycle disturbance in Het-1A cells, we knocked down SV40LT of Het-1A cells with siRNA and found that under this condition, ST-<span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span> arrest was significantly attenuated, whereas the proportion of cells in the <span class="hlt">G</span>1 phase was significantly increased. Furthermore, SV40LT-siRNA also inhibited the activation of the p53-p21 signaling pathway <span class="hlt">induced</span> by ST. In conclusion, our data indicated that ST could <span class="hlt">induce</span> DNA damage in both primary cultured and immortalized esophageal epithelial cells. In primary human esophageal epithelial cells, ST <span class="hlt">induced</span> DNA damage and then triggered the ATM-Chk<span class="hlt">2</span> pathway, resulting in <span class="hlt">G</span>1 phase arrest</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/1585081','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/1585081"><span>Effect of chloramphenicol on sister <span class="hlt">chromatid</span> exchange in bovine fibroblasts.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Arruga, M V; Catalan, J; Moreno, C</p> <p>1992-03-01</p> <p>The genotoxic potential of different chloramphenicol concentrations (5, 20, 40 and 60 micrograms ml-1) was investigated in bovine fibroblast primary lines by sister <span class="hlt">chromatid</span> exchange assay. Chloramphenicol acted for long enough to ensure similar effects to persistent storage in the kidney. In this experiment 10 micrograms ml-1 of 5-bromodeoxyuridine was added for 60 hours for all doses of chloramphenicol and to the control. When the tissue culture cells were exposed to increasing doses, increased numbers of sister <span class="hlt">chromatid</span> exchanges developed. Differences were significantly different to the control.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040173021&hterms=protein+synthesis&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Dprotein%2Bsynthesis','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040173021&hterms=protein+synthesis&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Dprotein%2Bsynthesis"><span>Activation of <span class="hlt">G</span> proteins mediates flow-<span class="hlt">induced</span> prostaglandin E<span class="hlt">2</span> production in osteoblasts</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Reich, K. M.; McAllister, T. N.; Gudi, S.; Frangos, J. A.</p> <p>1997-01-01</p> <p>Interstitial fluid flow may play a role in load-<span class="hlt">induced</span> bone remodeling. Previously, we have shown that fluid flow stimulates osteoblast production of cAMP inositol trisphosphate (IP3), and PGE<span class="hlt">2</span>. Flow-<span class="hlt">induced</span> increases in cAMP and IP3 were shown to be a result of PG production. Thus, PGE<span class="hlt">2</span> production appears to be an important component in fluid flow <span class="hlt">induced</span> signal transduction. In the present study, we investigated the mechanism of flow-<span class="hlt">induced</span> PGE<span class="hlt">2</span> synthesis. Flow-<span class="hlt">induced</span> a 20-fold increase in PGE<span class="hlt">2</span> production in osteoblasts. Increases were also observed with ALF4-(10mM) (98-fold), an activator of guanidine nucleotide-binding proteins (<span class="hlt">G</span> proteins), and calcium ionophore A23187 (<span class="hlt">2</span> microM) (100-fold) in stationary cells. We then investigated whether flow stimulation is mediated by <span class="hlt">G</span> proteins and increases in intracellular calcium. Flow-<span class="hlt">induced</span> PGE<span class="hlt">2</span> production was inhibited by the <span class="hlt">G</span> protein inhibitors GDP beta S (100 microM) and pertussis toxin (1 microgram/ml) by 83% and 72%, respectively. Chelation of extracellular calcium by EGTA (<span class="hlt">2</span> mM) and intracellular calcium by quin-<span class="hlt">2</span>/AM (30 microM) blocked flow stimulation by 87% and 67%, respectively. These results suggest that <span class="hlt">G</span> proteins and calcium play an important role in mediating mechanochemical signal transduction in osteoblasts.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2567865','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2567865"><span>Shugoshin1 May Play Important Roles in Separation of Homologous Chromosomes and Sister <span class="hlt">Chromatids</span> during Mouse Oocyte Meiosis</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Yin, Shen; Ai, Jun-Shu; Shi, Li-Hong; Wei, Liang; Yuan, Ju; Ouyang, Ying-Chun; Hou, Yi; Chen, Da-Yuan; Schatten, Heide; Sun, Qing-Yuan</p> <p>2008-01-01</p> <p>Background Homologous chromosomes separate in meiosis I and sister <span class="hlt">chromatids</span> separate in meiosis II, generating haploid gametes. To address the question why sister <span class="hlt">chromatids</span> do not separate in meiosis I, we explored the roles of Shogoshin1 (Sgo1) in chromosome separation during oocyte meiosis. Methodology/Principal Findings Sgo1 function was evaluated by exogenous overexpression to enhance its roles and RNAi to suppress its roles during two meioses of mouse oocytes. Immunocytochemistry and chromosome spread were used to evaluate phenotypes. The exogenous Sgo1 overexpression kept homologous chromosomes and sister <span class="hlt">chromatids</span> not to separate in meiosis I and meiosis II, respectively, while the Sgo1 RNAi promoted premature separation of sister <span class="hlt">chromatids</span>. Conclusions Our results reveal that prevention of premature separation of sister <span class="hlt">chromatids</span> in meiosis I requires the retention of centromeric Sgo1, while normal separation of sister <span class="hlt">chromatids</span> in meiosis II requires loss of centromeric Sgo1. PMID:18949044</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/10753184','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/10753184"><span>XPD polymorphisms: effects on DNA repair proficiency.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lunn, R M; Helzlsouer, K J; Parshad, R; Umbach, D M; Harris, E L; Sanford, K K; Bell, D A</p> <p>2000-04-01</p> <p>XPD codes for a DNA helicase involved in transcription and nucleotide excision repair. Rare XPD mutations diminish nucleotide excision repair resulting in hypersensitivity to UV light and increased risk of skin cancer. Several polymorphisms in this gene have been identified but their impact on DNA repair is not known. We compared XPD genotypes at codons 312 and 751 with DNA repair proficiency in 31 women. XPD genotypes were measured by PCR-RFLP. DNA repair proficiency was assessed using a cytogenetic assay that detects X-ray <span class="hlt">induced</span> <span class="hlt">chromatid</span> aberrations (<span class="hlt">breaks</span> and gaps). <span class="hlt">Chromatid</span> aberrations were scored per 100 metaphase cells following incubation at 37 degrees C (1.5 h after irradiation) to allow for repair of DNA damage. Individuals with the Lys/Lys codon 751 XPD genotype had a higher number of <span class="hlt">chromatid</span> aberrations (132/100 metaphase cells) than those having a 751Gln allele (34/100 metaphase cells). Individuals having greater than 60 <span class="hlt">chromatid</span> <span class="hlt">breaks</span> plus gaps were categorized as having sub-optimal repair. Possessing a Lys/Lys751 genotype increased the risk of sub-optimal DNA repair (odds ratio = 7.<span class="hlt">2</span>, 95% confidence interval = 1.01-87.7). The Asp312Asn XPD polymorphism did not appear to affect DNA repair proficiency. These results suggest that the Lys751 (common) allele may alter the XPD protein product resulting in sub-optimal repair of X-ray-<span class="hlt">induced</span> DNA damage.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2734174','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2734174"><span>The Cellular Phenotype of Roberts Syndrome Fibroblasts as Revealed by Ectopic Expression of ESCO<span class="hlt">2</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>van der Lelij, Petra; van Gosliga, Djoke; Oostra, Anneke B.; Steltenpool, Jûrgen; de Groot, Jan; Scheper, Rik J.; Wolthuis, Rob M.; Waisfisz, Quinten; Darroudi, Firouz; Joenje, Hans; de Winter, Johan P.</p> <p>2009-01-01</p> <p>Cohesion between sister <span class="hlt">chromatids</span> is essential for faithful chromosome segregation. In budding yeast, the acetyltransferase Eco1/Ctf7 establishes cohesion during DNA replication in S phase and in response to DNA double strand <span class="hlt">breaks</span> in <span class="hlt">G</span><span class="hlt">2</span>/M phase. In humans two Eco1 orthologs exist: ESCO1 and ESCO<span class="hlt">2</span>. Both proteins are required for proper sister <span class="hlt">chromatid</span> cohesion, but their exact function is unclear at present. Since ESCO<span class="hlt">2</span> has been identified as the gene defective in the rare autosomal recessive cohesinopathy Roberts syndrome (RBS), cells from RBS patients can be used to elucidate the role of ESCO<span class="hlt">2</span>. We investigated for the first time RBS cells in comparison to isogenic controls that stably express V5- or GFP-tagged ESCO<span class="hlt">2</span>. We show that the sister <span class="hlt">chromatid</span> cohesion defect in the transfected cell lines is rescued and suggest that ESCO<span class="hlt">2</span> is regulated by proteasomal degradation in a cell cycle-dependent manner. In comparison to the corrected cells RBS cells were hypersensitive to the DNA-damaging agents mitomycin C, camptothecin and etoposide, while no particular sensitivity to UV, ionizing radiation, hydroxyurea or aphidicolin was found. The cohesion defect of RBS cells and their hypersensitivity to DNA-damaging agents were not corrected by a patient-derived ESCO<span class="hlt">2</span> acetyltransferase mutant (W539<span class="hlt">G</span>), indicating that the acetyltransferase activity of ESCO<span class="hlt">2</span> is essential for its function. In contrast to a previous study on cells from patients with Cornelia de Lange syndrome, another cohesinopathy, RBS cells failed to exhibit excessive chromosome aberrations after irradiation in <span class="hlt">G</span><span class="hlt">2</span> phase of the cell cycle. Our results point at an S phase-specific role for ESCO<span class="hlt">2</span> in the maintenance of genome stability. PMID:19738907</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1469616','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1469616"><span>Cytogenetic Monitoring of Farmers exposed to pesticides in Colombia.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Hoyos, L S; Carvajal, S; Solano, L; Rodriguez, J; Orozco, L; López, Y; Au, W W</p> <p>1996-01-01</p> <p>We have monitored 30 pesticide-exposed workers and 30 matched controls for expression of chromosome aberrations (CA) and sister <span class="hlt">chromatid</span> exchanges (SCE) in their lymphocytes. Peripheral blood cultures were set up within 3 hr after the collection of samples, and four cultures were set up from each donor. For CA analysis, 100 complete metaphase cells from each donor were evaluated. For the SCE assay, 50 complete metaphase cells from each donor were analyzed. The CA and SCE data were analyzed for differences between the two groups using the chi <span class="hlt">2</span> and the Student's t-test, respectively. From the CA analysis it was obvious that the overwhelming majority of aberrations were <span class="hlt">chromatid</span> <span class="hlt">breaks</span> and isochromatid <span class="hlt">breaks</span>; therefore, only these data are presented and used for statistical analysis. Isochromatid <span class="hlt">breaks</span> were counted as two <span class="hlt">breaks</span> each and <span class="hlt">chromatid</span> <span class="hlt">breaks</span> as one in calculating the total <span class="hlt">chromatid</span> <span class="hlt">break</span> frequencies. Statistical evaluation of the data indicates that there is no significant difference (p > 0.05; chi <span class="hlt">2</span> test) between the exposed and the nonexposed groups based on <span class="hlt">chromatid</span> <span class="hlt">breaks</span> per 100 cells (1.<span class="hlt">2</span> +/- 0.3 and 1.5 +/- 0.<span class="hlt">2</span>, respectively) and total <span class="hlt">chromatid</span> <span class="hlt">breaks</span> per 100 cells (1.7 +/- 0.3 and <span class="hlt">2</span>.1 +/- 0.<span class="hlt">2</span>, respectively). No significantly difference between the two groups (p > 0.05, Student's t-test) was observed with SCE frequencies (5.0 +/- 1.1 and 4.8 +/- 0.9, respectively). Linear regression analysis indicates that the data were not influenced by age, cigarette smoking, or alcohol consumption. It is assuring that the exposure conditions among these Indian farmers have not caused detectable increases of chromosome damage using standard assays; this suggests the lack of serious long-term health problems. However, periodic monitoring of such exposed populations should be conducted using the same or other more sensitive assays. In addition, other populations with exposure to different types of pesticides in Colombia should also be investigated. PMID</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2969851','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2969851"><span>BubR1- and Polo-Coated DNA Tethers Facilitate Poleward Segregation of Acentric <span class="hlt">Chromatids</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Royou, Anne; Gagou, Mary E.; Karess, Roger; Sullivan, William</p> <p>2010-01-01</p> <p>Summary The mechanisms that safeguard cells against chromosomal instability (CIN) are of great interest, as CIN contributes to tumorigenesis. To gain insight into these mechanisms, we studied the behavior of cells entering mitosis with damaged chromosomes. We used the endonuclease I-CreI to generate acentric chromosomes in Drosophila larvae. While I-CreI expression produces acentric chromosomes in the majority of neuronal stem cells, remarkably, it has no effect on adult survival. Our live studies reveal that acentric <span class="hlt">chromatids</span> segregate efficiently to opposite poles. The acentric <span class="hlt">chromatid</span> poleward movement is mediated through DNA tethers decorated with BubR1, Polo, INCENP, and Aurora-B. Reduced BubR1 or Polo function results in abnormal segregation of acentric <span class="hlt">chromatids</span>, a decrease in acentric chromosome tethering, and a great reduction in adult survival. We propose that BubR1 and Polo facilitate the accurate segregation of acentric <span class="hlt">chromatids</span> by maintaining the integrity of the tethers that connect acentric chromosomes to their centric partners. PMID:20141837</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1241963','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1241963"><span>Magnetic-field-<span class="hlt">induced</span> DNA strand <span class="hlt">breaks</span> in brain cells of the rat.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Lai, Henry; Singh, Narendra P</p> <p>2004-01-01</p> <p>In previous research, we found that rats acutely (<span class="hlt">2</span> hr) exposed to a 60-Hz sinusoidal magnetic field at intensities of 0.1-0.5 millitesla (mT) showed increases in DNA single- and double-strand <span class="hlt">breaks</span> in their brain cells. Further research showed that these effects could be blocked by pretreating the rats with the free radical scavengers melatonin and N-tert-butyl-alpha-phenylnitrone, suggesting the involvement of free radicals. In the present study, effects of magnetic field exposure on brain cell DNA in the rat were further investigated. Exposure to a 60-Hz magnetic field at 0.01 mT for 24 hr caused a significant increase in DNA single- and double-strand <span class="hlt">breaks</span>. Prolonging the exposure to 48 hr caused a larger increase. This indicates that the effect is cumulative. In addition, treatment with Trolox (a vitamin E analog) or 7-nitroindazole (a nitric oxide synthase inhibitor) blocked magnetic-field-<span class="hlt">induced</span> DNA strand <span class="hlt">breaks</span>. These data further support a role of free radicals on the effects of magnetic fields. Treatment with the iron chelator deferiprone also blocked the effects of magnetic fields on brain cell DNA, suggesting the involvement of iron. Acute magnetic field exposure increased apoptosis and necrosis of brain cells in the rat. We hypothesize that exposure to a 60-Hz magnetic field initiates an iron-mediated process (e.<span class="hlt">g</span>., the Fenton reaction) that increases free radical formation in brain cells, leading to DNA strand <span class="hlt">breaks</span> and cell death. This hypothesis could have an important implication for the possible health effects associated with exposure to extremely low-frequency magnetic fields in the public and occupational environments. PMID:15121512</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28358361','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28358361"><span>Tumor-treating fields elicit a conditional vulnerability to ionizing radiation via the downregulation of BRCA1 signaling and reduced DNA double-strand <span class="hlt">break</span> repair capacity in non-small cell lung cancer cell lines.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Karanam, Narasimha Kumar; Srinivasan, Kalayarasan; Ding, Lianghao; Sishc, Brock; Saha, Debabrata; Story, Michael D</p> <p>2017-03-30</p> <p>The use of tumor-treating fields (TTFields) has revolutionized the treatment of recurrent and newly diagnosed glioblastoma (GBM). TTFields are low-intensity, intermediate frequency, alternating electric fields that are applied to tumor regions and cells using non-invasive arrays. The predominant mechanism by which TTFields are thought to kill tumor cells is the disruption of mitosis. Using five non-small cell lung cancer (NSCLC) cell lines we found that there is a variable response in cell proliferation and cell killing between these NSCLC cell lines that was independent of p53 status. TTFields treatment increased the <span class="hlt">G</span><span class="hlt">2</span>/M population, with a concomitant reduction in S-phase cells followed by the appearance of a sub-<span class="hlt">G</span>1 population indicative of apoptosis. Temporal changes in gene expression during TTFields exposure was evaluated to identify molecular signaling changes underlying the differential TTFields response. The most differentially expressed genes were associated with the cell cycle and cell proliferation pathways. However, the expression of genes found within the BRCA1 DNA-damage response were significantly downregulated (P<0.05) during TTFields treatment. DNA double-strand <span class="hlt">break</span> (DSB) repair foci increased when cells were exposed to TTFields as did the appearance of <span class="hlt">chromatid</span>-type aberrations, suggesting an interphase mechanism responsible for cell death involving DNA repair. Exposing cells to TTFields immediately following ionizing radiation resulted in increased <span class="hlt">chromatid</span> aberrations and a reduced capacity to repair DNA DSBs, which were likely responsible for at least a portion of the enhanced cell killing seen with the combination. These findings suggest that TTFields <span class="hlt">induce</span> a state of 'BRCAness' leading to a conditional susceptibility resulting in enhanced sensitivity to ionizing radiation and provides a strong rationale for the use of TTFields as a combined modality therapy with radiation or other DNA-damaging agents.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2230547','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2230547"><span>Homologous Recombination Repair Protects Against Particulate Chromate-<span class="hlt">induced</span> Chromosome Instability in Chinese Hamster Cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Stackpole, Megan M.; Wise, Sandra S.; Duzevik, Eliza Grlickova; Munroe, Ray C.; Thompson, W. Douglas; Thacker, John; Thompson, Larry H.; Hinz, John M.; Wise, John Pierce</p> <p>2008-01-01</p> <p>Particulate hexavalent chromium [Cr(VI)] compounds are well-established human carcinogens. Cr(VI)-<span class="hlt">induced</span> tumors are characterized by chromosomal instability (CIN); however, the mechanisms of this effect are unknown. We investigated the hypothesis that homologous recombination (HR) repair of DNA double strand <span class="hlt">breaks</span> protect cells from Cr(VI)-<span class="hlt">induced</span> CIN by focusing on the XRCC3 and RAD51C genes, which play an important role in cellular resistance to DNA double strand <span class="hlt">breaks</span>. We used Chinese hamster cells defective in each HR gene (irs3 for RAD51C and irs1SF for XRCC3) and compared with their wildtype parental and cDNA-complemented controls. We found that the intracellular Cr ion levels varied among the cell lines after particulate chromate treatment. Importantly, accounting for differences in Cr ion levels, we discovered that XRCC3 and RAD51C cells treated with lead chromate had increased cytotoxicity and chromosomal aberrations, relative to wild-type and cDNA-complimented cells. We also observed the emergence of high levels of <span class="hlt">chromatid</span> exchanges in the two mutant cell lines. For example, 1 ug/cm<span class="hlt">2</span> lead chromate <span class="hlt">induced</span> 20 and 32 exchanges in XRCC3- and RAD51C-deficient cells, respectively, whereas no exchanges were detected in the wildtype and cDNA-complemented cells. These observations suggest that HR protects cells from Cr(VI)-<span class="hlt">induced</span> CIN, consistent with the ability of particulate Cr(VI) to <span class="hlt">induce</span> double strand <span class="hlt">breaks</span>. PMID:17662313</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/5738888-enhanced-response-induction-sister-chromatid-exchange-gamma-radiation-neurofibromatosis','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/5738888-enhanced-response-induction-sister-chromatid-exchange-gamma-radiation-neurofibromatosis"><span>Enhanced response to the induction of sister <span class="hlt">chromatid</span> exchange by gamma radiation in neurofibromatosis</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Hafez, M.; Abd el-Nabi, S.M.; el-Wehedi, G.</p> <p></p> <p>The study included 8 unrelated patients with neurofibromatosis, and 10 unrelated normal and healthy persons as controls. Whole blood samples were divided into plastic T flasks and exposed at room temperature to gamma rays. The radiation dose was 36 rad/minute, and the doses delivered were 0, 75, 150 and 300 rad. The lymphocytes were cultured in (RPMI) 1640 tissue culture medium and autologous serum (20%). Phytohemagglutinin and bromodeoxyuridine (Brdu) (10 microM) were added at initiation of culture and harvesting was done 64 to 68 hours after culture initiation. Slides were coded, differential staining was done, and sister <span class="hlt">chromatid</span> exchanges (SCEs)more » and aberrations (gaps, <span class="hlt">breaks</span>, dicentrics, fragments and minutes) were counted. In the controls no significant increase in frequency of SCE has been found (P greater than 0.5). In the patients, the frequencies significantly increased with the increase of dose of irradiation (P less than 0.001). Furthermore, after irradiation, the incidence of gaps, <span class="hlt">breaks</span>, and dicentrics were significantly increased in patients compared with controls. Moreover, the incidence increased with the increase in the dose of radiation. The results are discussed with a conclusion that the results add to the indication of a genetic predisposition to develop cancer in neurofibromatosis patients.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22207796-vcc-over-expression-inhibits-cisplatin-induced-apoptosis-hepg2-cells','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22207796-vcc-over-expression-inhibits-cisplatin-induced-apoptosis-hepg2-cells"><span>VCC-1 over-expression inhibits cisplatin-<span class="hlt">induced</span> apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Zhou, Zhitao; Lu, Xiao; Zhu, Ping</p> <p></p> <p>Highlights: Black-Right-Pointing-Pointer VCC-1 is hypothesized to be associated with carcinogenesis. Black-Right-Pointing-Pointer Levels of VCC-1 are increased significantly in HCC. Black-Right-Pointing-Pointer Over-expression of VCC-1 could promotes cellular proliferation rate. Black-Right-Pointing-Pointer Over-expression of VCC-1 inhibit the cisplatin-provoked apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Black-Right-Pointing-Pointer VCC-1 plays an important role in control the tumor growth and apoptosis. -- Abstract: Vascular endothelial growth factor-correlated chemokine 1 (VCC-1), a recently described chemokine, is hypothesized to be associated with carcinogenesis. However, the molecular mechanisms by which aberrant VCC-1 expression determines poor outcomes of cancers are unknown. In this study, we found that VCC-1 was highly expressed in hepatocellularmore » carcinoma (HCC) tissue. It was also associated with proliferation of Hep<span class="hlt">G</span><span class="hlt">2</span> cells, and inhibition of cisplatin-<span class="hlt">induced</span> apoptosis of Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Conversely, down-regulation of VCC-1 in Hep<span class="hlt">G</span><span class="hlt">2</span> cells increased cisplatin-<span class="hlt">induced</span> apoptosis of Hep<span class="hlt">G</span><span class="hlt">2</span> cells. In summary, these results suggest that VCC-1 is involved in cisplatin-<span class="hlt">induced</span> apoptosis of Hep<span class="hlt">G</span><span class="hlt">2</span> cells, and also provides some evidence for VCC-1 as a potential cellular target for chemotherapy.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040172818&hterms=mitosis&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dmitosis','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040172818&hterms=mitosis&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dmitosis"><span>Rejoining and misrejoining of radiation-<span class="hlt">induced</span> chromatin <span class="hlt">breaks</span>. III. Hypertonic treatment</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Durante, M.; George, K.; Wu, H. L.; Yang, T. C.</p> <p>1998-01-01</p> <p>It has been shown that treatment in anisotonic medium modifies rejoining of radiation-<span class="hlt">induced</span> <span class="hlt">breaks</span> in interphase chromosomes. In previous work, we have demonstrated that formation of exchanges in human lymphocytes has a slow component (half-time of 1-<span class="hlt">2</span> h), but a fraction of exchanges are also observed in samples assayed soon after exposure. In this paper we studied the effect of hypertonic treatment on rejoining and misrejoining of radiation-<span class="hlt">induced</span> <span class="hlt">breaks</span> using fluorescence in situ hybridization of prematurely condensed chromosomes in human lymphocytes. Isolated lymphocytes were irradiated with 7 Gy gamma rays, fused to mitotic hamster cells and incubated in hypertonic solution (0.5 M NaCl) for the period normally allowed for interphase chromosome condensation to occur. The data from hypertonic treatment experiments indicate the presence of a class of interphase chromosome <span class="hlt">breaks</span> that rejoin and misrejoin very quickly (half-time of 5-6 min). The fast misrejoining of these lesions is considered to be responsible for the initial level of exchanges which we reported previously. No significant effect of hypertonic treatment on the yield of chromosome aberrations scored at the first postirradiation mitosis was detected.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4194320','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4194320"><span>High levels of BRC4 <span class="hlt">induced</span> by a Tet-On 3<span class="hlt">G</span> system suppress DNA repair and impair cell proliferation in vertebrate cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Abe, Takuya; Branzei, Dana</p> <p>2014-01-01</p> <p>Transient induction or suppression of target genes is useful to study the function of toxic or essential genes in cells. Here we apply a Tet-On 3<span class="hlt">G</span> system to DT40 lymphoma B cell lines, validating it for three different genes. Using this tool, we then show that overexpression of the chicken BRC4 repeat of the tumor suppressor BRCA<span class="hlt">2</span> impairs cell proliferation and <span class="hlt">induces</span> chromosomal <span class="hlt">breaks</span>. Mechanistically, high levels of BRC4 suppress double strand <span class="hlt">break-induced</span> homologous recombination, inhibit the formation of RAD51 recombination repair foci, reduce cellular resistance to DNA damaging agents and <span class="hlt">induce</span> a <span class="hlt">G</span><span class="hlt">2</span> damage checkpoint-mediated cell-cycle arrest. The above phenotypes are mediated by BRC4 capability to bind and inhibit RAD51. The toxicity associated with BRC4 overexpression is exacerbated by chemotherapeutic agents and reversed by RAD51 overexpression, but it is neither aggravated nor suppressed by a deficit in the non-homologous end-joining pathway of double strand <span class="hlt">break</span> repair. We further find that the endogenous BRCA<span class="hlt">2</span> mediates the cytotoxicity associated with BRC4 induction, thus underscoring the possibility that BRC4 or other domains of BRCA<span class="hlt">2</span> cooperate with ectopic BRC4 in regulating repair activities or mitotic cell division. In all, the results demonstrate the utility of the Tet-On 3<span class="hlt">G</span> system in DT40 research and underpin a model in which BRC4 role on cell proliferation and chromosome repair arises primarily from its suppressive role on RAD51 functions. PMID:25218467</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25218467','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25218467"><span>High levels of BRC4 <span class="hlt">induced</span> by a Tet-On 3<span class="hlt">G</span> system suppress DNA repair and impair cell proliferation in vertebrate cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Abe, Takuya; Branzei, Dana</p> <p>2014-10-01</p> <p>Transient induction or suppression of target genes is useful to study the function of toxic or essential genes in cells. Here we apply a Tet-On 3<span class="hlt">G</span> system to DT40 lymphoma B cell lines, validating it for three different genes. Using this tool, we then show that overexpression of the chicken BRC4 repeat of the tumor suppressor BRCA<span class="hlt">2</span> impairs cell proliferation and <span class="hlt">induces</span> chromosomal <span class="hlt">breaks</span>. Mechanistically, high levels of BRC4 suppress double strand <span class="hlt">break-induced</span> homologous recombination, inhibit the formation of RAD51 recombination repair foci, reduce cellular resistance to DNA damaging agents and <span class="hlt">induce</span> a <span class="hlt">G</span><span class="hlt">2</span> damage checkpoint-mediated cell-cycle arrest. The above phenotypes are mediated by BRC4 capability to bind and inhibit RAD51. The toxicity associated with BRC4 overexpression is exacerbated by chemotherapeutic agents and reversed by RAD51 overexpression, but it is neither aggravated nor suppressed by a deficit in the non-homologous end-joining pathway of double strand <span class="hlt">break</span> repair. We further find that the endogenous BRCA<span class="hlt">2</span> mediates the cytotoxicity associated with BRC4 induction, thus underscoring the possibility that BRC4 or other domains of BRCA<span class="hlt">2</span> cooperate with ectopic BRC4 in regulating repair activities or mitotic cell division. In all, the results demonstrate the utility of the Tet-On 3<span class="hlt">G</span> system in DT40 research and underpin a model in which BRC4 role on cell proliferation and chromosome repair arises primarily from its suppressive role on RAD51 functions. Copyright © 2014 The Authors. Published by Elsevier B.V. All rights reserved.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li class="active"><span>5</span></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_5 --> <div id="page_6" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li class="active"><span>6</span></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="101"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26036576','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26036576"><span>Parkin <span class="hlt">induces</span> <span class="hlt">G</span><span class="hlt">2</span>/M cell cycle arrest in TNF-α-treated HeLa cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lee, Min Ho; Cho, Yoonjung; Jung, Byung Chul; Kim, Sung Hoon; Kang, Yeo Wool; Pan, Cheol-Ho; Rhee, Ki-Jong; Kim, Yoon Suk</p> <p>2015-08-14</p> <p>Parkin is a known tumor suppressor. However, the mechanism by which parkin acts as a tumor suppressor remains to be fully elucidated. Previously, we reported that parkin expression <span class="hlt">induces</span> caspase-dependent apoptotic cell death in TNF-α-treated HeLa cells. However, at that time, we did not consider the involvement of parkin in cell cycle control. In the current study, we investigated whether parkin is involved in cell cycle regulation and suppression of cancer cell growth. In our cell cycle analyses, parkin expression <span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span>/M cell cycle arrest in TNF-α-treated HeLa cells. To elucidate the mechanism(s) by which parkin <span class="hlt">induces</span> this <span class="hlt">G</span><span class="hlt">2</span>/M arrest, we analyzed cell cycle regulatory molecules involved in the <span class="hlt">G</span><span class="hlt">2</span>/M transition. Parkin expression <span class="hlt">induced</span> CDC<span class="hlt">2</span> phosphorylation which is known to inhibit CDC<span class="hlt">2</span> activity and cause <span class="hlt">G</span><span class="hlt">2</span>/M arrest. Cyclin B1, which is degraded during the mitotic transition, accumulated in response to parkin expression, thereby indicating parkin-<span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span>/M arrest. Next, we established that Myt1, which is known to phosphorylate and inhibit CDC<span class="hlt">2</span>, increased following parkin expression. In addition, we found that parkin also <span class="hlt">induces</span> increased Myt1 expression, <span class="hlt">G</span><span class="hlt">2</span>/M arrest, and reduced cell viability in TNF-α-treated HCT15 cells. Furthermore, knockdown of parkin expression by parkin-specific siRNA decreased Myt1 expression and phosphorylation of CDC<span class="hlt">2</span> and resulted in recovered cell viability. These results suggest that parkin acts as a crucial molecule causing cell cycle arrest in <span class="hlt">G</span><span class="hlt">2</span>/M, thereby suppressing tumor cell growth. Copyright © 2015 Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26492105','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26492105"><span>Mitochondrial Dysfunction and Ca(<span class="hlt">2</span>+) Overload Contributes to Hesperidin <span class="hlt">Induced</span> Paraptosis in Hepatoblastoma Cells, Hep<span class="hlt">G</span><span class="hlt">2</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yumnam, Silvia; Hong, Gyeong Eun; Raha, Suchismita; Saralamma, Venu Venkatarame Gowda; Lee, Ho Jeong; Lee, Won-Sup; Kim, Eun-Hee; Kim, Gon Sup</p> <p>2016-06-01</p> <p>Paraptosis is a programmed cell death which is morphologically and biochemically different from apoptosis. In this study, we have investigated the role of Ca(<span class="hlt">2</span>+) in hesperidin-<span class="hlt">induced</span> paraptotic cell death in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Increase in mitochondrial Ca(<span class="hlt">2</span>+) level was observed in hesperidin treated Hep<span class="hlt">G</span><span class="hlt">2</span> cells but not in normal liver cancer cells. Inhibition of inositol-1,4,5-triphosphate receptor (IP3 R) and ryanodine receptor also block the mitochondrial Ca(<span class="hlt">2</span>+) accumulation suggesting that the release of Ca(<span class="hlt">2</span>+) from the endoplasmic reticulum (ER) may probably lead to the increase in mitochondrial Ca(<span class="hlt">2</span>+) level. Pretreatment with ruthenium red (RuRed), a Ca(<span class="hlt">2</span>+) uniporter inhibitor inhibited the hesperidin-<span class="hlt">induced</span> mitochondrial Ca(<span class="hlt">2</span>+) overload, swelling of mitochondria, and cell death in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. It has also been demonstrated that mitochondrial Ca(<span class="hlt">2</span>+) influxes act upstream of ROS and mitochondrial superoxide production. The increased ROS production further leads to mitochondrial membrane loss in hesperidin treated Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Taken together our results show that IP3 R and ryanodine receptor mediated release of Ca(<span class="hlt">2</span>+) from the ER and its subsequent influx through the uniporter into mitochondria contributes to hesperidin-<span class="hlt">induced</span> paraptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. © 2015 Wiley Periodicals, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5052648','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5052648"><span>Humic acid inhibits HBV-<span class="hlt">induced</span> autophagosome formation and <span class="hlt">induces</span> apoptosis in HBV-transfected Hep <span class="hlt">G</span><span class="hlt">2</span> cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Pant, Kishor; Yadav, Ajay K.; Gupta, Parul; Rathore, Abhishek Singh; Nayak, Baibaswata; Venugopal, Senthil K.</p> <p>2016-01-01</p> <p>Hepatitis B Virus (HBV) utilizes several mechanisms to survive in the host cells and one of the main pathways being autophagosome formation. Humic acid (HA), one of the major components of Mineral pitch, is an Ayurvedic medicinal food, commonly used by the people of the Himalayan regions of Nepal and India for various body ailments. We hypothesized that HA could <span class="hlt">induce</span> cell death and inhibit HBV-<span class="hlt">induced</span> autophagy in hepatic cells. Incubation of Hep <span class="hlt">G</span><span class="hlt">2.2</span>.1.5 cells (Hep<span class="hlt">G</span><span class="hlt">2</span> cells stably expressing HBV) with HA (100 μM) inhibited both cell proliferation and autophagosome formation significantly, while apoptosis induction was enhanced. Western blot results showed that HA incubation resulted in decreased levels of beclin-1, SIRT-1 and c-myc, while caspase-3 and β-catenin expression were up-regulated. Western blot results showed that HA significantly inhibited the expression of HBx (3-fold with 50 μM and 5-fold with 100 μM) compared to control cells. When HA was incubated with HBx-transfected Hep <span class="hlt">G</span><span class="hlt">2</span> cells, HBx-<span class="hlt">induced</span> autophagosome formation and beclin-1 levels were decreased. These data showed that HA <span class="hlt">induced</span> apoptosis and inhibited HBV-<span class="hlt">induced</span> autophagosome formation and proliferation in hepatoma cells. PMID:27708347</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1128680','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1128680"><span>DNA single strand breakage, DNA adducts, and sister <span class="hlt">chromatid</span> exchange in lymphocytes and phenanthrene and pyrene metabolites in urine of coke oven workers.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Popp, W; Vahrenholz, C; Schell, C; Grimmer, G; Dettbarn, G; Kraus, R; Brauksiepe, A; Schmeling, B; Gutzeit, T; von Bülow, J; Norpoth, K</p> <p>1997-01-01</p> <p>OBJECTIVES: To investigate the specificity of biological monitoring variables (excretion of phenanthrene and pyrene metabolites in urine) and the usefulness of some biomarkers of effect (alkaline filter elution, 32P postlabelling assay, measurement of sister <span class="hlt">chromatid</span> exchange) in workers exposed to polycyclic aromatic hydrocarbons (PAHs). METHODS: 29 coke oven workers and a standardised control group were investigated for frequencies of DNA single strand breakage, DNA protein cross links (alkaline filter elution assay), sister <span class="hlt">chromatid</span> exchange, and DNA adducts (32P postlabelling assay) in lymphocytes. Phenanthrene and pyrene metabolites were measured in 24 hour urine samples. 19 different PAHs (including benzo(a)pyrene, pyrene, and phenanthrene) were measured at the workplace by personal air monitoring. The GSTT1 activity in erythrocytes and lymphocyte subpopulations in blood was also measured. RESULTS: Concentrations of phenanthrene, pyrene, and benzo(a)pyrene in air correlated well with the concentration of total PAHs in air; they could be used for comparisons of different workplaces if the emission compositions were known. The measurement of phenanthrene metabolites in urine proved to be a better biological monitoring variable than the measurement of 1-hydroxypyrene. Significantly more DNA strand <span class="hlt">breaks</span> in lymphocytes of coke oven workers were found (alkaline filter elution assay); the DNA adduct rate was not significantly increased in workers, but correlated with exposure to PAHs in a semiquantitative manner. The number of sister <span class="hlt">chromatid</span> exchanges was lower in coke oven workers but this was not significant; thus counting sister <span class="hlt">chromatid</span> exchanges was not a good variable for biomonitoring of coke oven workers. Also, indications for immunotoxic influences (changes in lymphocyte subpopulations) were found. CONCLUSIONS: The measurement of phenanthrene metabolites in urine seems to be a better biological monitoring variable for exposure to PAHs than</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016NHESS..16.1629T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016NHESS..16.1629T"><span>Mangrove forest against dyke-<span class="hlt">break-induced</span> tsunami on rapidly subsiding coasts</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Takagi, Hiroshi; Mikami, Takahito; Fujii, Daisuke; Esteban, Miguel; Kurobe, Shota</p> <p>2016-07-01</p> <p>Thin coastal dykes typically found in developing countries may suddenly collapse due to rapid land subsidence, material ageing, sea-level rise, high wave attack, earthquakes, landslides, or a collision with vessels. Such a failure could trigger dam-<span class="hlt">break</span> tsunami-type flooding, or "dyke-<span class="hlt">break-induced</span> tsunami", a possibility which has so far been overlooked in the field of coastal disaster science and management. To analyse the potential consequences of one such flooding event caused by a dyke failure, a hydrodynamic model was constructed based on the authors' field surveys of a vulnerable coastal location in Jakarta, Indonesia. In a <span class="hlt">2</span> m land subsidence scenario - which is expected to take place in the study area after only about 10-20 years - the model results show that the floodwaters rapidly rise to a height of nearly 3 m, resembling the flooding pattern of earthquake-<span class="hlt">induced</span> tsunamis. The depth-velocity product criterion suggests that many of the narrow pedestrian paths behind the dyke could experience strong flows, which are far greater than the safe limits that would allow pedestrian evacuation. A couple of alternative scenarios were also considered to investigate how such flood impacts could be mitigated by creating a mangrove belt in front of the dyke as an additional safety measure. The dyke-<span class="hlt">break-induced</span> tsunamis, which in many areas are far more likely than regular earthquake tsunamis, cannot be overlooked and thus should be considered in disaster management and urban planning along the coasts of many developing countries.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4199498','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4199498"><span>The Chromosomal Association of the Smc5/6 Complex Depends on Cohesion and Predicts the Level of Sister <span class="hlt">Chromatid</span> Entanglement</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Jeppsson, Kristian; Carlborg, Kristian K.; Nakato, Ryuichiro; Berta, Davide G.; Lilienthal, Ingrid; Kanno, Takaharu; Lindqvist, Arne; Brink, Maartje C.; Dantuma, Nico P.; Katou, Yuki; Shirahige, Katsuhiko; Sjögren, Camilla</p> <p>2014-01-01</p> <p>The cohesin complex, which is essential for sister <span class="hlt">chromatid</span> cohesion and chromosome segregation, also inhibits resolution of sister <span class="hlt">chromatid</span> intertwinings (SCIs) by the topoisomerase Top<span class="hlt">2</span>. The cohesin-related Smc5/6 complex (Smc5/6) instead accumulates on chromosomes after Top<span class="hlt">2</span> inactivation, known to lead to a buildup of unresolved SCIs. This suggests that cohesin can influence the chromosomal association of Smc5/6 via its role in SCI protection. Using high-resolution ChIP-sequencing, we show that the localization of budding yeast Smc5/6 to duplicated chromosomes indeed depends on sister <span class="hlt">chromatid</span> cohesion in wild-type and top<span class="hlt">2</span>-4 cells. Smc5/6 is found to be enriched at cohesin binding sites in the centromere-proximal regions in both cell types, but also along chromosome arms when replication has occurred under Top<span class="hlt">2</span>-inhibiting conditions. Reactivation of Top<span class="hlt">2</span> after replication causes Smc5/6 to dissociate from chromosome arms, supporting the assumption that Smc5/6 associates with a Top<span class="hlt">2</span> substrate. It is also demonstrated that the amount of Smc5/6 on chromosomes positively correlates with the level of missegregation in top<span class="hlt">2</span>-4, and that Smc5/6 promotes segregation of short chromosomes in the mutant. Altogether, this shows that the chromosomal localization of Smc5/6 predicts the presence of the <span class="hlt">chromatid</span> segregation-inhibiting entities which accumulate in top<span class="hlt">2</span>-4 mutated cells. These are most likely SCIs, and our results thus indicate that, at least when Top<span class="hlt">2</span> is inhibited, Smc5/6 facilitates their resolution. PMID:25329383</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4699751','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4699751"><span>Anti-cancer Activity of Osmanthus matsumuranus Extract by <span class="hlt">Inducing</span> <span class="hlt">G</span><span class="hlt">2</span>/M Arrest and Apoptosis in Human Hepatocellular Carcinoma Hep <span class="hlt">G</span><span class="hlt">2</span> Cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Jin, Soojung; Park, Hyun-Jin; Oh, You Na; Kwon, Hyun Ju; Kim, Jeong-Hwan; Choi, Yung Hyun; Kim, Byung Woo</p> <p>2015-01-01</p> <p>Background: Osmanthus matsumuranus, a species of Oleaceae, is found in East Asia and Southeast Asia. The bioactivities of O. matsumuranus have not yet been fully understood. Here, we studied on the molecular mechanisms underlying anti-cancer effect of ethanol extract of O. matsumuranus (EEOM). Methods: Inhibitory effect of EEOM on cell growth and proliferation was determined by WST assay in various cancer cells. To investigate the mechanisms of EEOM-mediated cytotoxicity, Hep<span class="hlt">G</span><span class="hlt">2</span> cells were treated with various concentration of EEOM and analyzed the cell cycle arrest and apoptosis induction by flow cytometry, Western blot analysis, 4,6-diamidino-<span class="hlt">2</span>-phenylindole (DAPI) staining and DNA fragmentation. Results: EEOM showed the cytotoxic activities in a dose-dependent manner in various cancer cell lines but not in normal cells, and Hep<span class="hlt">G</span><span class="hlt">2</span> cells were most susceptible to EEOM-<span class="hlt">induced</span> cytotoxicity. EEOM <span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span>/M arrest in Hep<span class="hlt">G</span><span class="hlt">2</span> cells associated with decreased expression of cyclin-dependent kinase 1 (CDK1), cyclin A and cylcin B, and increased expression of phospho-checkpoint kinase <span class="hlt">2</span>, p53 and CDK inhibitor p21. Immunofluorescence staining showed that EEOM-treated Hep<span class="hlt">G</span><span class="hlt">2</span> increased doublet nuclei and condensed actin, resulting in cell rounding. Furthermore, EEOM-mediated apoptosis was determined by Annexin V staining, chromatin condensation and DNA fragmentation. EEOM caused upregulation of FAS and Bax, activation of caspase-3, -8, -9, and fragmentation of poly ADP ribose polymerase. Conclusions: These results suggest that EEOM efficiently inhibits proliferation of Hep<span class="hlt">G</span><span class="hlt">2</span> cells by <span class="hlt">inducing</span> both <span class="hlt">G</span><span class="hlt">2</span>/M arrest and apoptosis via intrinsic and extrinsic pathways, and EEOM may be used as a cancer chemopreventive agent in the food or nutraceutical industry. PMID:26734586</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26784746','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26784746"><span>Chromosome Bridges Maintain Kinetochore-Microtubule Attachment throughout Mitosis and Rarely <span class="hlt">Break</span> during Anaphase.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pampalona, Judit; Roscioli, Emanuele; Silkworth, William T; Bowden, Brent; Genescà, Anna; Tusell, Laura; Cimini, Daniela</p> <p>2016-01-01</p> <p>Accurate chromosome segregation during cell division is essential to maintain genome stability, and chromosome segregation errors are causally linked to genetic disorders and cancer. An anaphase chromosome bridge is a particular chromosome segregation error observed in cells that enter mitosis with fused chromosomes/sister <span class="hlt">chromatids</span>. The widely accepted Breakage/Fusion/Bridge cycle model proposes that anaphase chromosome bridges <span class="hlt">break</span> during mitosis to generate chromosome ends that will fuse during the following cell cycle, thus forming new bridges that will <span class="hlt">break</span>, and so on. However, various studies have also shown a link between chromosome bridges and aneuploidy and/or polyploidy. In this study, we investigated the behavior and properties of chromosome bridges during mitosis, with the idea to gain insight into the potential mechanism underlying chromosome bridge-<span class="hlt">induced</span> aneuploidy. We find that only a small number of chromosome bridges <span class="hlt">break</span> during anaphase, whereas the rest persist through mitosis into the subsequent cell cycle. We also find that the microtubule bundles (k-fibers) bound to bridge kinetochores are not prone to breakage/detachment, thus supporting the conclusion that k-fiber detachment is not the cause of chromosome bridge-<span class="hlt">induced</span> aneuploidy. Instead, our data suggest that while the microtubules bound to the kinetochores of normally segregating chromosomes shorten substantially during anaphase, the k-fibers bound to bridge kinetochores shorten only slightly, and may even lengthen, during anaphase. This causes some of the bridge kinetochores/chromosomes to lag behind in a position that is proximal to the cell/spindle equator and may cause the bridged chromosomes to be segregated into the same daughter nucleus or to form a micronucleus.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27388656','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27388656"><span>Determination of chloride in brazilian crude oils by ion chromatography after extraction <span class="hlt">induced</span> by emulsion <span class="hlt">breaking</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Robaina, Nicolle F; Feiteira, Fernanda N; Cassella, Alessandra R; Cassella, Ricardo J</p> <p>2016-08-05</p> <p>The present paper reports on the development of a novel extraction <span class="hlt">induced</span> by emulsion <span class="hlt">breaking</span> (EIEB) method for the determination of chloride in crude oils. The proposed method was based on the formation and <span class="hlt">breaking</span> of oil-in-water emulsions with the samples and the consequential transference of the highly water-soluble chloride to the aqueous phase during emulsion <span class="hlt">breaking</span>, which was achieved by centrifugation. The determination of chloride in the extracts was performed by ion chromatography (IC) with conductivity detection. Several parameters (oil phase:aqueous phase ratio, crude oil:mineral oil ratio, shaking time and type and concentration of surfactant) that could affect the performance of the method were evaluated. Total extraction of chloride from samples could be achieved when 1.0<span class="hlt">g</span> of oil phase (0.5<span class="hlt">g</span> of sample+0.5<span class="hlt">g</span> of mineral oil) was emulsified in 5mL of a <span class="hlt">2</span>.5% (m/v) solution of Triton X-114. The obtained emulsion was shaken for 60min and broken by centrifugation for 5min at 5000rpm. The separated aqueous phase was collected, filtered and diluted before analysis by IC. Under these conditions, the limit of detection was 0.5μgg(-1) NaCl and the limit of quantification was 1.6μgg(-1) NaCl. We applied the method to the determination of chloride in six Brazilian crude oils and the results did not differ statistically from those obtained by the ASTM D6470 method when the paired Student-t-test, at 95% confidence level, was applied. Copyright © 2016 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014JGRC..119.6952Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014JGRC..119.6952Z"><span>A numerical investigation of wave-<span class="hlt">breaking-induced</span> turbulent coherent structure under a solitary wave</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhou, Zheyu; Sangermano, Jacob; Hsu, Tian-Jian; Ting, Francis C. K.</p> <p>2014-10-01</p> <p>To better understand the effect of wave-<span class="hlt">breaking-induced</span> turbulence on the bed, we report a 3-D large-eddy simulation (LES) study of a <span class="hlt">breaking</span> solitary wave in spilling condition. Using a turbulence-resolving approach, we study the generation and the fate of wave-<span class="hlt">breaking-induced</span> turbulent coherent structures, commonly known as obliquely descending eddies (ODEs). Specifically, we focus on how these eddies may impinge onto bed. The numerical model is implemented using an open-source CFD library of solvers, called OpenFOAM, where the incompressible 3-D filtered Navier-Stokes equations for the water and the air phases are solved with a finite volume scheme. The evolution of the water-air interfaces is approximated with a volume of fluid method. Using the dynamic Smagorinsky closure, the numerical model has been validated with wave flume experiments of solitary wave <span class="hlt">breaking</span> over a 1/50 sloping beach. Simulation results show that during the initial overturning of the <span class="hlt">breaking</span> wave, <span class="hlt">2</span>-D horizontal rollers are generated, accelerated, and further evolve into a couple of 3-D hairpin vortices. Some of these vortices are sufficiently intense to impinge onto the bed. These hairpin vortices possess counter-rotating and downburst features, which are key characteristics of ODEs observed by earlier laboratory studies using Particle Image Velocimetry. Model results also suggest that those ODEs that impinge onto bed can <span class="hlt">induce</span> strong near-bed turbulence and bottom stress. The intensity and locations of these near-bed turbulent events could not be parameterized by near-surface (or depth integrated) turbulence unless in very shallow depth.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1289920','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1289920"><span>Tadpole-<span class="hlt">induced</span> electroweak symmetry <span class="hlt">breaking</span> and pNGB Higgs models</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Harnik, Roni; Howe, Kiel; Kearney, John</p> <p></p> <p>We investigate <span class="hlt">induced</span> electroweak symmetry <span class="hlt">breaking</span> (EWSB) in models in which the Higgs is a pseudo-Nambu-Goldstone boson (pNGB). In pNGB Higgs models, Higgs properties and precision electroweak measurements imply a hierarchy between the EWSB and global symmetry-<span class="hlt">breaking</span> scales,more » $$v_H \\ll f_H$$. When the pNGB potential is generated radiatively, this hierarchy requires fine-tuning to a degree of at least $$\\sim v_H^<span class="hlt">2</span>/f_H^<span class="hlt">2</span>$$. We show that if Higgs EWSB is <span class="hlt">induced</span> by a tadpole arising from an auxiliary sector at scale $$f_\\Sigma \\ll v_H$$, this tuning is significantly ameliorated or can even be removed. We present explicit examples both in Twin Higgs models and in Composite Higgs models based on $SO(5)/SO(4)$. For the Twin case, the result is a fully natural model with $$f_H \\sim 1$$ TeV and the lightest colored top partners at <span class="hlt">2</span> TeV. These models also have an appealing mechanism to generate the scales of the auxiliary sector and Higgs EWSB directly from the scale $$f_H$$, with a natural hierarchy $$f_\\Sigma \\ll v_H \\ll f_H \\sim{\\rm TeV}$$. Finally, the framework predicts modified Higgs coupling as well as new Higgs and vector states at LHC13.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1289920-tadpole-induced-electroweak-symmetry-breaking-pngb-higgs-models','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1289920-tadpole-induced-electroweak-symmetry-breaking-pngb-higgs-models"><span>Tadpole-<span class="hlt">induced</span> electroweak symmetry <span class="hlt">breaking</span> and pNGB Higgs models</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Harnik, Roni; Howe, Kiel; Kearney, John</p> <p>2017-03-22</p> <p>We investigate <span class="hlt">induced</span> electroweak symmetry <span class="hlt">breaking</span> (EWSB) in models in which the Higgs is a pseudo-Nambu-Goldstone boson (pNGB). In pNGB Higgs models, Higgs properties and precision electroweak measurements imply a hierarchy between the EWSB and global symmetry-<span class="hlt">breaking</span> scales,more » $$v_H \\ll f_H$$. When the pNGB potential is generated radiatively, this hierarchy requires fine-tuning to a degree of at least $$\\sim v_H^<span class="hlt">2</span>/f_H^<span class="hlt">2</span>$$. We show that if Higgs EWSB is <span class="hlt">induced</span> by a tadpole arising from an auxiliary sector at scale $$f_\\Sigma \\ll v_H$$, this tuning is significantly ameliorated or can even be removed. We present explicit examples both in Twin Higgs models and in Composite Higgs models based on $SO(5)/SO(4)$. For the Twin case, the result is a fully natural model with $$f_H \\sim 1$$ TeV and the lightest colored top partners at <span class="hlt">2</span> TeV. These models also have an appealing mechanism to generate the scales of the auxiliary sector and Higgs EWSB directly from the scale $$f_H$$, with a natural hierarchy $$f_\\Sigma \\ll v_H \\ll f_H \\sim{\\rm TeV}$$. Finally, the framework predicts modified Higgs coupling as well as new Higgs and vector states at LHC13.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2847229','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2847229"><span>Phosphorylation of Exo1 modulates homologous recombination repair of DNA double-strand <span class="hlt">breaks</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Bolderson, Emma; Tomimatsu, Nozomi; Richard, Derek J.; Boucher, Didier; Kumar, Rakesh; Pandita, Tej K.; Burma, Sandeep; Khanna, Kum Kum</p> <p>2010-01-01</p> <p>DNA double-strand <span class="hlt">break</span> (DSB) repair via the homologous recombination pathway is a multi-stage process, which results in repair of the DSB without loss of genetic information or fidelity. One essential step in this process is the generation of extended single-stranded DNA (ssDNA) regions at the <span class="hlt">break</span> site. This ssDNA serves to <span class="hlt">induce</span> cell cycle checkpoints and is required for Rad51 mediated strand invasion of the sister <span class="hlt">chromatid</span>. Here, we show that human Exonuclease 1 (Exo1) is required for the normal repair of DSBs by HR. Cells depleted of Exo1 show chromosomal instability and hypersensitivity to ionising radiation (IR) exposure. We find that Exo1 accumulates rapidly at DSBs and is required for the recruitment of RPA and Rad51 to sites of DSBs, suggesting a role for Exo1 in ssDNA generation. Interestingly, the phosphorylation of Exo1 by ATM appears to regulate the activity of Exo1 following resection, allowing optimal Rad51 loading and the completion of HR repair. These data establish a role for Exo1 in resection of DSBs in human cells, highlighting the critical requirement of Exo1 for DSB repair via HR and thus the maintenance of genomic stability. PMID:20019063</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24236568','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24236568"><span>Osthole <span class="hlt">induces</span> <span class="hlt">G</span><span class="hlt">2</span>/M cell cycle arrest and apoptosis in human hepatocellular carcinoma Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chao, Xu; Zhou, Xiaojun; Zheng, Gang; Dong, Changhu; Zhang, Wei; Song, Xiaomei; Jin, Tianbo</p> <p>2014-05-01</p> <p>Osthole [7-methoxy-8-(3-methyl-<span class="hlt">2</span>-butenyl) coumarin] isolated from the fruit of Cnidium monnieri (L.) Cuss, one of the commonly used Chinese medicines listed in the Shennong's Classic of Materia Medica in the Han Dynasty, had remarkable antiproliferative activity against human hepatocellular carcinoma Hep<span class="hlt">G</span><span class="hlt">2</span> cells in culture. This study evaluated the effects of osthole on cell growth, nuclear morphology, cell cycle distribution, and expression of apoptosis-related proteins in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Cytotoxic activity of osthole was determined by the MTT assay at various concentrations ranging from 0.004 to 1.0 µmol/ml in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Cell morphology was assessed by Hoechst staining and fluorescence microscopy. Apoptosis and cell-cycle distribution was determined by annexin V staining and flow cytometry. Apoptotic protein levels were assessed by Western blot. Osthole exhibited significant inhibition of the survival of Hep<span class="hlt">G</span><span class="hlt">2</span> cells and the half inhibitory concentration (IC₅₀) values were 0.186, 0.158 and 0.123 µmol/ml at 24, 48 and 72 h, respectively. Cells treated with osthole at concentrations of 0, 0.004, 0.02, 0.1 and 0.5 μmol/ml showed a statistically significant increase in the <span class="hlt">G</span><span class="hlt">2</span>/M fraction accompanied by a decrease in the <span class="hlt">G</span>0/<span class="hlt">G</span>1 fraction. The increase of apoptosis <span class="hlt">induced</span> by osthole was correlated with down-regulation expression of anti-apoptotic Bcl-<span class="hlt">2</span> protein and up-regulation expression of pro-apoptotic Bax and p53 proteins. Osthole had significant growth inhibitory activity and the pro-apoptotic effect of osthole is mediated through the activation of caspases and mitochondria in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Results suggest that osthole has promising therapeutic potential against hepatocellular carcinoma.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26616367','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26616367"><span>The Nitric Oxide Prodrug JS-K <span class="hlt">Induces</span> Ca(<span class="hlt">2</span>+)-Mediated Apoptosis in Human Hepatocellular Carcinoma Hep<span class="hlt">G</span><span class="hlt">2</span> Cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liu, Ling; Wang, Dongmei; Wang, Jiangang; Wang, Shuying</p> <p>2016-04-01</p> <p>Hepatocellular carcinoma is one of the most common and deadly forms of human malignancies. JS-K, O(<span class="hlt">2)-(2</span>, 4-dinitrophenyl) 1-[(4-ethoxycarbonyl) piperazin-1-yl] diazen-1-ium-1, <span class="hlt">2</span>-diolate, has the ability to <span class="hlt">induce</span> apoptosis of tumor cell lines. In the present study, JS-K inhibited the proliferation of Hep<span class="hlt">G</span><span class="hlt">2</span> cells in a time- and concentration-dependent manner and significantly <span class="hlt">induced</span> apoptosis. JS-K enhanced the ratio of Bax-to-Bcl-<span class="hlt">2</span>, released of cytochrome c (Cyt c) from mitochondria and the activated caspase-9/3. JS-K caused an increasing cytosolic Ca(<span class="hlt">2</span>+) and the loss of mitochondrial membrane potential. Carboxy-PTIO (a NO scavenger) and BAPTA-AM (an intracellular Ca(<span class="hlt">2</span>+) chelator) significantly blocked an increasing cytosolic Ca(<span class="hlt">2</span>+) in JS-K-<span class="hlt">induced</span> Hep<span class="hlt">G</span><span class="hlt">2</span> cells apoptosis, especially Carboxy-PTIO. Meanwhile, Carboxy-PTIO and BAPTA-AM treatment both attenuate JS-K-<span class="hlt">induced</span> apoptosis through upregulation of Bcl-<span class="hlt">2</span>, downregulation of Bax, reduction of Cyt c release from mitochondria to cytoplasm and inactivation of caspase-9/3. In summary, JS-K <span class="hlt">induced</span> Hep<span class="hlt">G</span><span class="hlt">2</span> cells apoptosis via Ca(<span class="hlt">2</span>+)/caspase-3-mediated mitochondrial pathway. © 2015 Wiley Periodicals, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27049592','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27049592"><span>Heteromerization of <span class="hlt">G</span><span class="hlt">2</span>A and OGR1 enhances proton sensitivity and proton-<span class="hlt">induced</span> calcium signals.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Huang, Ya-Han; Su, Yeu-Shiuan; Chang, Chung-Jen; Sun, Wei-Hsin</p> <p>2016-12-01</p> <p>Proton-sensing <span class="hlt">G</span>-protein-coupled receptors (GPCRs; OGR1, GPR4, <span class="hlt">G</span><span class="hlt">2</span>A, TDAG8), with full activation at pH 6.4 ∼ 6.8, are important to pH homeostasis, immune responses and acid-<span class="hlt">induced</span> pain. Although <span class="hlt">G</span><span class="hlt">2</span>A mediates the <span class="hlt">G</span>13-Rho pathway in response to acid, whether <span class="hlt">G</span><span class="hlt">2</span>A activates Gs, Gi or Gq proteins remains debated. In this study, we examined the response of this fluorescence protein-tagged OGR1 family to acid stimulation in HEK293T cells. <span class="hlt">G</span><span class="hlt">2</span>A did not generate detectable intracellular calcium or cAMP signals or show apparent receptor redistribution with moderate acid (pH ≥ 6.0) stimulation but reduced cAMP accumulation under strong acid stimulation (pH ≤ 5.5). Surprisingly, coexpression of OGR1- and <span class="hlt">G</span><span class="hlt">2</span>A-enhanced proton sensitivity and proton-<span class="hlt">induced</span> calcium signals. This alteration is attributed to oligomerization of OGR1 and <span class="hlt">G</span><span class="hlt">2</span>A. The oligomeric potential locates receptors at a specific site, which leads to enhanced proton-<span class="hlt">induced</span> calcium signals through channels.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24127286','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24127286"><span>Silymarin suppresses the PGE<span class="hlt">2</span> -<span class="hlt">induced</span> cell migration through inhibition of EP<span class="hlt">2</span> activation; <span class="hlt">G</span> protein-dependent PKA-CREB and <span class="hlt">G</span> protein-independent Src-STAT3 signal pathways.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Woo, Seon Min; Min, Kyoung-Jin; Chae, In Gyeong; Chun, Kyung-Soo; Kwon, Taeg Kyu</p> <p>2015-03-01</p> <p>Silymarin has been known as a chemopreventive agent, and possesses multiple anti-cancer activities including induction of apoptosis, inhibition of proliferation and growth, and blockade of migration and invasion. However, whether silymarin could inhibit prostaglandin (PG) E<span class="hlt">2</span> -<span class="hlt">induced</span> renal cell carcinoma (RCC) migration and what are the underlying mechanisms are not well elucidated. Here, we found that silymarin markedly inhibited PGE<span class="hlt">2</span> -stimulated migration. PGE<span class="hlt">2</span> <span class="hlt">induced</span> <span class="hlt">G</span> protein-dependent CREB phosphorylation via protein kinase A (PKA) signaling, and PKA inhibitor (H89) inhibited PGE<span class="hlt">2</span> -mediated migration. Silymarin reduced PGE<span class="hlt">2</span> -<span class="hlt">induced</span> CREB phosphorylation and CRE-promoter activity. PGE<span class="hlt">2</span> also activated <span class="hlt">G</span> protien-independent signaling pathways (Src and STAT3) and silymarin reduced PGE<span class="hlt">2</span> -<span class="hlt">induced</span> phosphorylation of Src and STAT3. Inhibitor of Src (Saracatinib) markedly reduced PGE<span class="hlt">2</span> -mediated migration. We found that EP<span class="hlt">2</span>, a PGE<span class="hlt">2</span> receptor, is involved in PGE<span class="hlt">2</span> -mediated cell migration. Down regulation of EP<span class="hlt">2</span> by EP<span class="hlt">2</span> siRNA and EP<span class="hlt">2</span> antagonist (AH6809) reduced PGE<span class="hlt">2</span> -inudced migration. In contrast, EP<span class="hlt">2</span> agonist (Butaprost) increased cell migration and silymarin effectively reduced butaprost-mediated cell migration. Moreover, PGE<span class="hlt">2</span> increased EP<span class="hlt">2</span> expression through activation of positive feedback mechanism, and PGE<span class="hlt">2</span> -<span class="hlt">induced</span> EP<span class="hlt">2</span> expression, as well as basal EP<span class="hlt">2</span> levels, were reduced in silymarin-treated cells. Taken together, our study demonstrates that silymarin inhibited PGE<span class="hlt">2</span> -<span class="hlt">induced</span> cell migration through inhibition of EP<span class="hlt">2</span> signaling pathways (<span class="hlt">G</span> protein dependent PKA-CREB and <span class="hlt">G</span> protein-independent Src-STAT3). © 2013 Wiley Periodicals, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11531014','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11531014"><span>A comparative study on cytogenetic and antineoplastic effects <span class="hlt">induced</span> by two modified steroidal alkylating agents.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Papageorgiou, A; Tsavdaridis, D; Geromichalos, G D; Camoutsis, C; Karaberis, E; Mourelatos, D; Chrysogelou, E; Houvartas, S; Kotsis, A</p> <p>2001-01-01</p> <p>We investigated the effects of two newly synthesized steroidal derivatives of nitrogen mustard on sister <span class="hlt">chromatid</span> exchange rates and on human lymphocyte proliferation kinetics. The compound 33-hydroxy-5alpha,22alpha-spirostan- 12-one-p-(N,N-bis(<span class="hlt">2</span>-chloroethyl)amino)phenylacetate(1) was, on a molar basis, less effective in <span class="hlt">inducing</span> sister <span class="hlt">chromatid</span> exchange and suppressing cell proliferation rate indices than compound 3beta-hydroxy-12alpha-aza-C-homo-5alpha,22alpha-spirostan-12-one-p-(N,N-bis(<span class="hlt">2</span>-chloroethyl)amino)phenylacetate(<span class="hlt">2</span>). A correlation was observed between the magnitude of the sister <span class="hlt">chromatid</span> exchange response and the depression of cell proliferation index. We also studied the effects of the aforementioned compounds on Lewis lung carcinoma. The order of the percent inhibition of tumor growth achieved by the compounds coincides with the order of the cytogenetic effects they <span class="hlt">induce</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25211769','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25211769"><span>OVA-bound nanoparticles <span class="hlt">induce</span> OVA-specific Ig<span class="hlt">G</span>1, Ig<span class="hlt">G</span><span class="hlt">2</span>a, and Ig<span class="hlt">G</span><span class="hlt">2</span>b responses with low IgE synthesis.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yanase, Noriko; Toyota, Hiroko; Hata, Kikumi; Yagyu, Seina; Seki, Takahiro; Harada, Mitsunori; Kato, Yasuki; Mizuguchi, Junichiro</p> <p>2014-10-14</p> <p>There is an urgent requirement for a novel vaccine that can stimulate immune responses without unwanted toxicity, including IgE elevation. We examined whether antigen ovalbumin (OVA) conjugated to the surface of nanoparticles (NPs) (OVA-NPs) with average diameter of 110nm would serve as an immune adjuvant. When BALB/c mice were immunized with OVA-NPs, they developed sufficient levels of OVA-specific Ig<span class="hlt">G</span>1 antibody responses with low levels of IgE synthesis, representing helper T (Th)<span class="hlt">2</span>-mediated humoral immunity. OVA-specific Ig<span class="hlt">G</span><span class="hlt">2</span>a and Ig<span class="hlt">G</span><span class="hlt">2</span>b responses (i.e., Th1-mediated immunity) were also <span class="hlt">induced</span> by secondary immunization with OVA-NPs. As expected, immunization with OVA in alum (OVA-alum) stimulated humoral immune responses, including Ig<span class="hlt">G</span>1 and IgE antibodies, with only low levels of Ig<span class="hlt">G</span><span class="hlt">2</span>a/Ig<span class="hlt">G</span><span class="hlt">2</span>b antibodies. CD4-positive T cells from mice primed with OVA-NPs produced substantial levels of IL-21 and IL-4, comparable to those from OVA-alum group. The irradiated mice receiving OVA-NPs-primed B cells together with OVA-alum-primed T cells exhibited enhanced anti-OVA Ig<span class="hlt">G</span><span class="hlt">2</span>b responses relative to OVA-alum-primed B cells and T cells following stimulation with OVA-NPs. Moreover, when OVA-NPs-primed, but not OVA-alum-primed, B cells were cultured in the presence of anti-CD40 monoclonal antibody, IL-4, and IL-21, or LPS plus TGF-β in vitro, OVA-specific Ig<span class="hlt">G</span>1 or Ig<span class="hlt">G</span><span class="hlt">2</span>b antibody responses were elicited, suggesting that immunization with OVA-NPs modulates B cells to generate Ig<span class="hlt">G</span>1 and Ig<span class="hlt">G</span><span class="hlt">2</span>b responses. Thus, OVA-NPs might exert their adjuvant action on B cells, and they represent a promising potential vaccine for generating both Ig<span class="hlt">G</span>1 and Ig<span class="hlt">G</span><span class="hlt">2</span>a/Ig<span class="hlt">G</span><span class="hlt">2</span>b antibody responses with low IgE synthesis. Copyright © 2014 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=242769','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=242769"><span><span class="hlt">Breaks</span> <span class="hlt">induced</span> in the deoxyribonucleic acid of aerosolized Escherichia coli by ozonized cyclohexene.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>De Mik, G; De Groot, I</p> <p>1978-01-01</p> <p>The inactivation of aerosolized Escherichia coli by ozone, cyclohexene, and ozonized cyclohexene was studied. The parameters for damage were loss of reproduction and introduction of <span class="hlt">breaks</span> in the deoxyribonucleic acid (DNA). Aerosolization of E. coli in clean air at 80 percent relative humidity or in air containing either ozone or cyclohexene hardly affected survival; however, some <span class="hlt">breaks</span> per DNA molecule were <span class="hlt">induced</span>, as shown by sucrose gradient sedimentation of the DNA. Aerosolization of E. coli in air containing ozonized cyclohexene at 80 percent relative humidity decreased the survival by a factor of 10(3) or more after 1 h of exposure and <span class="hlt">induced</span> many <span class="hlt">breaks</span> in the DNA. PMID:341811</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li class="active"><span>6</span></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_6 --> <div id="page_7" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li class="active"><span>7</span></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="121"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29925947','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29925947"><span>Nuclear ARP<span class="hlt">2</span>/3 drives DNA <span class="hlt">break</span> clustering for homology-directed repair.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Schrank, Benjamin R; Aparicio, Tomas; Li, Yinyin; Chang, Wakam; Chait, Brian T; Gundersen, Gregg G; Gottesman, Max E; Gautier, Jean</p> <p>2018-06-20</p> <p>DNA double-strand <span class="hlt">breaks</span> repaired by non-homologous end joining display limited DNA end-processing and chromosomal mobility. By contrast, double-strand <span class="hlt">breaks</span> undergoing homology-directed repair exhibit extensive processing and enhanced motion. The molecular basis of this movement is unknown. Here, using Xenopus laevis cell-free extracts and mammalian cells, we establish that nuclear actin, WASP, and the actin-nucleating ARP<span class="hlt">2</span>/3 complex are recruited to damaged chromatin undergoing homology-directed repair. We demonstrate that nuclear actin polymerization is required for the migration of a subset of double-strand <span class="hlt">breaks</span> into discrete sub-nuclear clusters. Actin-driven movements specifically affect double-strand <span class="hlt">breaks</span> repaired by homology-directed repair in <span class="hlt">G</span><span class="hlt">2</span> cell cycle phase; inhibition of actin nucleation impairs DNA end-processing and homology-directed repair. By contrast, ARP<span class="hlt">2</span>/3 is not enriched at double-strand <span class="hlt">breaks</span> repaired by non-homologous end joining and does not regulate non-homologous end joining. Our findings establish that nuclear actin-based mobility shapes chromatin organization by generating repair domains that are essential for homology-directed repair in eukaryotic cells.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4532491','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4532491"><span>Nbs1 ChIP-Seq Identifies Off-Target DNA Double-Strand <span class="hlt">Breaks</span> <span class="hlt">Induced</span> by AID in Activated Splenic B Cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Linehan, Erin K.; Schrader, Carol E.; Stavnezer, Janet</p> <p>2015-01-01</p> <p>Activation-<span class="hlt">induced</span> cytidine deaminase (AID) is required for initiation of Ig class switch recombination (CSR) and somatic hypermutation (SHM) of antibody genes during immune responses. AID has also been shown to <span class="hlt">induce</span> chromosomal translocations, mutations, and DNA double-strand <span class="hlt">breaks</span> (DSBs) involving non-Ig genes in activated B cells. To determine what makes a DNA site a target for AID-<span class="hlt">induced</span> DSBs, we identify off-target DSBs <span class="hlt">induced</span> by AID by performing chromatin immunoprecipitation (ChIP) for Nbs1, a protein that binds DSBs, followed by deep sequencing (ChIP-Seq). We detect and characterize hundreds of off-target AID-dependent DSBs. Two types of tandem repeats are highly enriched within the Nbs1-binding sites: long CA repeats, which can form Z-DNA, and tandem pentamers containing the AID target hotspot WGCW. These tandem repeats are not nearly as enriched at AID-independent DSBs, which we also identified. Msh<span class="hlt">2</span>, a component of the mismatch repair pathway and important for genome stability, increases off-target DSBs, similar to its effect on Ig switch region DSBs, which are required intermediates during CSR. Most of the off-target DSBs are two-ended, consistent with generation during <span class="hlt">G</span>1 phase, similar to DSBs in Ig switch regions. However, a minority are one-ended, presumably due to conversion of single-strand <span class="hlt">breaks</span> to DSBs during replication. One-ended DSBs are repaired by processes involving homologous recombination, including <span class="hlt">break-induced</span> replication repair, which can lead to genome instability. Off-target DSBs, especially those present during S phase, can lead to chromosomal translocations, deletions and gene amplifications, resulting in the high frequency of B cell lymphomas derived from cells that express or have expressed AID. PMID:26263206</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24614013','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24614013"><span>Genistein protects hematopoietic stem cells against <span class="hlt">G-CSF-induced</span> DNA damage.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Souza, Liliana R; Silva, Erica; Calloway, Elissa; Kucuk, Omer; Rossi, Michael; McLemore, Morgan L</p> <p>2014-05-01</p> <p>Granulocyte colony-stimulating factor (<span class="hlt">G</span>-CSF) has been used to treat neutropenia in various clinical settings. Although clearly beneficial, there are concerns that the chronic use of <span class="hlt">G</span>-CSF in certain conditions increases the risk of myelodysplastic syndrome (MDS) and/or acute myeloid leukemia (AML). The most striking example is in severe congenital neutropenia (SCN). Patients with SCN develop MDS/AML at a high rate that is directly correlated to the cumulative lifetime dosage of <span class="hlt">G</span>-CSF. Myelodysplastic syndrome and AML that arise in these settings are commonly associated with chromosomal deletions. We have demonstrated in this study that chronic <span class="hlt">G</span>-CSF treatment in mice results in expansion of the hematopoietic stem cell (HSC) population. In addition, primitive hematopoietic progenitors from <span class="hlt">G</span>-CSF-treated mice show evidence of DNA damage as demonstrated by an increase in double-strand <span class="hlt">breaks</span> and recurrent chromosomal deletions. Concurrent treatment with genistein, a natural soy isoflavone, limits DNA damage in this population. The protective effect of genistein seems to be related to its preferential inhibition of <span class="hlt">G-CSF-induced</span> proliferation of HSCs. Importantly, genistein does not impair <span class="hlt">G-CSF-induced</span> proliferation of committed hematopoietic progenitors, nor diminishes neutrophil production. The protective effect of genistein was accomplished with plasma levels that are attainable through dietary supplementation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25529445','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25529445"><span>Eicosapentaenoic acid (EPA) <span class="hlt">induced</span> apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells through ROS-Ca(<span class="hlt">2</span>+)-JNK mitochondrial pathways.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Yuanyuan; Han, Lirong; Qi, Wentao; Cheng, Dai; Ma, Xiaolei; Hou, Lihua; Cao, Xiaohong; Wang, Chunling</p> <p>2015-01-24</p> <p>Eicosapentaenoic acid (EPA), a well-known dietary n-3 PUFAS, has been considered to inhibit proliferation of tumor cells. However, the molecular mechanism related to EPA-<span class="hlt">induced</span> liver cancer cells apoptosis has not been reported. In this study, we investigated the effect of EPA on Hep<span class="hlt">G</span><span class="hlt">2</span> cells proliferation and apoptosis mechanism through mitochondrial pathways. EPA inhibited proliferation of Hep<span class="hlt">G</span><span class="hlt">2</span> cells in a dose-dependent manner and had no significant effect on the cell viability of humor normal liver L-02 cells. It was found that EPA initially evoked ROS formation, leading to [Ca(<span class="hlt">2</span>+)]c accumulation and the mitochondrial permeability transition pore (MPTP) opening; EPA-<span class="hlt">induced</span> Hep<span class="hlt">G</span><span class="hlt">2</span> cells apoptosis was inhibited by N-acetylcysteine (NAC, an inhibitor of ROS), 1,<span class="hlt">2</span>-bis (<span class="hlt">2</span>-aminophenoxy) ethane-N,N,N',N'-tetraacetic acid (BAPTA-AM, a chelator of calcium) and CsA (inhibitor of MPTP). The relationship between ROS production, the increase of cytoplasmic Ca and MPTP opening was detected. It seems that ROS may act as an upstream regulator of EPA-<span class="hlt">induced</span> [Ca(<span class="hlt">2</span>+)]c generation, moreover, generation of ROS, overload of mitochondrial [Ca(<span class="hlt">2</span>+)]c, and JNK activated cause the opening of MPTP. Western blotting results showed that EPA elevated the phosphorylation status of JNK, processes associated with the ROS generation. Simultaneously, the apoptosis <span class="hlt">induced</span> by EPA was related to release of cytochrome C from mitochondria to cytoplasm through the MPTP and activation of caspase-9 and caspase-3. These results suggest that EPA <span class="hlt">induces</span> apoptosis through ROS-Ca(<span class="hlt">2</span>+)-JNK mitochondrial pathways. Copyright © 2014 Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22416907-eicosapentaenoic-acid-epa-induced-apoptosis-hepg2-cells-through-rosca-sup-jnk-mitochondrial-pathways','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22416907-eicosapentaenoic-acid-epa-induced-apoptosis-hepg2-cells-through-rosca-sup-jnk-mitochondrial-pathways"><span>Eicosapentaenoic acid (EPA) <span class="hlt">induced</span> apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells through ROS–Ca{sup <span class="hlt">2</span>+}–JNK mitochondrial pathways</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Zhang, Yuanyuan; Han, Lirong; Qi, Wentao</p> <p></p> <p>Highlights: • EPA evoked ROS formation, [Ca{sup <span class="hlt">2</span>+}]{sub c} accumulation, the opening of MPTP and the phosphorylation of JNK. • EPA-<span class="hlt">induced</span> [Ca{sup <span class="hlt">2</span>+}]{sub c} elevation was depended on production of ROS. • EPA-<span class="hlt">induced</span> ROS generation, [Ca{sup <span class="hlt">2</span>+}]{sub c} increase, and JNK activated caused MPTP opening. • The apoptosis <span class="hlt">induced</span> by EPA was related to release of cytochrome C through the MPTP. • EPA <span class="hlt">induced</span> Hep<span class="hlt">G</span><span class="hlt">2</span> cells apoptosis through ROS–Ca{sup <span class="hlt">2</span>+}–JNK mitochondrial pathways. - Abstract: Eicosapentaenoic acid (EPA), a well-known dietary n−3 PUFAS, has been considered to inhibit proliferation of tumor cells. However, the molecular mechanism related to EPA-<span class="hlt">induced</span> liver cancermore » cells apoptosis has not been reported. In this study, we investigated the effect of EPA on Hep<span class="hlt">G</span><span class="hlt">2</span> cells proliferation and apoptosis mechanism through mitochondrial pathways. EPA inhibited proliferation of Hep<span class="hlt">G</span><span class="hlt">2</span> cells in a dose-dependent manner and had no significant effect on the cell viability of humor normal liver L-02 cells. It was found that EPA initially evoked ROS formation, leading to [Ca{sup <span class="hlt">2</span>+}]{sub c} accumulation and the mitochondrial permeability transition pore (MPTP) opening; EPA-<span class="hlt">induced</span> Hep<span class="hlt">G</span><span class="hlt">2</span> cells apoptosis was inhibited by N-acetylcysteine (NAC, an inhibitor of ROS), 1,<span class="hlt">2</span>-bis (<span class="hlt">2</span>-aminophenoxy) ethane-N,N,N′,N′-tetraacetic acid (BAPTA-AM, a chelator of calcium) and CsA (inhibitor of MPTP). The relationship between ROS production, the increase of cytoplasmic Ca and MPTP opening was detected. It seems that ROS may act as an upstream regulator of EPA-<span class="hlt">induced</span> [Ca{sup <span class="hlt">2</span>+}]{sub c} generation, moreover, generation of ROS, overload of mitochondrial [Ca{sup <span class="hlt">2</span>+}]{sub c}, and JNK activated cause the opening of MPTP. Western blotting results showed that EPA elevated the phosphorylation status of JNK, processes associated with the ROS generation. Simultaneously, the apoptosis <span class="hlt">induced</span> by EPA was related to release of cytochrome C from mitochondria to cytoplasm through</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/4069583','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/4069583"><span>Abnormal centromere-<span class="hlt">chromatid</span> apposition (ACCA) and Peters' anomaly.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wertelecki, W; Dev, V G; Superneau, D W</p> <p>1985-08-01</p> <p>Abnormal centromere-<span class="hlt">chromatid</span> apposition (ACCA) was noted in a patient with Peters' anomaly. Previous reports of ACCA emphasized its association with tetraphocomelia and other congenital malformations (Roberts, SC Phocomelia, Pseudothalidomide Syndromes). This report expands the array of congenital malformations associated with ACCA and emphasizes the diagnostic importance of ocular defects for the ascertainment of additional cases of ACCA and its possible relationship with abnormal cell division.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4041464','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4041464"><span>Requirement for Parp-1 and DNA ligases 1 or 3 but not of Xrcc1 in chromosomal translocation formation by backup end joining</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Soni, Aashish; Siemann, Maria; Grabos, Martha; Murmann, Tamara; Pantelias, Gabriel E.; Iliakis, George</p> <p>2014-01-01</p> <p>In mammalian cells, ionizing radiation (IR)-<span class="hlt">induced</span> DNA double-strand <span class="hlt">breaks</span> (DSBs) are repaired in all phases of the cell cycle predominantly by classical, DNA-PK-dependent nonhomologous end joining (D-NHEJ). Homologous recombination repair (HRR) is functional during the S- and <span class="hlt">G</span><span class="hlt">2</span>-phases, when a sister <span class="hlt">chromatid</span> becomes available. An error-prone, alternative form of end joining, operating as backup (B-NHEJ) functions robustly throughout the cell cycle and particularly in the <span class="hlt">G</span><span class="hlt">2</span>-phase and is thought to backup predominantly D-NHEJ. Parp-1, DNA-ligases 1 (Lig1) and 3 (Lig3), and Xrcc1 are implicated in B-NHEJ. Chromosome and <span class="hlt">chromatid</span> translocations are manifestations of erroneous DSB repair and are crucial culprits in malignant transformation and IR-<span class="hlt">induced</span> cell lethality. We analyzed shifts in translocation formation deriving from defects in D-NHEJ or HRR in cells irradiated in the <span class="hlt">G</span><span class="hlt">2</span>-phase and identify B-NHEJ as the main DSB repair pathway backing up both of these defects at the cost of a large increase in translocation formation. Our results identify Parp-1 and Lig1 and 3 as factors involved in translocation formation and show that Xrcc1 reinforces the function of Lig3 in the process without being required for it. Finally, we demonstrate intriguing connections between B-NHEJ and DNA end resection in translocation formation and show that, as for D-NHEJ and HRR, the function of B-NHEJ facilitates the recovery from the <span class="hlt">G</span><span class="hlt">2</span>-checkpoint. These observations advance our understanding of chromosome aberration formation and have implications for the mechanism of action of Parp inhibitors. PMID:24748665</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040173044&hterms=Culture+differences&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3DCulture%2Bdifferences','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040173044&hterms=Culture+differences&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3DCulture%2Bdifferences"><span>Radiation-<span class="hlt">induced</span> chromosomal instability in BALB/c and C57BL/6 mice: the difference is as clear as black and white</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ponnaiya, B.; Cornforth, M. N.; Ullrich, R. L.</p> <p>1997-01-01</p> <p>Genomic instability has been proposed to be the earliest step in radiation-<span class="hlt">induced</span> tumorigenesis. It follows from this hypothesis that individuals highly susceptible to induction of tumors by radiation should exhibit enhanced radiation-<span class="hlt">induced</span> instability. BALB/c white mice are considerably more sensitive to radiation-<span class="hlt">induced</span> mammary cancer than C57BL/6 black mice. In this study, primary mammary epithelial cell cultures from these two strains were examined for the "delayed" appearance of chromosomal aberrations after exposure to 137Cs gamma radiation, as a measure of radiation-<span class="hlt">induced</span> genomic instability. As expected, actively dividing cultures from both strains showed a rapid decline of initial asymmetrical aberrations with time postirradiation. However, after 16 population doublings, cells from BALB/c mice exhibited a marked increase in the frequency of <span class="hlt">chromatid</span>-type <span class="hlt">breaks</span> and gaps which remained elevated throughout the time course of the experiment (28 doublings). No such effect was observed for the cells of C57BL/6 mice; after the rapid clearance of initial aberrations, the frequency of <span class="hlt">chromatid</span>-type aberrations in the irradiated population remained at or near those of nonirradiated controls. These results demonstrate a correlation between the latent expression of chromosomal damage in vitro and susceptibility for mammary tumors, and provide further support for the central role of radiation-<span class="hlt">induced</span> instability in the process of tumorigenesis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23291097','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23291097"><span>RPA accumulation during class switch recombination represents 5'-3' DNA-end resection during the S-<span class="hlt">G</span><span class="hlt">2</span>/M phase of the cell cycle.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yamane, Arito; Robbiani, Davide F; Resch, Wolfgang; Bothmer, Anne; Nakahashi, Hirotaka; Oliveira, Thiago; Rommel, Philipp C; Brown, Eric J; Nussenzweig, Andre; Nussenzweig, Michel C; Casellas, Rafael</p> <p>2013-01-31</p> <p>Activation-<span class="hlt">induced</span> cytidine deaminase (AID) promotes chromosomal translocations by <span class="hlt">inducing</span> DNA double-strand <span class="hlt">breaks</span> (DSBs) at immunoglobulin (Ig) genes and oncogenes in the <span class="hlt">G</span>1 phase. RPA is a single-stranded DNA (ssDNA)-binding protein that associates with resected DSBs in the S phase and facilitates the assembly of factors involved in homologous repair (HR), such as Rad51. Notably, RPA deposition also marks sites of AID-mediated damage, but its role in Ig gene recombination remains unclear. Here, we demonstrate that RPA associates asymmetrically with resected ssDNA in response to lesions created by AID, recombination-activating genes (RAG), or other nucleases. Small amounts of RPA are deposited at AID targets in <span class="hlt">G</span>1 in an ATM-dependent manner. In contrast, recruitment in the S-<span class="hlt">G</span><span class="hlt">2</span>/M phase is extensive, ATM independent, and associated with Rad51 accumulation. In the S-<span class="hlt">G</span><span class="hlt">2</span>/M phase, RPA increases in nonhomologous-end-joining-deficient lymphocytes, where there is more extensive DNA-end resection. Thus, most RPA recruitment during class switch recombination represents salvage of unrepaired <span class="hlt">breaks</span> by homology-based pathways during the S-<span class="hlt">G</span><span class="hlt">2</span>/M phase of the cell cycle. Copyright © 2013 The Authors. Published by Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22596308-caspase-independent-cell-death-mediated-apoptosis-inducing-factor-aif-nuclear-translocation-involved-ionizing-radiation-induced-hepg2-cell-death','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22596308-caspase-independent-cell-death-mediated-apoptosis-inducing-factor-aif-nuclear-translocation-involved-ionizing-radiation-induced-hepg2-cell-death"><span>Caspase-independent cell death mediated by apoptosis-<span class="hlt">inducing</span> factor (AIF) nuclear translocation is involved in ionizing radiation <span class="hlt">induced</span> Hep<span class="hlt">G</span><span class="hlt">2</span> cell death</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Sun, Hengwen; Yang, Shana; Li, Jianhua</p> <p></p> <p>Hepatocellular carcinoma (HCC) is the fifth most common cancer in the world. The aim of radiotherapy is to eradicate cancer cells with ionizing radiation. Except for the caspase-dependent mechanism, several lines of evidence demonstrated that caspase-independent mechanism is directly involved in the cell death responding to irradiation. For this reason, defining the contribution of caspase-independent molecular mechanisms represents the main goal in radiotherapy. In this study, we focused on the role of apoptosis-<span class="hlt">inducing</span> factor (AIF), the caspase-independent molecular, in ionizing radiation <span class="hlt">induced</span> hepatocellular carcinoma cell line (Hep<span class="hlt">G</span><span class="hlt">2</span>) cell death. We found that ionizing radiation has no function on AIF expressionmore » in Hep<span class="hlt">G</span><span class="hlt">2</span> cells, but could <span class="hlt">induce</span> AIF release from the mitochondria and translocate into nuclei. Inhibition of AIF could reduce ionizing radiation <span class="hlt">induced</span> Hep<span class="hlt">G</span><span class="hlt">2</span> cell death. These studies strongly support a direct relationship between AIF nuclear translocation and radiation <span class="hlt">induced</span> cell death. What's more, AIF nuclear translocation is caspase-independent manner, but not caspase-dependent manner, in this process. These new findings add a further attractive point of investigation to better define the complex interplay between caspase-independent cell death and radiation therapy. - Highlights: • AIF nuclear translocation is involved in ionizing radiation <span class="hlt">induced</span> hepatocellular carcinoma cell line Hep<span class="hlt">G</span><span class="hlt">2</span> cell death. • AIF mediated cell death <span class="hlt">induced</span> by ionizing radiation is caspase-independent. • Caspase-independent pathway is involved in ionzing radiation <span class="hlt">induced</span> Hep<span class="hlt">G</span><span class="hlt">2</span> cell death.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25744245','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25744245"><span><span class="hlt">G</span> protein-coupled estrogen receptor (GPER) mediates NSCLC progression <span class="hlt">induced</span> by 17β-estradiol (E<span class="hlt">2</span>) and selective agonist <span class="hlt">G</span>1.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liu, Changyu; Liao, Yongde; Fan, Sheng; Tang, Hexiao; Jiang, Zhixiao; Zhou, Bo; Xiong, Jing; Zhou, Sheng; Zou, Man; Wang, Jianmiao</p> <p>2015-04-01</p> <p>Estrogen classically drives lung cancer development via estrogen receptor β (ERβ). However, fulvestrant, an anti-estrogen-based endocrine therapeutic treatment, shows limited effects for non-small cell lung cancer (NSCLC) in phase II clinical trials. <span class="hlt">G</span> protein-coupled estrogen receptor (GPER), a third estrogen receptor that binds to estrogen, has been found to be activated by fulvestrant, stimulating the progression of breast, endometrial, and ovarian cancers. We here demonstrated that cytoplasm-GPER (cGPER) (80.49 %) and nucleus-GPER (53.05 %) were detected by immunohistochemical analysis in NSCLC samples. cGPER expression was related to stages IIIA-IV, lymph node metastasis, and poorly differentiated NSCLC. Selective agonist <span class="hlt">G</span>1 and 17β-estradiol (E<span class="hlt">2</span>) promoted the GPER-mediated proliferation, invasion, and migration of NSCLC cells. Additionally, in vitro administration of E<span class="hlt">2</span> and <span class="hlt">G</span>1 increased the number of tumor nodules, tumor grade, and tumor index in a urethane-<span class="hlt">induced</span> adenocarcinoma model. Importantly, the pro-tumorigenic effects of GPER <span class="hlt">induced</span> by E<span class="hlt">2</span> were significantly reduced by co-administering the GPER inhibitor <span class="hlt">G</span>15 and the ERβ inhibitor fulvestrant, as compared to administering fulvestrant alone both in vitro and in vivo. Moreover, the phosphorylation of MAPK and Akt was involved in E<span class="hlt">2</span>/<span class="hlt">G</span>1-<span class="hlt">induced</span> GPER activation. In conclusion, our results indicated that a pro-tumor function of GPER exists that mediated E<span class="hlt">2</span>-/<span class="hlt">G</span>1-dependent NSCLC progression and showed better efficiency regarding the co-targeting of GPER and ERβ, providing a rationale for further investigation of anti-estrogen clinical therapy.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29410075','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29410075"><span>The small molecule CS1 inhibits mitosis and sister <span class="hlt">chromatid</span> resolution in HeLa cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wu, Xingkang; Li, Zhenyu; Shen, Yuemao</p> <p>2018-05-01</p> <p>Mitosis, the most dramatic event in the cell cycle, involves the reorganization of virtually all cellular components. Antimitotic agents are useful for dissecting the mechanism of this reorganization. Previously, we found that the small molecule CS1 accumulates cells in <span class="hlt">G</span><span class="hlt">2</span>/M phase [1], but the mechanism of its action remains unknown. Cell cycle analysis, live cell imaging and nuclear staining were used. Chromosomal morphology was detected by chromosome spreading. The effects of CS1 on microtubules were confirmed by tubulin polymerization, colchicine tubulin-binding, cellular tubulin polymerization and immunofluorescence assays and by analysis of microtubule dynamics and molecular modeling. Histone phosphoproteomics was performed using mass spectrometry. Cell signaling cascades were analyzed using immunofluorescence, immunoprecipitation, immunoblotting, siRNA knockdown and chemical inhibition of specific proteins. The small molecule CS1 was shown to be an antimitotic agent. CS1 potently inhibited microtubule polymerization via interaction with the colchicine-binding pocket of tubulin in vitro and inhibited the formation of the spindle apparatus by reducing the bulk of growing microtubules in HeLa cells, which led to activation of the spindle assembly checkpoint (SAC) and mitotic arrest of HeLa cells. Compared with colchicine, CS1 impaired the progression of sister <span class="hlt">chromatid</span> resolution independent of cohesin dissociation, and this was reversed by the removal of CS1. Additionally, CS1 <span class="hlt">induced</span> unique histone phosphorylation patterns distinct from those <span class="hlt">induced</span> by colchicine. CS1 is a unique antimitotic small molecule and a powerful tool with unprecedented value over colchicine that makes it possible to specifically and conditionally perturb mitotic progression. Copyright © 2018 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017RSPSA.47370258M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017RSPSA.47370258M"><span>Modelling wave-<span class="hlt">induced</span> sea ice <span class="hlt">break</span>-up in the marginal ice zone</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Montiel, F.; Squire, V. A.</p> <p>2017-10-01</p> <p>A model of ice floe <span class="hlt">break</span>-up under ocean wave forcing in the marginal ice zone (MIZ) is proposed to investigate how floe size distribution (FSD) evolves under repeated wave <span class="hlt">break</span>-up events. A three-dimensional linear model of ocean wave scattering by a finite array of compliant circular ice floes is coupled to a flexural failure model, which <span class="hlt">breaks</span> a floe into two floes provided the two-dimensional stress field satisfies a <span class="hlt">break</span>-up criterion. A closed-feedback loop algorithm is devised, which (i) solves the wave-scattering problem for a given FSD under time-harmonic plane wave forcing, (ii) computes the stress field in all the floes, (iii) fractures the floes satisfying the <span class="hlt">break</span>-up criterion, and (iv) generates an updated FSD, initializing the geometry for the next iteration of the loop. The FSD after 50 <span class="hlt">break</span>-up events is unimodal and near normal, or bimodal, suggesting waves alone do not govern the power law observed in some field studies. Multiple scattering is found to enhance <span class="hlt">break</span>-up for long waves and thin ice, but to reduce <span class="hlt">break</span>-up for short waves and thick ice. A <span class="hlt">break</span>-up front marches forward in the latter regime, as wave-<span class="hlt">induced</span> fracture weakens the ice cover, allowing waves to travel deeper into the MIZ.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28781233','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28781233"><span>Phospho-H1 Decorates the Inter-<span class="hlt">chromatid</span> Axis and Is Evicted along with Shugoshin by SET during Mitosis.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Krishnan, Swathi; Smits, Arne H; Vermeulen, Michiel; Reinberg, Danny</p> <p>2017-08-17</p> <p>Precise control of sister <span class="hlt">chromatid</span> separation during mitosis is pivotal to maintaining genomic integrity. Yet, the regulatory mechanisms involved are not well understood. Remarkably, we discovered that linker histone H1 phosphorylated at S/T18 decorated the inter-<span class="hlt">chromatid</span> axial DNA on mitotic chromosomes. Sister <span class="hlt">chromatid</span> resolution during mitosis required the eviction of such H1S/T18ph by the chaperone SET, with this process being independent of and most likely downstream of arm-cohesin dissociation. SET also directed the disassembly of Shugoshins in a polo-like kinase 1-augmented manner, aiding centromere resolution. SET ablation compromised mitotic fidelity as evidenced by unresolved sister <span class="hlt">chromatids</span> with marked accumulation of H1S/T18ph and centromeric Shugoshin. Thus, chaperone-assisted eviction of linker histones and Shugoshins is a fundamental step in mammalian mitotic progression. Our findings also elucidate the functional implications of the decades-old observation of mitotic linker histone phosphorylation, serving as a paradigm to explore the role of linker histones in bio-signaling processes. Copyright © 2017 Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010OcMod..35..105S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010OcMod..35..105S"><span>Modeling quiescent phase transport of air bubbles <span class="hlt">induced</span> by <span class="hlt">breaking</span> waves</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shi, Fengyan; Kirby, James T.; Ma, Gangfeng</p> <p></p> <p>Simultaneous modeling of both the acoustic phase and quiescent phase of <span class="hlt">breaking</span> wave-<span class="hlt">induced</span> air bubbles involves a large range of length scales from microns to meters and time scales from milliseconds to seconds, and thus is computational unaffordable in a surfzone-scale computational domain. In this study, we use an air bubble entrainment formula in a two-fluid model to predict air bubble evolution in the quiescent phase in a <span class="hlt">breaking</span> wave event. The <span class="hlt">breaking</span> wave-<span class="hlt">induced</span> air bubble entrainment is formulated by connecting the shear production at the air-water interface and the bubble number intensity with a certain bubble size spectra observed in laboratory experiments. A two-fluid model is developed based on the partial differential equations of the gas-liquid mixture phase and the continuum bubble phase, which has multiple size bubble groups representing a polydisperse bubble population. An enhanced <span class="hlt">2</span>-DV VOF (Volume of Fluid) model with a k - ɛ turbulence closure is used to model the mixture phase. The bubble phase is governed by the advection-diffusion equations of the gas molar concentration and bubble intensity for groups of bubbles with different sizes. The model is used to simulate air bubble plumes measured in laboratory experiments. Numerical results indicate that, with an appropriate parameter in the air entrainment formula, the model is able to predict the main features of bubbly flows as evidenced by reasonable agreement with measured void fraction. Bubbles larger than an intermediate radius of O(1 mm) make a major contribution to void fraction in the near-crest region. Smaller bubbles tend to penetrate deeper and stay longer in the water column, resulting in significant contribution to the cross-sectional area of the bubble cloud. An underprediction of void fraction is found at the beginning of wave <span class="hlt">breaking</span> when large air pockets take place. The core region of high void fraction predicted by the model is dislocated due to use of the shear</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17384201','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17384201"><span>JS-K, a GST-activated nitric oxide generator, <span class="hlt">induces</span> DNA double-strand <span class="hlt">breaks</span>, activates DNA damage response pathways, and <span class="hlt">induces</span> apoptosis in vitro and in vivo in human multiple myeloma cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kiziltepe, Tanyel; Hideshima, Teru; Ishitsuka, Kenji; Ocio, Enrique M; Raje, Noopur; Catley, Laurence; Li, Chun-Qi; Trudel, Laura J; Yasui, Hiroshi; Vallet, Sonia; Kutok, Jeffery L; Chauhan, Dharminder; Mitsiades, Constantine S; Saavedra, Joseph E; Wogan, Gerald N; Keefer, Larry K; Shami, Paul J; Anderson, Kenneth C</p> <p>2007-07-15</p> <p>Here we investigated the cytotoxicity of JS-K, a prodrug designed to release nitric oxide (NO(*)) following reaction with glutathione S-transferases, in multiple myeloma (MM). JS-K showed significant cytotoxicity in both conventional therapy-sensitive and -resistant MM cell lines, as well as patient-derived MM cells. JS-K <span class="hlt">induced</span> apoptosis in MM cells, which was associated with PARP, caspase-8, and caspase-9 cleavage; increased Fas/CD95 expression; Mcl-1 cleavage; and Bcl-<span class="hlt">2</span> phosphorylation, as well as cytochrome c, apoptosis-<span class="hlt">inducing</span> factor (AIF), and endonuclease <span class="hlt">G</span> (Endo<span class="hlt">G</span>) release. Moreover, JS-K overcame the survival advantages conferred by interleukin-6 (IL-6) and insulin-like growth factor 1 (IGF-1), or by adherence of MM cells to bone marrow stromal cells. Mechanistic studies revealed that JS-K-<span class="hlt">induced</span> cytotoxicity was mediated via NO(*) in MM cells. Furthermore, JS-K <span class="hlt">induced</span> DNA double-strand <span class="hlt">breaks</span> (DSBs) and activated DNA damage responses, as evidenced by neutral comet assay, as well as H<span class="hlt">2</span>AX, Chk<span class="hlt">2</span> and p53 phosphorylation. JS-K also activated c-Jun NH(<span class="hlt">2</span>)-terminal kinase (JNK) in MM cells; conversely, inhibition of JNK markedly decreased JS-K-<span class="hlt">induced</span> cytotoxicity. Importantly, bortezomib significantly enhanced JS-K-<span class="hlt">induced</span> cytotoxicity. Finally, JS-K is well tolerated, inhibits tumor growth, and prolongs survival in a human MM xenograft mouse model. Taken together, these data provide the preclinical rationale for the clinical evaluation of JS-K to improve patient outcome in MM.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18160429','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18160429"><span>Human immunodeficiency virus type 1 Vpr <span class="hlt">induces</span> cell cycle <span class="hlt">G</span><span class="hlt">2</span> arrest through Srk1/MK<span class="hlt">2</span>-mediated phosphorylation of Cdc25.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Huard, Sylvain; Elder, Robert T; Liang, Dong; Li, Ge; Zhao, Richard Y</p> <p>2008-03-01</p> <p>Human immunodeficiency virus type 1 (HIV-1) Vpr <span class="hlt">induces</span> cell cycle <span class="hlt">G</span>(<span class="hlt">2</span>) arrest in fission yeast (Schizosaccharomyces pombe) and mammalian cells, suggesting the cellular pathway(s) targeted by Vpr is conserved among eukaryotes. Our previous studies in fission yeast demonstrated that Vpr <span class="hlt">induces</span> <span class="hlt">G</span>(<span class="hlt">2</span>) arrest in part through inhibition of Cdc25, a Cdc<span class="hlt">2</span>-specific phosphatase that promotes <span class="hlt">G</span>(<span class="hlt">2</span>)/M transition. The goal of this study was to further elucidate molecular mechanism underlying the inhibitory effect of Vpr on Cdc25. We show here that, similar to the DNA checkpoint controls, expression of vpr promotes subcellular relocalization of Cdc25 from nuclear to cytoplasm and thereby prevents activation of Cdc<span class="hlt">2</span> by Cdc25. Vpr-<span class="hlt">induced</span> nuclear exclusion of Cdc25 appears to depend on the serine/threonine phosphorylation of Cdc25 and the presence of Rad24/14-3-3 protein, since amino acid substitutions of the nine possible phosphorylation sites of Cdc25 with Ala (9A) or deletion of the rad24 gene abolished nuclear exclusion <span class="hlt">induced</span> by Vpr. Interestingly, Vpr is still able to promote Cdc25 nuclear export in mutants defective in the checkpoints (rad3 and chk1/cds1), the kinases that are normally required for Cdc25 phosphorylation and nuclear exclusion of Cdc25, suggesting that others kinase(s) might modulate phosphorylation of Cdc25 for the Vpr-<span class="hlt">induced</span> <span class="hlt">G</span>(<span class="hlt">2</span>) arrest. We report here that this kinase is Srk1. Deletion of the srk1 gene blocks the nuclear exclusion of Cdc25 caused by Vpr. Overexpression of srk1 <span class="hlt">induces</span> cell elongation, an indication of cell cycle <span class="hlt">G</span>(<span class="hlt">2</span>) delay, in a similar fashion to Vpr; however, no additive effect of cell elongation was observed when srk1 and vpr were coexpressed, indicating Srk1 and Vpr are likely affecting the cell cycle <span class="hlt">G</span>(<span class="hlt">2</span>)/M transition through the same cellular pathway. Immunoprecipitation further shows that Vpr and Srk1 are part of the same protein complex. Consistent with our findings in fission yeast, depletion of the MK<span class="hlt">2</span> gene, a human homologue</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22592808-traf1-knockdown-alleviates-palmitate-induced-insulin-resistance-hepg2-cells-through-nf-pathway','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22592808-traf1-knockdown-alleviates-palmitate-induced-insulin-resistance-hepg2-cells-through-nf-pathway"><span>TRAF1 knockdown alleviates palmitate-<span class="hlt">induced</span> insulin resistance in Hep<span class="hlt">G</span><span class="hlt">2</span> cells through NF-κB pathway</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Zhang, Wanlu; Jiangsu Province Key Laboratory for Inflammation and Molecular Drug Target, Nantong University, 19 Qixiu Road, Nantong 226001, Jiangsu Province; Tang, Zhuqi</p> <p></p> <p>High-fat diet (HFD) and inflammation are key contributors to insulin resistance (IR) and Type <span class="hlt">2</span> diabetes mellitus (T<span class="hlt">2</span>DM). With HFD, plasma free fatty acids (FFAs) can activate the nuclear factor-κB (NF-κB) in target tissues, then initiate negative crosstalk between FFAs and insulin signaling. However, the molecular link between IR and inflammation remains to be identified. We here reported that tumor necrosis factor receptor-associated factor 1 (TRAF1), an adapter in signal transduction, was involved in the onset of IR in hepatocytes. TRAF1 was significantly up-regulated in insulin-resistant liver tissues and palmitate (PA)-treated Hep<span class="hlt">G</span><span class="hlt">2</span> cells. In addition, we showed that depletion ofmore » TRAF1 led to inhibition of the activity of NF-κB. Given the fact that the activation of NF-κB played a facilitating role in IR, the phosphorylation of Akt and GSK3β was also analyzed. We found that depletion of TRAF1 markedly reversed PA-<span class="hlt">induced</span> attenuation of the phosphorylation of Akt and GSK3β in the cells. The accumulation of lipid droplets in hepatocyte and expression of two key gluconeogenic enzymes, PEPCK and <span class="hlt">G</span>6Pase, were also determined and found to display a similar tendency with the phosphorylation of Akt and GSK3β. Glucose uptake assay indicated that knocking down TRAF1 blocked the effect of PA on the suppression of glucose uptake. These data implicated that TRAF1 knockdown might alleviate PA-<span class="hlt">induced</span> IR in Hep<span class="hlt">G</span><span class="hlt">2</span> cells through NF-κB pathway. - Highlights: • TRAF1 accelerated PA-<span class="hlt">induced</span> IR in Hep<span class="hlt">G</span><span class="hlt">2</span> cells mediated through NF-κB signaling. • Knockdown of TRAF1 alleviated PA-<span class="hlt">induced</span> IR in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. • Knockdown of TRAF1 alleviated PA-<span class="hlt">induced</span> lipid accumulation in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. • Knockdown of TRAF1 reversed PA-<span class="hlt">induced</span> suppression of glucose uptake in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. • Knockdown of TRAF1 reversed PA-<span class="hlt">induced</span> gluconeogenesis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5666229','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5666229"><span>Modelling wave-<span class="hlt">induced</span> sea ice <span class="hlt">break</span>-up in the marginal ice zone</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Squire, V. A.</p> <p>2017-01-01</p> <p>A model of ice floe <span class="hlt">break</span>-up under ocean wave forcing in the marginal ice zone (MIZ) is proposed to investigate how floe size distribution (FSD) evolves under repeated wave <span class="hlt">break</span>-up events. A three-dimensional linear model of ocean wave scattering by a finite array of compliant circular ice floes is coupled to a flexural failure model, which <span class="hlt">breaks</span> a floe into two floes provided the two-dimensional stress field satisfies a <span class="hlt">break</span>-up criterion. A closed-feedback loop algorithm is devised, which (i) solves the wave-scattering problem for a given FSD under time-harmonic plane wave forcing, (ii) computes the stress field in all the floes, (iii) fractures the floes satisfying the <span class="hlt">break</span>-up criterion, and (iv) generates an updated FSD, initializing the geometry for the next iteration of the loop. The FSD after 50 <span class="hlt">break</span>-up events is unimodal and near normal, or bimodal, suggesting waves alone do not govern the power law observed in some field studies. Multiple scattering is found to enhance <span class="hlt">break</span>-up for long waves and thin ice, but to reduce <span class="hlt">break</span>-up for short waves and thick ice. A <span class="hlt">break</span>-up front marches forward in the latter regime, as wave-<span class="hlt">induced</span> fracture weakens the ice cover, allowing waves to travel deeper into the MIZ. PMID:29118659</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29118659','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29118659"><span>Modelling wave-<span class="hlt">induced</span> sea ice <span class="hlt">break</span>-up in the marginal ice zone.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Montiel, F; Squire, V A</p> <p>2017-10-01</p> <p>A model of ice floe <span class="hlt">break</span>-up under ocean wave forcing in the marginal ice zone (MIZ) is proposed to investigate how floe size distribution (FSD) evolves under repeated wave <span class="hlt">break</span>-up events. A three-dimensional linear model of ocean wave scattering by a finite array of compliant circular ice floes is coupled to a flexural failure model, which <span class="hlt">breaks</span> a floe into two floes provided the two-dimensional stress field satisfies a <span class="hlt">break</span>-up criterion. A closed-feedback loop algorithm is devised, which (i) solves the wave-scattering problem for a given FSD under time-harmonic plane wave forcing, (ii) computes the stress field in all the floes, (iii) fractures the floes satisfying the <span class="hlt">break</span>-up criterion, and (iv) generates an updated FSD, initializing the geometry for the next iteration of the loop. The FSD after 50 <span class="hlt">break</span>-up events is unimodal and near normal, or bimodal, suggesting waves alone do not govern the power law observed in some field studies. Multiple scattering is found to enhance <span class="hlt">break</span>-up for long waves and thin ice, but to reduce <span class="hlt">break</span>-up for short waves and thick ice. A <span class="hlt">break</span>-up front marches forward in the latter regime, as wave-<span class="hlt">induced</span> fracture weakens the ice cover, allowing waves to travel deeper into the MIZ.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li class="active"><span>7</span></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_7 --> <div id="page_8" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li class="active"><span>8</span></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="141"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28395717','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28395717"><span>[Arginase inhibitor nor-NOHA <span class="hlt">induces</span> apoptosis and inhibits invasion and migration of Hep<span class="hlt">G</span><span class="hlt">2</span> cells].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Li, Xiangnan; Zhu, Fangyu; He, Yongsong; Luo, Fang</p> <p>2017-04-01</p> <p>Objective To investigate the cell inhibitory effect of arginase inhibitor nor-NOHA on Hep<span class="hlt">G</span><span class="hlt">2</span> hepatocellular carcinoma cells and related mechanism. Methods CCK-8 assay was used to detect the cell proliferation and flow cytometry to detect the apoptosis of Hep<span class="hlt">G</span><span class="hlt">2</span> cells treated with (0, 0.5, 1.0, <span class="hlt">2</span>.0, 3.0) ng/μL nor-NOHA. The protein levels of arginase 1 (Arg1), P53, matrix metalloproteinase-<span class="hlt">2</span> (MMP-<span class="hlt">2</span>), E-cadherin (ECD) were determined by Western blotting. Real time quantitative PCR was employed to examine the changes in the mRNA level of <span class="hlt">inducible</span> nitric oxide synthase (iNOS). Griess assay was used to measure the concentration of nitric oxide (NO) in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Transwell TM assay and wound-healing assay were performed to evaluate the changes of the cell invasion and migration ability, respectively. Results nor-NOHA inhibited the proliferation and <span class="hlt">induced</span> the apoptosis of Hep<span class="hlt">G</span><span class="hlt">2</span> cells. It also decreased the expression levels of Arg1 and MMP-<span class="hlt">2</span>, increased the expression levels of P53 and ECD as well as the production of NO; in addition, nor-NOHA inhibited the invasion and migration of Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Conclusion Nor-NOHA can <span class="hlt">induce</span> cell apoptosis and inhibit the ability of invasion and migration of Hep<span class="hlt">G</span><span class="hlt">2</span> cells by inhibiting Arg1, which is related with the increase of iNOS expression and the high concentration of NO.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22416703-carboxyethenyl-isatin-derivative-induces-sub-cell-cycle-arrest-apoptosis-human-leukemia-k562-cells','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22416703-carboxyethenyl-isatin-derivative-induces-sub-cell-cycle-arrest-apoptosis-human-leukemia-k562-cells"><span>5-(<span class="hlt">2</span>-Carboxyethenyl) isatin derivative <span class="hlt">induces</span> <span class="hlt">G</span>{sub <span class="hlt">2</span>}/M cell cycle arrest and apoptosis in human leukemia K562 cells</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Zhou, Yao; Zhao, Hong-Ye; Han, Kai-Lin</p> <p>2014-08-08</p> <p>Highlights: • 5-(<span class="hlt">2</span>-Carboxyethenyl) isatin derivative (HKL <span class="hlt">2</span>H) inhibited K562’s proliferation. • HKL <span class="hlt">2</span>H caused the morphology change of <span class="hlt">G</span>{sub <span class="hlt">2</span>}/M phase arrest and typical apoptosis. • HKL <span class="hlt">2</span>H <span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span>/M cell cycle phase arrest in K562 cells. • HKL <span class="hlt">2</span>H <span class="hlt">induced</span> apoptosis in K562 cells through the mitochondrial pathway. - Abstract: Our previous study successfully identified that the novel isatin derivative (E)-methyl 3-(1-(4-methoxybenzyl)-<span class="hlt">2</span>,3-dioxoindolin-5-yl) acrylate (HKL <span class="hlt">2</span>H) acts as an anticancer agent at an inhibitory concentration (IC{sub 50}) level of 3 nM. In this study, the molecular mechanism how HKL <span class="hlt">2</span>H <span class="hlt">induces</span> cytotoxic activity in the human chronic myelogenous leukemia K562more » cells was investigated. Flow cytometric analysis showed that the cells were arrested in the <span class="hlt">G</span>{sub <span class="hlt">2</span>}/M phase and accumulated subsequently in the sub-<span class="hlt">G</span>{sub 1} phase in the presence of HKL <span class="hlt">2</span>H. HKL <span class="hlt">2</span>H treatment down-regulated the expressions of CDK1 and cyclin B but up-regulated the level of phosphorylated CDK1. Annexin-V staining and the classic DNA ladder studies showed that HKL <span class="hlt">2</span>H <span class="hlt">induced</span> the apoptosis of K562 cells. Our study further showed that HKL <span class="hlt">2</span>H treatment caused the dissipation of mitochondrial membrane potential, activated caspase-3 and lowered the Bcl-<span class="hlt">2</span>/Bax ratio in K562 cells, suggesting that the HKL <span class="hlt">2</span>H-causing programmed cell death of K562 cells was caused via the mitochondrial apoptotic pathway. Taken together, our data demonstrated that HKL <span class="hlt">2</span>H, a 5-(<span class="hlt">2</span>-carboxyethenyl) isatin derivative, notably <span class="hlt">induces</span> <span class="hlt">G</span>{sub <span class="hlt">2</span>}/M cell cycle arrest and mitochondrial-mediated apoptosis in K562 cells, indicating that this compound could be a promising anticancer candidate for further investigation.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4840299','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4840299"><span>PolySUMOylation by Siz<span class="hlt">2</span> and Mms21 triggers relocation of DNA <span class="hlt">breaks</span> to nuclear pores through the Slx5/Slx8 STUbL</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Horigome, Chihiro; Bustard, Denise E.; Marcomini, Isabella; Delgoshaie, Neda; Tsai-Pflugfelder, Monika; Cobb, Jennifer A.; Gasser, Susan M.</p> <p>2016-01-01</p> <p>High-resolution imaging shows that persistent DNA damage in budding yeast localizes in distinct perinuclear foci for repair. The signals that trigger DNA double-strand <span class="hlt">break</span> (DSB) relocation or determine their destination are unknown. We show here that DSB relocation to the nuclear envelope depends on SUMOylation mediated by the E3 ligases Siz<span class="hlt">2</span> and Mms21. In <span class="hlt">G</span>1, a polySUMOylation signal deposited coordinately by Mms21 and Siz<span class="hlt">2</span> recruits the SUMO targeted ubiquitin ligase Slx5/Slx8 to persistent <span class="hlt">breaks</span>. Both Slx5 and Slx8 are necessary for damage relocation to nuclear pores. When targeted to an undamaged locus, however, Slx5 alone can mediate relocation in <span class="hlt">G</span>1-phase cells, bypassing the requirement for polySUMOylation. In contrast, in S-phase cells, monoSUMOylation mediated by the Rtt107-stabilized SMC5/6–Mms21 E3 complex drives DSBs to the SUN domain protein Mps3 in a manner independent of Slx5. Slx5/Slx8 and binding to pores favor repair by ectopic <span class="hlt">break-induced</span> replication and imprecise end-joining. PMID:27056668</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27155388','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27155388"><span>Exogenous FABP4 <span class="hlt">induces</span> endoplasmic reticulum stress in Hep<span class="hlt">G</span><span class="hlt">2</span> liver cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bosquet, Alba; Guaita-Esteruelas, Sandra; Saavedra, Paula; Rodríguez-Calvo, Ricardo; Heras, Mercedes; Girona, Josefa; Masana, Lluís</p> <p>2016-06-01</p> <p>Fatty acid binding protein 4 (FABP4) is an intracellular fatty acid (FA) carrier protein that is, in part, secreted into circulation. Circulating FABP4 levels are increased in obesity, diabetes and other insulin resistance (IR) diseases. FAs contribute to IR by promoting endoplasmic reticulum stress (ER stress) and altering the insulin signaling pathway. The effect of FABP4 on ER stress in the liver is not known. The aim of this study was to investigate whether exogenous FABP4 (eFABP4) is involved in the lipid-<span class="hlt">induced</span> ER stress in the liver. Hep<span class="hlt">G</span><span class="hlt">2</span> cells were cultured with eFABP4 (40 ng/ml) with or without linoleic acid (LA, 200 μM) for 18 h. The expression of ER stress-related markers was determined by Western blotting (ATF6, EIF<span class="hlt">2</span>α, IRE1 and ubiquitin) and real-time PCR (ATF6, CHOP, EIF<span class="hlt">2</span>α and IRE1). Apoptosis was studied by flow cytometry using Annexin V-FITC and propidium iodide staining. eFABP4 increased the ER stress markers ATF6 and IRE1 in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. This effect led to insulin resistance mediated by changes in AKT and JNK phosphorylation. Furthermore, eFABP4 significantly <span class="hlt">induced</span> both apoptosis, as assessed by flow cytometry, and CHOP expression, without affecting necrosis and ubiquitination. The presence of LA increased the ER stress response <span class="hlt">induced</span> by eFABP4. eFABP4, per se, <span class="hlt">induces</span> ER stress and potentiates the effect of LA in Hep<span class="hlt">G</span><span class="hlt">2</span> cells, suggesting that FABP4 could be a link between obesity-associated metabolic abnormalities and hepatic IR mechanisms. Copyright © 2016 Elsevier Ireland Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27715492','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27715492"><span>Connecting <span class="hlt">G</span> protein signaling to chemoattractant-mediated cell polarity and cytoskeletal reorganization.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liu, Youtao; Lacal, Jesus; Firtel, Richard A; Kortholt, Arjan</p> <p>2018-07-04</p> <p>The directional movement toward extracellular chemical gradients, a process called chemotaxis, is an important property of cells. Central to eukaryotic chemotaxis is the molecular mechanism by which chemoattractant-mediated activation of <span class="hlt">G</span>-protein coupled receptors (GPCRs) <span class="hlt">induces</span> symmetry <span class="hlt">breaking</span> in the activated downstream signaling pathways. Studies with mainly Dictyostelium and mammalian neutrophils as experimental systems have shown that chemotaxis is mediated by a complex network of signaling pathways. Recently, several labs have used extensive and efficient proteomic approaches to further unravel this dynamic signaling network. Together these studies showed the critical role of the interplay between heterotrimeric <span class="hlt">G</span>-protein subunits and monomeric <span class="hlt">G</span> proteins in regulating cytoskeletal rearrangements during chemotaxis. Here we highlight how these proteomic studies have provided greater insight into the mechanisms by which the heterotrimeric <span class="hlt">G</span> protein cycle is regulated, how heterotrimeric <span class="hlt">G</span> proteins-<span class="hlt">induced</span> symmetry <span class="hlt">breaking</span> is mediated through small <span class="hlt">G</span> protein signaling, and how symmetry <span class="hlt">breaking</span> in <span class="hlt">G</span> protein signaling subsequently <span class="hlt">induces</span> cytoskeleton rearrangements and cell migration.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=43487&Lab=ORD&keyword=bone&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50','EPA-EIMS'); return false;" href="https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=43487&Lab=ORD&keyword=bone&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50"><span>EVIDENCE FOR THE CHROMOSOMAL REPLICONS AS UNITS OF SISTER <span class="hlt">CHROMATID</span> EXCHANGES</span></a></p> <p><a target="_blank" href="http://oaspub.epa.gov/eims/query.page">EPA Science Inventory</a></p> <p></p> <p></p> <p>Current hypotheses of sister <span class="hlt">chromatid</span> exchange (SCE) formation postulate that sites of SCE induction are associated with active replicons or replicon clusters. We have applied the FCC-SCD technique to in vivo studies of mouse bone marrow cells that have been treated with cycloph...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4513114','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4513114"><span>hMSH5 Facilitates the Repair of Camptothecin-<span class="hlt">induced</span> Double-strand <span class="hlt">Breaks</span> through an Interaction with FANCJ*</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Xu, Yang; Wu, Xiling; Her, Chengtao</p> <p>2015-01-01</p> <p>Replication stress from stalled or collapsed replication forks is a major challenge to genomic integrity. The anticancer agent camptothecin (CPT) is a DNA topoisomerase I inhibitor that causes fork collapse and double-strand <span class="hlt">breaks</span> amid DNA replication. Here we report that hMSH5 promotes cell survival in response to CPT-<span class="hlt">induced</span> DNA damage. Cells deficient in hMSH5 show elevated CPT-<span class="hlt">induced</span> γ-H<span class="hlt">2</span>AX and RPA<span class="hlt">2</span> foci with concomitant reduction of Rad51 foci, indicative of impaired homologous recombination. In addition, CPT-treated hMSH5-deficient cells exhibit aberrant activation of Chk1 and Chk<span class="hlt">2</span> kinases and therefore abnormal cell cycle progression. Furthermore, the hMSH5-FANCJ chromatin recruitment underlies the effects of hMSH5 on homologous recombination and Chk1 activation. Intriguingly, FANCJ depletion desensitizes hMSH5-deficient cells to CPT-elicited cell killing. Collectively, our data point to the existence of a functional interplay between hMSH5 and FANCJ in double-strand <span class="hlt">break</span> repair <span class="hlt">induced</span> by replication stress. PMID:26055704</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040088538&hterms=potential+action&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dpotential%2Baction','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040088538&hterms=potential+action&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dpotential%2Baction"><span>Evidence for factors modulating radiation-<span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span>-delay: potential application as radioprotectors</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Cheong, N.; Zeng, Z. C.; Wang, Y.; Iliakis, G.</p> <p>2001-01-01</p> <p>Manipulation of checkpoint response to DNA damage can be developed as a means for protecting astronauts from the adverse effects of unexpected, or background exposures to ionizing radiation. To achieve this goal reagents need to be developed that protect cells from radiation injury by prolonging checkpoint response, thus promoting repair. We present evidence for a low molecular weight substance excreted by cells that dramatically increases the duration of the <span class="hlt">G</span><span class="hlt">2</span>-delay. This compound is termed <span class="hlt">G</span><span class="hlt">2</span>-Arrest Modulating Activity (GAMA). A rat cell line (A1-5) generated by transforming rat embryo fibroblasts with a temperature sensitive form of p53 plus H-ras demonstrates a dramatic increase in radiation resistance after exposure to low LET radiation that is not associated with an increase in the efficiency of rejoining of DNA double strand <span class="hlt">breaks</span>. Radioresistance in this cell line correlates with a dramatic increase in the duration of the <span class="hlt">G</span><span class="hlt">2</span> arrest that is modulated by a GAMA produced by actively growing cells. The properties of GAMA suggest that it is a low molecular weight heat-stable peptide. Further characterization of this substance and elucidation of its mechanism of action may allow the development of a biological response modifier with potential applications as a radioprotector. GAMA may be useful for protecting astronauts from radiation injury as preliminary evidence suggests that it is able to modulate the response of cells exposed to heavy ion radiation, similar to that encountered in outer space.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20167213','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20167213"><span>Genotoxic effect of 6-gingerol on human hepatoma <span class="hlt">G</span><span class="hlt">2</span> cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yang, Guang; Zhong, Laifu; Jiang, Liping; Geng, Chengyan; Cao, Jun; Sun, Xiance; Ma, Yufang</p> <p>2010-04-15</p> <p>6-gingerol, a major component of ginger, has antioxidant, anti-apoptotic, and anti-inflammatory activities. However, some dietary phytochemicals possess pro-oxidant effects as well, and the risk of adverse effects is increased by raising the use of doses. The aim of this study was to assess the genotoxic effects of 6-gingerol and to clarify the mechanisms, using human hepatoma <span class="hlt">G</span><span class="hlt">2</span> (Hep<span class="hlt">G</span><span class="hlt">2</span>) cells. Exposure of the cells to 6-gingerol caused significant increase of DNA migration in comet assay, increase of micronuclei frequencies at high concentrations at 20-80 and 20-40 microM, respectively. These results indicate that 6-gingerol caused DNA strand <span class="hlt">breaks</span> and chromosome damage. To further elucidate the underlying mechanisms, we tested lysosomal membrane stability, mitochondrial membrane potential, the intracellular generation of reactive oxygen species (ROS) and reduced glutathione (GSH). In addition, the level of oxidative DNA damage was evaluated by immunocytochemical analysis on 8-hydroxydeoxyguanosine (8-OHd<span class="hlt">G</span>). Results showed that lysosomal membrane stability was reduced after treatment by 6-gingerol (20-80 microM) for 40 min, mitochondrial membrane potential decreased after treatment for 50 min, GSH and ROS levels were significantly increased after treatment for 60 min. These suggest 6-gingerol <span class="hlt">induces</span> genotoxicity probably by oxidative stress; lysosomal and mitochondrial damage were observed in 6-gingerol-<span class="hlt">induced</span> toxicity. Copyright 2010 Elsevier Ireland Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24748665','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24748665"><span>Requirement for Parp-1 and DNA ligases 1 or 3 but not of Xrcc1 in chromosomal translocation formation by backup end joining.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Soni, Aashish; Siemann, Maria; Grabos, Martha; Murmann, Tamara; Pantelias, Gabriel E; Iliakis, George</p> <p>2014-06-01</p> <p>In mammalian cells, ionizing radiation (IR)-<span class="hlt">induced</span> DNA double-strand <span class="hlt">breaks</span> (DSBs) are repaired in all phases of the cell cycle predominantly by classical, DNA-PK-dependent nonhomologous end joining (D-NHEJ). Homologous recombination repair (HRR) is functional during the S- and <span class="hlt">G</span><span class="hlt">2</span>-phases, when a sister <span class="hlt">chromatid</span> becomes available. An error-prone, alternative form of end joining, operating as backup (B-NHEJ) functions robustly throughout the cell cycle and particularly in the <span class="hlt">G</span><span class="hlt">2</span>-phase and is thought to backup predominantly D-NHEJ. Parp-1, DNA-ligases 1 (Lig1) and 3 (Lig3), and Xrcc1 are implicated in B-NHEJ. Chromosome and <span class="hlt">chromatid</span> translocations are manifestations of erroneous DSB repair and are crucial culprits in malignant transformation and IR-<span class="hlt">induced</span> cell lethality. We analyzed shifts in translocation formation deriving from defects in D-NHEJ or HRR in cells irradiated in the <span class="hlt">G</span><span class="hlt">2</span>-phase and identify B-NHEJ as the main DSB repair pathway backing up both of these defects at the cost of a large increase in translocation formation. Our results identify Parp-1 and Lig1 and 3 as factors involved in translocation formation and show that Xrcc1 reinforces the function of Lig3 in the process without being required for it. Finally, we demonstrate intriguing connections between B-NHEJ and DNA end resection in translocation formation and show that, as for D-NHEJ and HRR, the function of B-NHEJ facilitates the recovery from the <span class="hlt">G</span><span class="hlt">2</span>-checkpoint. These observations advance our understanding of chromosome aberration formation and have implications for the mechanism of action of Parp inhibitors. © The Author(s) 2014. Published by Oxford University Press on behalf of Nucleic Acids Research.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28137606','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28137606"><span>Protective effects of rambutan (Nephelium lappaceum) peel phenolics on H<span class="hlt">2</span>O<span class="hlt">2</span>-<span class="hlt">induced</span> oxidative damages in Hep<span class="hlt">G</span><span class="hlt">2</span> cells and d-galactose-<span class="hlt">induced</span> aging mice.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhuang, Yongliang; Ma, Qingyu; Guo, Yan; Sun, Liping</p> <p>2017-10-01</p> <p>Rambutan peel phenolic (RPP) extracts were prepared via dynamic separation with macroporous resin. The total phenolic content and individual phenolics in RPP were determined. Results showed that the total phenolic content of RPP was 877.11 mg gallic acid equivalents (GAE)/<span class="hlt">g</span> extract. The content of geranin (122.18 mg/<span class="hlt">g</span> extract) was the highest among those of the 39 identified phenolic compounds. RPP protected against oxidative stress in H <span class="hlt">2</span> O <span class="hlt">2</span> -<span class="hlt">induced</span> Hep<span class="hlt">G</span><span class="hlt">2</span> cells in a dose-response manner. The inhibitory effects of RPP on cell apoptosis might be related to its inhibitory effects on the generation of intracellular reactive oxygen species and increased effects on superoxide dismutase activity. The in vivo anti-aging activity of RPP was evaluated using an aging mice model that was <span class="hlt">induced</span> by d-galactose (d-gal). The results showed that RPP enhanced the antioxidative status of experimental mice. Moreover, histological analysis indicated that RPP effectively reduced d-gal-<span class="hlt">induced</span> liver and kidney tissue damage in a dose-dependent manner. Therefore, RPP can be used as a natural antioxidant and anti-aging agent in the pharmaceutical and food industries. Copyright © 2017 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26847723','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26847723"><span>Kaempferol <span class="hlt">induces</span> apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells via activation of the endoplasmic reticulum stress pathway.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Guo, Haiqing; Ren, Feng; Zhang, Li; Zhang, Xiangying; Yang, Rongrong; Xie, Bangxiang; Li, Zhuo; Hu, Zhongjie; Duan, Zhongping; Zhang, Jing</p> <p>2016-03-01</p> <p>Kaempferol is a flavonoid compound that has gained importance due to its antitumor properties; however, the underlying mechanisms remain to be fully understood. The present study aimed to investigate the molecular mechanisms of the antitumor function of kaempferol in Hep<span class="hlt">G</span><span class="hlt">2</span> hepatocellular carcinoma cells. Kaempferol was determined to reduce cell viability, increase lactate dehydrogenase activity and <span class="hlt">induce</span> apoptosis in a concentration‑ and time‑dependent manner in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Additionally, kaempferol‑<span class="hlt">induced</span> apoptosis possibly acts via the endoplasmic reticulum (ER) stress pathway, due to the significant increase in the protein expression levels of glucose‑regulated protein 78, glucose‑regulated protein 94, protein kinase R‑like ER kinase, inositol‑requiring enzyme 1α, partial activating transcription factor 6 cleavage, caspase‑4, C/EBP homologous protein (CHOP) and cleaved caspase‑3. The pro‑apoptotic activity of kaempferol was determined to be due to induction of the ER stress‑CHOP pathway, as: i) ER stress was blocked by 4‑phenyl butyric acid (4‑PBA) pretreatment and knockdown of CHOP with small interfering RNA, which resulted in alleviation of kaempferol‑<span class="hlt">induced</span> Hep<span class="hlt">G</span><span class="hlt">2</span> cell apoptosis; and ii) transfection with plasmid overexpressing CHOP reversed the protective effect of 4‑PBA in kaempferol‑<span class="hlt">induced</span> Hep<span class="hlt">G</span><span class="hlt">2</span> cells and increased the apoptotic rate. Thus, kaempferol promoted Hep<span class="hlt">G</span><span class="hlt">2</span> cell apoptosis via induction of the ER stress‑CHOP signaling pathway. These observations indicate that kaempferol may be used as a potential chemopreventive treatment strategy for patients with hepatocellular carcinoma.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/8052689','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/8052689"><span>DNA double-strand <span class="hlt">breaks</span> <span class="hlt">induced</span> by high-energy neon and iron ions in human fibroblasts. II. Probing individual notI fragments by hybridization.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Löbrich, M; Rydberg, B; Cooper, P K</p> <p>1994-08-01</p> <p>The initial yields of DNA double-strand <span class="hlt">breaks</span> <span class="hlt">induced</span> by energetic heavy ions (425 MeV/u neon and 250, 400 and 600 MeV/u iron) in comparison to X rays were measured in normal human diploid fibroblast cells within three small areas of the genome, defined by NotI fragments of 3.<span class="hlt">2</span>, <span class="hlt">2</span>.0 and 1.<span class="hlt">2</span> Mbp. The methodology involves NotI restriction endonuclease digestion of DNA from irradiated cells, followed by pulsed-field gel electrophoresis, Southern blotting and hybridization with probes recognizing single-copy sequences within the three NotI fragments. The gradual disappearance of the full-size NotI fragment with dose and the appearance of a smear of broken DNA molecules are quantified. Assuming Poisson statistics for the number of double-strand <span class="hlt">breaks</span> <span class="hlt">induced</span> per NotI fragment of known size, absolute yields of DNA double-strand <span class="hlt">breaks</span> were calculated and determined to be linear with dose in all cases, with the neon ion (LET 32 keV/microns) producing 4.4 x 10(-3) <span class="hlt">breaks</span>/Mbp/Gy and all three iron-ion beams (LETs from 190 to 350 keV/microns) producing <span class="hlt">2</span>.8 x 10(-3) <span class="hlt">breaks</span>/Mbp/Gy, giving RBE values for production of double-strand <span class="hlt">breaks</span> of 0.76 for neon and 0.48 for iron in comparison to our previously determined X-ray induction rate of 5.8 x 10(-3) <span class="hlt">breaks</span>/Mbp/Gy. These RBE values are in good agreement with results of measurements over the whole genome as reported in the accompanying paper (B. Rydberg, M. Löbrich and P. Cooper, Radiat. Res. 139, 133-141, 1994). The distribution of broken DNA molecules was similar for the various radiations, supporting a random distribution of double-strand <span class="hlt">breaks</span> <span class="hlt">induced</span> by the heavy ions over Mbp distances; however, correlated <span class="hlt">breaks</span> (clusters) over much smaller distances are not ruled out. Reconstitution of the 3.<span class="hlt">2</span> Mbp NotI fragment was studied during postirradiation incubation of the cells as a measure of rejoining of correct DNA ends. The proportion of <span class="hlt">breaks</span> repaired decreased with increasing LET.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26734525','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26734525"><span>Evidence for non-conservative current-<span class="hlt">induced</span> forces in the <span class="hlt">breaking</span> of Au and Pt atomic chains.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sabater, Carlos; Untiedt, Carlos; van Ruitenbeek, Jan M</p> <p>2015-01-01</p> <p>This experimental work aims at probing current-<span class="hlt">induced</span> forces at the atomic scale. Specifically it addresses predictions in recent work regarding the appearance of run-away modes as a result of a combined effect of the non-conservative wind force and a 'Berry force'. The systems we consider here are atomic chains of Au and Pt atoms, for which we investigate the distribution of <span class="hlt">break</span> down voltage values. We observe two distinct modes of <span class="hlt">breaking</span> for Au atomic chains. The <span class="hlt">breaking</span> at high voltage appears to behave as expected for regular <span class="hlt">break</span> down by thermal excitation due to Joule heating. However, there is a low-voltage <span class="hlt">breaking</span> mode that has characteristics expected for the mechanism of current-<span class="hlt">induced</span> forces. Although a full comparison would require more detailed information on the individual atomic configurations, the systems we consider are very similar to those considered in recent model calculations and the comparison between experiment and theory is very encouraging for the interpretation we propose.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4685917','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4685917"><span>Evidence for non-conservative current-<span class="hlt">induced</span> forces in the <span class="hlt">breaking</span> of Au and Pt atomic chains</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Sabater, Carlos; Untiedt, Carlos</p> <p>2015-01-01</p> <p>Summary This experimental work aims at probing current-<span class="hlt">induced</span> forces at the atomic scale. Specifically it addresses predictions in recent work regarding the appearance of run-away modes as a result of a combined effect of the non-conservative wind force and a ‘Berry force’. The systems we consider here are atomic chains of Au and Pt atoms, for which we investigate the distribution of <span class="hlt">break</span> down voltage values. We observe two distinct modes of <span class="hlt">breaking</span> for Au atomic chains. The <span class="hlt">breaking</span> at high voltage appears to behave as expected for regular <span class="hlt">break</span> down by thermal excitation due to Joule heating. However, there is a low-voltage <span class="hlt">breaking</span> mode that has characteristics expected for the mechanism of current-<span class="hlt">induced</span> forces. Although a full comparison would require more detailed information on the individual atomic configurations, the systems we consider are very similar to those considered in recent model calculations and the comparison between experiment and theory is very encouraging for the interpretation we propose. PMID:26734525</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5321743','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5321743"><span>CX-5461 is a DNA <span class="hlt">G</span>-quadruplex stabilizer with selective lethality in BRCA1/<span class="hlt">2</span> deficient tumours</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Xu, Hong; Di Antonio, Marco; McKinney, Steven; Mathew, Veena; Ho, Brandon; O'Neil, Nigel J.; Santos, Nancy Dos; Silvester, Jennifer; Wei, Vivien; Garcia, Jessica; Kabeer, Farhia; Lai, Daniel; Soriano, Priscilla; Banáth, Judit; Chiu, Derek S.; Yap, Damian; Le, Daniel D.; Ye, Frank B.; Zhang, Anni; Thu, Kelsie; Soong, John; Lin, Shu-chuan; Tsai, Angela Hsin Chin; Osako, Tomo; Algara, Teresa; Saunders, Darren N.; Wong, Jason; Xian, Jian; Bally, Marcel B.; Brenton, James D.; Brown, Grant W.; Shah, Sohrab P.; Cescon, David; Mak, Tak W.; Caldas, Carlos; Stirling, Peter C.; Hieter, Phil; Balasubramanian, Shankar; Aparicio, Samuel</p> <p>2017-01-01</p> <p><span class="hlt">G</span>-quadruplex DNAs form four-stranded helical structures and are proposed to play key roles in different cellular processes. Targeting <span class="hlt">G</span>-quadruplex DNAs for cancer treatment is a very promising prospect. Here, we show that CX-5461 is a <span class="hlt">G</span>-quadruplex stabilizer, with specific toxicity against BRCA deficiencies in cancer cells and polyclonal patient-derived xenograft models, including tumours resistant to PARP inhibition. Exposure to CX-5461, and its related drug CX-3543, blocks replication forks and <span class="hlt">induces</span> ssDNA gaps or <span class="hlt">breaks</span>. The BRCA and NHEJ pathways are required for the repair of CX-5461 and CX-3543-<span class="hlt">induced</span> DNA damage and failure to do so leads to lethality. These data strengthen the concept of <span class="hlt">G</span>4 targeting as a therapeutic approach, specifically for targeting HR and NHEJ deficient cancers and other tumours deficient for DNA damage repair. CX-5461 is now in advanced phase I clinical trial for patients with BRCA1/<span class="hlt">2</span> deficient tumours (Canadian trial, NCT02719977, opened May 2016). PMID:28211448</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28211448','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28211448"><span>CX-5461 is a DNA <span class="hlt">G</span>-quadruplex stabilizer with selective lethality in BRCA1/<span class="hlt">2</span> deficient tumours.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Xu, Hong; Di Antonio, Marco; McKinney, Steven; Mathew, Veena; Ho, Brandon; O'Neil, Nigel J; Santos, Nancy Dos; Silvester, Jennifer; Wei, Vivien; Garcia, Jessica; Kabeer, Farhia; Lai, Daniel; Soriano, Priscilla; Banáth, Judit; Chiu, Derek S; Yap, Damian; Le, Daniel D; Ye, Frank B; Zhang, Anni; Thu, Kelsie; Soong, John; Lin, Shu-Chuan; Tsai, Angela Hsin Chin; Osako, Tomo; Algara, Teresa; Saunders, Darren N; Wong, Jason; Xian, Jian; Bally, Marcel B; Brenton, James D; Brown, Grant W; Shah, Sohrab P; Cescon, David; Mak, Tak W; Caldas, Carlos; Stirling, Peter C; Hieter, Phil; Balasubramanian, Shankar; Aparicio, Samuel</p> <p>2017-02-17</p> <p><span class="hlt">G</span>-quadruplex DNAs form four-stranded helical structures and are proposed to play key roles in different cellular processes. Targeting <span class="hlt">G</span>-quadruplex DNAs for cancer treatment is a very promising prospect. Here, we show that CX-5461 is a <span class="hlt">G</span>-quadruplex stabilizer, with specific toxicity against BRCA deficiencies in cancer cells and polyclonal patient-derived xenograft models, including tumours resistant to PARP inhibition. Exposure to CX-5461, and its related drug CX-3543, blocks replication forks and <span class="hlt">induces</span> ssDNA gaps or <span class="hlt">breaks</span>. The BRCA and NHEJ pathways are required for the repair of CX-5461 and CX-3543-<span class="hlt">induced</span> DNA damage and failure to do so leads to lethality. These data strengthen the concept of <span class="hlt">G</span>4 targeting as a therapeutic approach, specifically for targeting HR and NHEJ deficient cancers and other tumours deficient for DNA damage repair. CX-5461 is now in advanced phase I clinical trial for patients with BRCA1/<span class="hlt">2</span> deficient tumours (Canadian trial, NCT02719977, opened May 2016).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26074399','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26074399"><span>High levels of Porphyromonas gingivalis-<span class="hlt">induced</span> immunoglobulin <span class="hlt">G</span><span class="hlt">2</span> are associated with lower high-density lipoprotein levels in chronic periodontitis.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ardila, Carlos M; Guzmán, Isabel C</p> <p>2016-11-01</p> <p>To investigate the association between the presence of Porphyromonas gingivalis-<span class="hlt">induced</span> immunoglobulin <span class="hlt">G</span> antibodies and the high-density lipoprotein (HDL) level. A total of 108 individuals were examined. The presence of P. gingivalis was detected using primers designed to target the 16S rRNA gene sequence. Peripheral blood was collected from each subject to determine the levels of P. gingivalis-<span class="hlt">induced</span> Ig<span class="hlt">G</span>1 and Ig<span class="hlt">G</span><span class="hlt">2</span> serum antibodies. The HDL levels were determined using fully enzymatic methods. A higher proportion of periodontitis patients had high levels of P. gingivalis-<span class="hlt">induced</span> Ig<span class="hlt">G</span>1 and Ig<span class="hlt">G</span><span class="hlt">2</span>, and the proportion of subjects with a HDL level of < 35 md/dL was higher in the group of chronic periodontitis patients. In the unadjusted regression model, the presence of high levels of P. gingivalis-<span class="hlt">induced</span> Ig<span class="hlt">G</span><span class="hlt">2</span> was associated with a HDL level of < 35 md/dL. The adjusted model indicated that periodontitis patients with high levels of P. gingivalis-<span class="hlt">induced</span> Ig<span class="hlt">G</span><span class="hlt">2</span> showed 3.<span class="hlt">2</span> more chances of having pathological HDL levels (odds ratio = 3.<span class="hlt">2</span>, 95% confidence interval = 1.<span class="hlt">2</span>-9.8). High levels of P. gingivalis-<span class="hlt">induced</span> Ig<span class="hlt">G</span><span class="hlt">2</span> were associated with low HDL concentrations in patients with periodontitis, which suggests that the response of the host to periodontal infection may play an important role in the pathogenesis of cardiovascular diseases. © 2015 Wiley Publishing Asia Pty Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27984746','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27984746"><span>Mammalian RAD52 Functions in <span class="hlt">Break-Induced</span> Replication Repair of Collapsed DNA Replication Forks.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sotiriou, Sotirios K; Kamileri, Irene; Lugli, Natalia; Evangelou, Konstantinos; Da-Ré, Caterina; Huber, Florian; Padayachy, Laura; Tardy, Sebastien; Nicati, Noemie L; Barriot, Samia; Ochs, Fena; Lukas, Claudia; Lukas, Jiri; Gorgoulis, Vassilis G; Scapozza, Leonardo; Halazonetis, Thanos D</p> <p>2016-12-15</p> <p>Human cancers are characterized by the presence of oncogene-<span class="hlt">induced</span> DNA replication stress (DRS), making them dependent on repair pathways such as <span class="hlt">break-induced</span> replication (BIR) for damaged DNA replication forks. To better understand BIR, we performed a targeted siRNA screen for genes whose depletion inhibited <span class="hlt">G</span>1 to S phase progression when oncogenic cyclin E was overexpressed. RAD52, a gene dispensable for normal development in mice, was among the top hits. In cells in which fork collapse was <span class="hlt">induced</span> by oncogenes or chemicals, the Rad52 protein localized to DRS foci. Depletion of Rad52 by siRNA or knockout of the gene by CRISPR/Cas9 compromised restart of collapsed forks and led to DNA damage in cells experiencing DRS. Furthermore, in cancer-prone, heterozygous APC mutant mice, homozygous deletion of the Rad52 gene suppressed tumor growth and prolonged lifespan. We therefore propose that mammalian RAD52 facilitates repair of collapsed DNA replication forks in cancer cells. Copyright © 2016 The Author(s). Published by Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21255774-sulforaphane-induces-dna-double-strand-breaks-predominantly-repaired-homologous-recombination-pathway-human-cancer-cells','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21255774-sulforaphane-induces-dna-double-strand-breaks-predominantly-repaired-homologous-recombination-pathway-human-cancer-cells"><span>Sulforaphane <span class="hlt">induces</span> DNA double strand <span class="hlt">breaks</span> predominantly repaired by homologous recombination pathway in human cancer cells</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Sekine-Suzuki, Emiko; Graduate School of Advanced Integration Science, Chiba University, 1-33 Yayoi-cho, Inage-ku, Chiba 263-8522; Yu, Dong</p> <p>2008-12-12</p> <p>Cytotoxicity and DNA double strand <span class="hlt">breaks</span> (DSBs) were studied in HeLa cells treated with sulforaphane (SFN), a well-known chemo-preventive agent. Cell survival was impaired by SFN in a concentration and treatment time-dependent manner. Both constant field gel electrophoresis (CFGE) and {gamma}-H<span class="hlt">2</span>AX assay unambiguously indicated formation of DSBs by SFN, reflecting the cell survival data. These DSBs were predominantly processed by homologous recombination repair (HRR), judging from the SFN concentration-dependent manner of Rad51 foci formation. On the other hand, the phosphorylation of DNA-PKcs, a key non-homologous end joining (NHEJ) protein, was not observed by SFN treatment, suggesting that NHEJ may notmore » be involved in DSBs <span class="hlt">induced</span> by this chemical. <span class="hlt">G</span><span class="hlt">2</span>/M arrest by SFN, a typical response for cells exposed to ionizing radiation was also observed. Our new data indicate the clear induction of DSBs by SFN and a useful anti-tumor aspect of SFN through the induction of DNA DSBs.« less</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li class="active"><span>8</span></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_8 --> <div id="page_9" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li class="active"><span>9</span></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="161"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28009302','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28009302"><span>RPA Stabilization of Single-Stranded DNA Is Critical for <span class="hlt">Break-Induced</span> Replication.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ruff, Patrick; Donnianni, Roberto A; Glancy, Eleanor; Oh, Julyun; Symington, Lorraine S</p> <p>2016-12-20</p> <p>DNA double-strand <span class="hlt">breaks</span> (DSBs) are cytotoxic lesions that must be accurately repaired to maintain genome stability. Replication protein A (RPA) plays an important role in homology-dependent repair of DSBs by protecting the single-stranded DNA (ssDNA) intermediates formed by end resection and by facilitating Rad51 loading. We found that hypomorphic mutants of RFA1 that support intra-chromosomal homologous recombination are profoundly defective for repair processes involving long tracts of DNA synthesis, in particular <span class="hlt">break-induced</span> replication (BIR). The BIR defects of the rfa1 mutants could be partially suppressed by eliminating the Sgs1-Dna<span class="hlt">2</span> resection pathway, suggesting that Dna<span class="hlt">2</span> nuclease attacks the ssDNA formed during end resection when not fully protected by RPA. Overexpression of Rad51 was also found to suppress the rfa1 BIR defects. We suggest that Rad51 binding to the ssDNA formed by excessive end resection and during D-loop migration can partially compensate for dysfunctional RPA. Copyright © 2016 The Author(s). Published by Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3356935','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3356935"><span>Ethanol Extract of Dianthus chinensis L. <span class="hlt">Induces</span> Apoptosis in Human Hepatocellular Carcinoma Hep<span class="hlt">G</span><span class="hlt">2</span> Cells In Vitro</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Nho, Kyoung Jin; Chun, Jin Mi; Kim, Ho Kyoung</p> <p>2012-01-01</p> <p>Dianthus chinensis L. is used to treat various diseases including cancer; however, the molecular mechanism by which the ethanol extract of Dianthus chinensis L. (EDCL) <span class="hlt">induces</span> apoptosis is unknown. In this study, the apoptotic effects of EDCL were investigated in human Hep<span class="hlt">G</span><span class="hlt">2</span> hepatocellular carcinoma cells. Treatment with EDCL significantly inhibited cell growth in a concentration- and time-dependent manner by <span class="hlt">inducing</span> apoptosis. This induction was associated with chromatin condensation, activation of caspases, and cleavage of poly (ADP-ribose) polymerase protein. However, apoptosis <span class="hlt">induced</span> by EDCL was attenuated by caspase inhibitor, indicating an important role for caspases in EDCL responses. Furthermore, EDCL did not alter the expression of bax in Hep<span class="hlt">G</span><span class="hlt">2</span> cells but did selectively downregulate the expression of bcl-<span class="hlt">2</span> and bcl-xl, resulting in an increase in the ratio of bax:bcl-<span class="hlt">2</span> and bax:bcl-xl. These results support a mechanism whereby EDCL <span class="hlt">induces</span> apoptosis through the mitochondrial pathway and caspase activation in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. PMID:22645629</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16427178','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16427178"><span>Physalis angulata <span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span>/M phase arrest in human breast cancer cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hsieh, Wen-Tsong; Huang, Kuan-Yuh; Lin, Hui-Yi; Chung, Jing-Gung</p> <p>2006-07-01</p> <p>Physalis angulata (PA) is employed in herbal medicine around the world. It is used to treat diabetes, hepatitis, asthma and malaria in Taiwan. We have evaluated PA as a cancer chemopreventive agent in vitro by studying the role of PA in regulation of proliferation, cell cycle and apoptosis in human breast cancer cell lines. PA inhibited cell proliferation and <span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span>/M arrest and apoptosis in human breast cancer MAD-MB 231 and MCF-7 cell lines. In this study, under treatment with various concentrations of PA in MDA-MB 231 cell line, we checked mRNA levels for cyclin A and cyclin B1 and the protein levels of cyclin A and cyclin B1, Cdc<span class="hlt">2</span> (cyclin-dependent kinases), p21(waf1/cip1) and P27(Kip1) (cyclin-dependent kinase inhibitors), Cdc25C, Chk<span class="hlt">2</span> and Wee1 kinase (cyclin-dependent kinase relative factors) in cell cycle <span class="hlt">G</span><span class="hlt">2</span>/M phase. From those results, we determined that PA arrests MDA-MB 231 cells at the <span class="hlt">G</span><span class="hlt">2</span>/M phase by (i) inhibiting synthesis or stability of mRNA and their downstream protein levels of cyclin A and cyclin B1, (ii) increasing p21(waf1/cip1) and P27(kip1) levels, (iii) increasing Chk<span class="hlt">2</span>, thus causing an increase in Cdc25C phosphorylation/inactivation and <span class="hlt">inducing</span> a decrease in Cdc<span class="hlt">2</span> levels and an increase in Wee1 level. According to the results obtained, PA appears to possess anticarcinogenic properties; these results suggest that the effect of PA on the levels of phosphorylated/inactivated Cdc25C are mediated by Chk<span class="hlt">2</span> activation, at least in part, via p21(waf1/cip1) and P27(kip1) cyclin-dependent kinase inhibitors pathway to arrest cells at <span class="hlt">G</span><span class="hlt">2</span>/M phase in breast cancer carcinoma cells.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12646262','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12646262"><span>Loss of p53 <span class="hlt">induces</span> M-phase retardation following <span class="hlt">G</span><span class="hlt">2</span> DNA damage checkpoint abrogation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Minemoto, Yuzuru; Uchida, Sanae; Ohtsubo, Motoaki; Shimura, Mari; Sasagawa, Toshiyuki; Hirata, Masato; Nakagama, Hitoshi; Ishizaka, Yukihito; Yamashita, Katsumi</p> <p>2003-04-01</p> <p>Most cell lines that lack functional p53 protein are arrested in the <span class="hlt">G</span><span class="hlt">2</span> phase of the cell cycle due to DNA damage. When the <span class="hlt">G</span><span class="hlt">2</span> checkpoint is abrogated, these cells are forced into mitotic catastrophe. A549 lung adenocarcinoma cells, in which p53 was eliminated with the HPV16 E6 gene, exhibited efficient arrest in the <span class="hlt">G</span><span class="hlt">2</span> phase when treated with adriamycin. Administration of caffeine to <span class="hlt">G</span><span class="hlt">2</span>-arrested cells <span class="hlt">induced</span> a drastic change in cell phenotype, the nature of which depended on the status of p53. Flow cytometric and microscopic observations revealed that cells that either contained or lacked p53 resumed their cell cycles and entered mitosis upon caffeine treatment. However, transit to the M phase was slower in p53-negative cells than in p53-positive cells. Consistent with these observations, CDK1 activity was maintained at high levels, along with stable cyclin B1, in p53-negative cells. The addition of butyrolactone I, which is an inhibitor of CDK1 and CDK<span class="hlt">2</span>, to the p53-negative cells reduced the floating round cell population and <span class="hlt">induced</span> the disappearance of cyclin B1. These results suggest a relationship between the p53 pathway and the ubiquitin-mediated degradation of mitotic cyclins and possible cross-talk between the <span class="hlt">G</span><span class="hlt">2</span>-DNA damage checkpoint and the mitotic checkpoint.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1410605-dynamical-time-reversal-symmetry-breaking-photo-induced-chiral-spin-liquids-frustrated-mott-insulators','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1410605-dynamical-time-reversal-symmetry-breaking-photo-induced-chiral-spin-liquids-frustrated-mott-insulators"><span>Dynamical time-reversal symmetry <span class="hlt">breaking</span> and photo-<span class="hlt">induced</span> chiral spin liquids in frustrated Mott insulators</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Claassen, Martin; Jiang, Hong -Chen; Moritz, Brian; ...</p> <p>2017-10-30</p> <p>The search for quantum spin liquids in frustrated quantum magnets recently has enjoyed a surge of interest, with various candidate materials under intense scrutiny. However, an experimental confirmation of a gapped topological spin liquid remains an open question. Here, we show that circularly polarized light can provide a knob to drive frustrated Mott insulators into a chiral spin liquid, realizing an elusive quantum spin liquid with topological order. We find that the dynamics of a driven Kagome Mott insulator is well-captured by an effective Floquet spin model, with heating strongly suppressed, <span class="hlt">inducing</span> a scalar spin chirality S i · (Smore » j × S k) term which dynamically <span class="hlt">breaks</span> time-reversal while preserving SU(<span class="hlt">2</span>) spin symmetry. We fingerprint the transient phase diagram and find a stable photo-<span class="hlt">induced</span> chiral spin liquid near the equilibrium state. Furthermore, the results presented suggest employing dynamical symmetry <span class="hlt">breaking</span> to engineer quantum spin liquids and access elusive phase transitions that are not readily accessible in equilibrium.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21254192-sub-mssm-theory-motivated-model-particle-physics','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21254192-sub-mssm-theory-motivated-model-particle-physics"><span><span class="hlt">G</span>{sub <span class="hlt">2</span>}-MSSM: An M theory motivated model of particle physics</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Acharya, Bobby S.; Bobkov, Konstantin; Kane, Gordon L.</p> <p>2008-09-15</p> <p>We continue our study of the low energy implications of M theory vacua on <span class="hlt">G</span>{sub <span class="hlt">2</span>}-manifolds, undertaken in B. S. Acharya, K. Bobkov, <span class="hlt">G</span>. L. Kane, P. Kumar, and J. Shao, Phys. Rev. D 76, 126010 (2007); B. Acharya, K. Bobkov, <span class="hlt">G</span>. Kane, P. Kumar, and D. Vaman, Phys. Rev. Lett. 97, 191601 (2006), where it was shown that the moduli can be stabilized and a TeV scale generated, with the Planck scale as the only dimensionful input. A well-motivated phenomenological model, the <span class="hlt">G</span>{sub <span class="hlt">2</span>}-MSSM, can be naturally defined within the above framework. In this paper, we study some ofmore » the important phenomenological features of the <span class="hlt">G</span>{sub <span class="hlt">2</span>}-MSSM. In particular, the soft supersymmetry <span class="hlt">breaking</span> parameters and the superpartner spectrum are computed. The <span class="hlt">G</span>{sub <span class="hlt">2</span>}-MSSM generically gives rise to light gauginos and heavy scalars with wino lightest supersymmetric particles when one tunes the cosmological constant. Electroweak symmetry <span class="hlt">breaking</span> is present but fine-tuned. The <span class="hlt">G</span>{sub <span class="hlt">2</span>}-MSSM is also naturally consistent with precision gauge coupling unification. The phenomenological consequences for cosmology and collider physics of the <span class="hlt">G</span>{sub <span class="hlt">2</span>}-MSSM will be reported in more detail soon.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5370529','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5370529"><span>Black soybean seed coat polyphenols prevent AAPH-<span class="hlt">induced</span> oxidative DNA-damage in Hep<span class="hlt">G</span><span class="hlt">2</span> cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Yoshioka, Yasukiyo; Li, Xiu; Zhang, Tianshun; Mitani, Takakazu; Yasuda, Michiko; Nanba, Fumio; Toda, Toshiya; Yamashita, Yoko; Ashida, Hitoshi</p> <p>2017-01-01</p> <p>Black soybean seed coat extract (BE), which contains abundant polyphenols such as procyanidins, cyanidin 3-glucoside, (+)-catechin, and (−)­epicatechin, has been reported on health beneficial functions such as antioxidant activity, anti-inflammatory, anti-obesity, and anti-diabetic activities. In this study, we investigated that prevention of BE and its polyphenols on <span class="hlt">2,2</span>'-azobis(<span class="hlt">2</span>-methylpropionamide) dihydrochloride (AAPH)-<span class="hlt">induced</span> oxidative DNA damage, and found that these polyphenols inhibited AAPH-<span class="hlt">induced</span> formation of 8-hydroxy-<span class="hlt">2</span>'-deoxyguanosine (8-OHd<span class="hlt">G</span>) as a biomarker for oxidative DNA damage in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Under the same conditions, these polyphenols also inhibited AAPH-<span class="hlt">induced</span> accumulation of reactive oxygen species (ROS) in the cells. Inhibition of ROS accumulation was observed in both cytosol and nucleus. It was confirmed that these polyphenols inhibited formation of AAPH radical using oxygen radical absorbance capacity assay under the cell-free conditions. These results indicate that polyphenols in BE inhibit free radical-<span class="hlt">induced</span> oxidative DNA damages by their potent antioxidant activity. Thus, BE is an effective food material for prevention of oxidative stress and oxidative DNA damages. PMID:28366989</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24662601','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24662601"><span>Role of 6-shogaol in tert -butyl hydroperoxide-<span class="hlt">induced</span> apoptosis of Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kim, Sang Chan; Lee, Jong Rok; Park, Sook Jahr</p> <p>2014-01-01</p> <p>The aim of this study was to investigate the protective effects of 6-shogaol on tert-butyl hydroperoxide (tBHP)-<span class="hlt">induced</span> oxidative stress leading to apoptosis in human hepatoma cell line Hep<span class="hlt">G</span><span class="hlt">2</span>. The cells were exposed to tBHP (100 μmol/l) after pretreatment with 6-shogaol (<span class="hlt">2</span>.5 and 5 μmol/l), and then cell viability was measured. 6-Shogaol fully prevented Hep<span class="hlt">G</span><span class="hlt">2</span> cell death caused by tBHP. Treatment of tBHP resulted in apoptotic cell death as assessed by TUNEL assay and the expression of apoptosis regulator proteins, Bcl-<span class="hlt">2</span> family, caspases and cytochrome c. Cells treated with 6-shogaol showed rapid reduction of apoptosis by restoring these markers of apoptotic cells. In addition, 6-shogaol significantly recovered disruption of mitochondrial membrane potential as a start sign of hepatic apoptosis <span class="hlt">induced</span> by oxidative stress. In line with this observation, antioxidative 6-shogaol inhibited generation of reactive oxygen species and depletion of reduced glutathione in tBHP-stimulated Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Taken together, these results for the first time showed antioxidative and antiapoptotic activities of 6-shogaol in tBHP-treated hepatoma Hep<span class="hlt">G</span><span class="hlt">2</span> cells, suggesting that 6-shogaol could be beneficial in hepatic disorders caused by oxidative stress. © 2014 S. Karger AG, Basel.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29723526','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29723526"><span>AFP<span class="hlt">2</span> as the novel regulator <span class="hlt">breaks</span> high-temperature-<span class="hlt">induced</span> seeds secondary dormancy through ABI5 and SOM in Arabidopsis thaliana.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chang, Guanxiao; Wang, Chuntao; Kong, Xiangxiang; Chen, Qian; Yang, Yongping; Hu, Xiangyang</p> <p>2018-06-18</p> <p>Imbibed seeds monitor environmental and endogenous signals to <span class="hlt">break</span> dormancy and initiate growth under appropriate conditions. In Arabidopsis thaliana, high temperature (HT) <span class="hlt">induces</span> secondary seed dormancy, but the underlying mechanism remains unclear. In this study, we found that the abi5-1 mutant was insensitive to high temperature, whereas plants overexpressing ABI5 displayed sensitivity. We then identified ABA-insensitive five-binding protein <span class="hlt">2</span> (AFP<span class="hlt">2</span>), which interacts with ABI5 and is involved in HT-<span class="hlt">induced</span> secondary seed dormancy. Under HT stress, the loss-of-function afp<span class="hlt">2</span> mutant showed lower seeds germination frequency, reversely, AFP<span class="hlt">2</span> overexpressing lines (OE-AFP<span class="hlt">2</span>) showed high germination frequency. Similar to the abi5 mutant, the crossed OE-AFP<span class="hlt">2</span> abi5 or afp<span class="hlt">2</span> abi5 lines showed high germination under HT, suggesting that ABI5 is epistatic to AFP<span class="hlt">2</span>. SOM is reported to negatively regulate seeds germination by altering GA/ABA metabolism, here we found that AFP<span class="hlt">2</span> and ABI5 altered SOM transcription. Specifically, overexpressing AFP<span class="hlt">2</span> suppressed SOM transcription, resulting in high expression of GA biosynthesis-related genes and low expression of ABA biosynthesis-related genes, ultimately promoting seed germination under HT. Thus, our data demonstrate that AFP<span class="hlt">2</span> is a novel regulator to control HT-<span class="hlt">induced</span> secondary seed dormancy through ABI5 and SOM. Copyright © 2018 Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16238096','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16238096"><span>Roles of nibrin and AtM/ATR kinases on the <span class="hlt">G</span><span class="hlt">2</span> checkpoint under endogenous or radio-<span class="hlt">induced</span> DNA damage.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Marcelain, Katherine; De La Torre, Consuelo; González, Patricio; Pincheira, Juana</p> <p>2005-01-01</p> <p>Checkpoint response to DNA damage involves the activation of DNA repair and <span class="hlt">G</span><span class="hlt">2</span> lengthening subpathways. The roles of nibrin (NBS1) and the ATM/ATR kinases in the <span class="hlt">G</span><span class="hlt">2</span> DNA damage checkpoint, evoked by endogenous and radio-<span class="hlt">induced</span> DNA damage, were analyzed in control, A-T and NBS lymphoblast cell lines. Short-term responses to <span class="hlt">G</span><span class="hlt">2</span> treatments were evaluated by recording changes in the yield of chromosomal aberrations in the ensuing mitosis, due to <span class="hlt">G</span><span class="hlt">2</span> checkpoint adaptation, and also in the duration of <span class="hlt">G</span><span class="hlt">2</span> itself. The role of ATM/ATR in the <span class="hlt">G</span><span class="hlt">2</span> checkpoint pathway repairing chromosomal aberrations was unveiled by caffeine inhibition of both kinases in <span class="hlt">G</span><span class="hlt">2</span>. In the control cell lines, nibrin and ATM cooperated to provide optimum <span class="hlt">G</span><span class="hlt">2</span> repair for endogenous DNA damage. In the A-T cells, ATR kinase substituted successfully for ATM, even though no <span class="hlt">G</span><span class="hlt">2</span> lengthening occurred. X-ray irradiation (0.4 Gy) in <span class="hlt">G</span><span class="hlt">2</span> increased chromosomal aberrations and lengthened <span class="hlt">G</span><span class="hlt">2</span>, in both mutant and control cells. However, the repair of radio-<span class="hlt">induced</span> DNA damage took place only in the controls. It was associated with nibrin-ATM interaction, and ATR did not substitute for ATM. The absence of nibrin prevented the repair of both endogenous and radio-<span class="hlt">induced</span> DNA damage in the NBS cells and partially affected the induction of <span class="hlt">G</span><span class="hlt">2</span> lengthening.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhPl...25a2512L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhPl...25a2512L"><span>Mode structure symmetry <span class="hlt">breaking</span> of energetic particle driven beta-<span class="hlt">induced</span> Alfvén eigenmode</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lu, Z. X.; Wang, X.; Lauber, Ph.; Zonca, F.</p> <p>2018-01-01</p> <p>The mode structure symmetry <span class="hlt">breaking</span> of energetic particle driven Beta-<span class="hlt">induced</span> Alfvén Eigenmode (BAE) is studied based on global theory and simulation. The weak coupling formula gives a reasonable estimate of the local eigenvalue compared with global hybrid simulation using XHMGC. The non-perturbative effect of energetic particles on global mode structure symmetry <span class="hlt">breaking</span> in radial and parallel (along B) directions is demonstrated. With the contribution from energetic particles, two dimensional (radial and poloidal) BAE mode structures with symmetric/asymmetric tails are produced using an analytical model. It is demonstrated that the symmetry <span class="hlt">breaking</span> in radial and parallel directions is intimately connected. The effects of mode structure symmetry <span class="hlt">breaking</span> on nonlinear physics, energetic particle transport, and the possible insight for experimental studies are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20661717','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20661717"><span>Silencing of cytosolic NADP+-dependent isocitrate dehydrogenase gene enhances ethanol-<span class="hlt">induced</span> toxicity in Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yang, Eun Sun; Lee, Su-Min; Park, Jeen-Woo</p> <p>2010-07-01</p> <p>It has been shown that acute and chronic alcohol administrations increase the production of reactive oxygen species, lower cellular antioxidant levels and enhance oxidative stress in many tissues. We recently reported that cytosolic NADP(+)-dependent isocitrate dehydrogenase (IDPc) functions as an antioxidant enzyme by supplying NADPH to the cytosol. Upon exposure to ethanol, IDPc was susceptible to the loss of its enzyme activity in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Transfection of Hep<span class="hlt">G</span><span class="hlt">2</span> cells with an IDPc small interfering RNA noticeably downregulated IDPc and enhanced the cells' vulnerability to ethanol-<span class="hlt">induced</span> cytotoxicity. Our results suggest that suppressing the expression of IDPc enhances ethanol-<span class="hlt">induced</span> toxicity in Hep<span class="hlt">G</span><span class="hlt">2</span> cells by further disruption of the cellular redox status.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1985NuPhB.256..130M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1985NuPhB.256..130M"><span>Effects of grand unification interactions on weak symmetry <span class="hlt">breaking</span> in supergravity theories</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Moxhay, Peter; Yamamoto, Katsuji</p> <p></p> <p>Possible effects of grand unification interactions on SU(<span class="hlt">2</span>) × U(1) <span class="hlt">breaking</span> are investigated by explicitly considering a supersymmetric SU(5) model coupled to N = 1 supergravity. Some remarkable features concerning the effects of renormalization on the effective soft supersymmetry <span class="hlt">breaking</span> terms of SU(5) in the GUT region MP - MG are clarified, which are relevant for determining the SU(3) × SU(<span class="hlt">2</span>) × U(1) theory below MG. In particular, the (mass) <span class="hlt">2</span> of the Higgs doublets, <span class="hlt">g</span> Hm <span class="hlt">g</span><span class="hlt">2</span>and <span class="hlt">g</span> overlineHm <span class="hlt">g</span><span class="hlt">2</span>, might become significantly small at M <span class="hlt">G</span> (<span class="hlt">g</span> H ⋍ <span class="hlt">g</span> overlineH ≈ 0.1) through the effect of SU(5) couplings such as overlineHø EH . Then, <span class="hlt">g</span>H can rather easily become negative below MG, so as to realize SU(<span class="hlt">2</span>) × U(1) <span class="hlt">breaking</span> naturally even for the "diet" top quark case ( mt ≈ 40 GeV). On the other hand, if <span class="hlt">g</span> H ⋍ <span class="hlt">g</span> overlineH ⋍ 1 at M <span class="hlt">G</span> by neglecting the grand unification interactions, some careful tuning of μ32/ mg<span class="hlt">2</span> is required with an accuracy ⪅10 -<span class="hlt">2</span> to achieve SU(<span class="hlt">2</span>) × U(1) <span class="hlt">breaking</span> with "diet" top quark, though a mass term μ 32( overlineHH) may be present.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28627253','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28627253"><span>Protective effects of quercetin on nicotine <span class="hlt">induced</span> oxidative stress in 'Hep<span class="hlt">G</span><span class="hlt">2</span> cells'.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yarahmadi, Amir; Zal, Fatemeh; Bolouki, Ayeh</p> <p>2017-10-01</p> <p>Nicotine is a natural component of tobacco plants and is responsible for the addictive properties of tobacco. Nicotine has been recognized to result in oxidative stress by <span class="hlt">inducing</span> the generation of reactive oxygen species (ROS). The purpose of this work was to estimate the hepatotoxicity effect of nicotine on viability and on antioxidant defense system in cultures of Hep<span class="hlt">G</span><span class="hlt">2</span> cell line and the other hand, ameliorative effect of quercetin (Q) as an antioxidant was analyzed. Nicotine <span class="hlt">induced</span> concentration dependent loss in Hep<span class="hlt">G</span><span class="hlt">2</span> cell line viability. The results indicated that nicotine decreased activity of superoxide dismutase (SOD) and glutathione reductase (GR) and increased activities of catalase (CAT) and glutathione peroxidase (GPx) and glutathione (GSH) content in the Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Q significantly increased activity of SOD, GR and GSH content and decreased activity of GPX in nicotine + Q groups. Our data demonstrate that Q plays a protective role against the imbalance elicited by nicotine between the production of free radicals and antioxidant defense systems, and suggest that administration of this antioxidant may find clinical application where cellular damage is a consequence of ROS.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28327192','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28327192"><span>DNA double-strand <span class="hlt">breaks</span> in human <span class="hlt">induced</span> pluripotent stem cell reprogramming and long-term in vitro culturing.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Simara, Pavel; Tesarova, Lenka; Rehakova, Daniela; Matula, Pavel; Stejskal, Stanislav; Hampl, Ales; Koutna, Irena</p> <p>2017-03-21</p> <p>Human <span class="hlt">induced</span> pluripotent stem cells (hiPSCs) play roles in both disease modelling and regenerative medicine. It is critical that the genomic integrity of the cells remains intact and that the DNA repair systems are fully functional. In this article, we focused on the detection of DNA double-strand <span class="hlt">breaks</span> (DSBs) by phosphorylated histone H<span class="hlt">2</span>AX (known as γH<span class="hlt">2</span>AX) and p53-binding protein 1 (53BP1) in three distinct lines of hiPSCs, their source cells, and one line of human embryonic stem cells (hESCs). We measured spontaneously occurring DSBs throughout the process of fibroblast reprogramming and during long-term in vitro culturing. To assess the variations in the functionality of the DNA repair system among the samples, the number of DSBs <span class="hlt">induced</span> by γ-irradiation and the decrease over time was analysed. The foci number was detected by fluorescence microscopy separately for the <span class="hlt">G</span>1 and S/<span class="hlt">G</span><span class="hlt">2</span> cell cycle phases. We demonstrated that fibroblasts contained a low number of non-replication-related DSBs, while this number increased after reprogramming into hiPSCs and then decreased again after long-term in vitro passaging. The artificial induction of DSBs revealed that the repair mechanisms function well in the source cells and hiPSCs at low passages, but fail to recognize a substantial proportion of DSBs at high passages. Our observations suggest that cellular reprogramming increases the DSB number but that the repair mechanism functions well. However, after prolonged in vitro culturing of hiPSCs, the repair capacity decreases.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1783675','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1783675"><span>Roles of the sister <span class="hlt">chromatid</span> cohesion apparatus in gene expression, development, and human syndromes</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Dorsett, Dale</p> <p>2006-01-01</p> <p>The sister <span class="hlt">chromatid</span> cohesion apparatus mediates physical pairing of duplicated chromosomes. This pairing is essential for appropriate distribution of chromosomes into the daughter cells upon cell division. Recent evidence shows that the cohesion apparatus, which is a significant structural component of chromosomes during interphase, also affects gene expression and development. The Cornelia de Lange (CdLS) and Roberts/SC phocomelia (RBS/SC) genetic syndromes in humans are caused by mutations affecting components of the cohesion apparatus. Studies in Drosophila suggest that effects on gene expression are most likely responsible for developmental alterations in CdLS. Effects on <span class="hlt">chromatid</span> cohesion are apparent in RBS/SC syndrome, but data from yeast and Drosophila point to the likelihood that changes in expression of genes located in heterochromatin could contribute to the developmental deficits. PMID:16819604</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17390723','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17390723"><span>Determination and analysis of site-specific 125I decay-<span class="hlt">induced</span> DNA double-strand <span class="hlt">break</span> end-group structures.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Datta, Kamal; Weinfeld, Michael; Neumann, Ronald D; Winters, Thomas A</p> <p>2007-02-01</p> <p>End groups contribute to the structural complexity of radiation-<span class="hlt">induced</span> DNA double-strand <span class="hlt">breaks</span> (DSBs). As such, end-group structures may affect a cell's ability to repair DSBs. The 3'-end groups of strand <span class="hlt">breaks</span> caused by gamma radiation, or oxidative processes, under oxygenated aqueous conditions have been shown to be distributed primarily between 3'-phosphoglycolate and 3'-phosphate, with 5'-phosphate ends in both cases. In this study, end groups of the high-LET-like DSBs caused by 125I decay were investigated. Site-specific DNA double-strand <span class="hlt">breaks</span> were produced in plasmid pTC27 in the presence or absence of <span class="hlt">2</span> M DMSO by 125I-labeled triplex-forming oligonucleotide targeting. End-group structure was assessed enzymatically as a function of the DSB end to serve as a substrate for ligation and various forms of end labeling. Using this approach, we have demonstrated 3'-hydroxyl (3'-OH) and 3'-phosphate (3'-P) end groups and 5'-ends (> or = 42%) terminated by phosphate. A 32P postlabeling assay failed to detect 3'-phosphoglycolate in a restriction fragment terminated by the 125I-<span class="hlt">induced</span> DNA double-strand <span class="hlt">break</span>, and this is likely due to restricted oxygen diffusion during irradiation as a frozen aqueous solution. Even so, end-group structure and relative distribution varied as a function of the free radical scavenging capacity of the irradiation buffer.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3662696','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3662696"><span>Staphylococcus aureus-<span class="hlt">Induced</span> <span class="hlt">G</span><span class="hlt">2</span>/M Phase Transition Delay in Host Epithelial Cells Increases Bacterial Infective Efficiency</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Almeida, Sintia; Legembre, Patrick; Edmond, Valérie; Azevedo, Vasco; Miyoshi, Anderson; Even, Sergine; Taieb, Frédéric; Arlot-Bonnemains, Yannick; Le Loir, Yves; Berkova, Nadia</p> <p>2013-01-01</p> <p>Staphylococcus aureus is a highly versatile, opportunistic pathogen and the etiological agent of a wide range of infections in humans and warm-blooded animals. The epithelial surface is its principal site of colonization and infection. In this work, we investigated the cytopathic effect of S. aureus strains from human and animal origins and their ability to affect the host cell cycle in human HeLa and bovine MAC-T epithelial cell lines. S. aureus invasion slowed down cell proliferation and <span class="hlt">induced</span> a cytopathic effect, resulting in the enlargement of host cells. A dramatic decrease in the number of mitotic cells was observed in the infected cultures. Flow cytometry analysis revealed an S. aureus-<span class="hlt">induced</span> delay in the <span class="hlt">G</span><span class="hlt">2</span>/M phase transition in synchronous HeLa cells. This delay required the presence of live S. aureus since the addition of the heat-killed bacteria did not alter the cell cycle. The results of Western blot experiments showed that the <span class="hlt">G</span><span class="hlt">2</span>/M transition delay was associated with the accumulation of inactive cyclin-dependent kinase Cdk1, a key <span class="hlt">inducer</span> of mitosis entry, and with the accumulation of unphosphorylated histone H3, which was correlated with a reduction of the mitotic cell number. Analysis of S. aureus proliferation in asynchronous, <span class="hlt">G</span>1- and <span class="hlt">G</span><span class="hlt">2</span>-phase-enriched HeLa cells showed that the <span class="hlt">G</span><span class="hlt">2</span> phase was preferential for bacterial infective efficiency, suggesting that the <span class="hlt">G</span><span class="hlt">2</span> phase delay may be used by S. aureus for propagation within the host. Taken together, our results divulge the potential of S. aureus in the subversion of key cellular processes such as cell cycle progression, and shed light on the biological significance of S. aureus-<span class="hlt">induced</span> host cell cycle alteration. PMID:23717407</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005GeoRL..3217807J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005GeoRL..3217807J"><span>Wave <span class="hlt">breaking</span> <span class="hlt">induced</span> surface wakes and jets observed during a bora event</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jiang, Qingfang; Doyle, James D.</p> <p>2005-09-01</p> <p>An observational and modeling study of a bora event that occurred during the field phase of the Mesoscale Alpine Programme is presented. Research aircraft in-situ measurements and airborne remote-sensing observations indicate the presence of strong low-level wave <span class="hlt">breaking</span> and alternating surface wakes and jets along the Croatian coastline over the Adriatic Sea. The observed features are well captured by a high-resolution COAMPS simulation. Analysis of the observations and modeling results indicate that the long-extending wakes above the boundary layer are <span class="hlt">induced</span> by dissipation associated with the low-level wave <span class="hlt">breaking</span>, which locally tends to accelerate the boundary layer flow beneath the <span class="hlt">breaking</span>. Farther downstream of the high peaks, a hydraulic jump occurs in the boundary layer, which creates surface wakes. Downstream of lower-terrain (passes), the boundary layer flow stays strong, resembling supercritical flow.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16756487','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16756487"><span><span class="hlt">Break-induced</span> replication and recombinational telomere elongation in yeast.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>McEachern, Michael J; Haber, James E</p> <p>2006-01-01</p> <p>When a telomere becomes unprotected or if only one end of a chromosomal double-strand <span class="hlt">break</span> succeeds in recombining with a template sequence, DNA can be repaired by a recombination-dependent DNA replication process termed <span class="hlt">break-induced</span> replication (BIR). In budding yeasts, there are two BIR pathways, one dependent on the Rad51 recombinase protein and one Rad51 independent; these two repair processes lead to different types of survivors in cells lacking the telomerase enzyme that is required for normal telomere maintenance. Recombination at telomeres is triggered by either excessive telomere shortening or disruptions in the function of telomere-binding proteins. Telomere elongation by BIR appears to often occur through a "roll and spread" mechanism. In this process, a telomeric circle produced by recombination at a dysfunctional telomere acts as a template for a rolling circle BIR event to form an elongated telomere. Additional BIR events can then copy the elongated sequence to all other telomeres.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li class="active"><span>9</span></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_9 --> <div id="page_10" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li class="active"><span>10</span></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="181"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28969111','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28969111"><span>Citral, A Monoterpene Protect Against High Glucose <span class="hlt">Induced</span> Oxidative Injury in Hep<span class="hlt">G</span><span class="hlt">2</span> Cell In Vitro-An Experimental Study.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Subramaniyan, Sri Devi; Natarajan, Ashok Kumar</p> <p>2017-08-01</p> <p>Diabetes mellitus, a major metabolic disorder associated with hyperglycaemia is one of the leading cause of death in many developed countries. However, use of natural phytochemicals have been proved to have a protective effect against oxidative damage. To investigate the effect of citral, a monoterpene on high glucose <span class="hlt">induced</span> cytotoxicity and oxidative stress in human hepatocellular liver carcinoma (Hep <span class="hlt">G</span><span class="hlt">2</span>) cell line. Cells were treated with 50 mM concentration of glucose for 24 hours incubation following citral (30 μM) was added to confluent Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Cell viability, Reactive Oxygen Species (ROS) generation, DNA damage, lipid peroxidation, antioxidants and Mitogen Activated Protein Kinases (MAPKs) signaling were assessed in citral and/or high glucose <span class="hlt">induced</span> Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Cells treated with glucose (50 mM), resulted in increased cytotoxicity, ROS generation, DNA damage, lipid peroxidation and depletion of enzymatic and non enzymatic antioxidants. In contrast, treatment with citral (30 μM) significantly decreased cell cytotoxicity, ROS generation, DNA damage, lipid peroxidation and increased antioxidants enzymes in high glucose <span class="hlt">induced</span> Hep<span class="hlt">G</span><span class="hlt">2</span> cells. In addition, the present study highlighted that high glucose treated cells showed increased expression of Extracellular Signal Regulated Protein Kinase-1 (ERK-1), c-Jun N-terminal Kinase (JNK) and p38 in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. On the other hand treatment with citral significantly suppressed the expression of ERK-1, JNK and p38 in high glucose <span class="hlt">induced</span> Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Citral protects against high glucose <span class="hlt">induced</span> oxidative stress through inhibiting ROS activated MAPK signaling pathway in Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/5120376-induction-sister-chromatid-exchange-preimplantation-mouse-embryos-vitro-sup-thymidine-ultraviolet-light-combination-caffeine','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/5120376-induction-sister-chromatid-exchange-preimplantation-mouse-embryos-vitro-sup-thymidine-ultraviolet-light-combination-caffeine"><span>Induction of sister <span class="hlt">chromatid</span> exchange in preimplantation mouse embryos in vitro by /sup 3/H-thymidine or ultraviolet light in combination with caffeine</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Mueller, W.U.S.; Spindle, A.</p> <p>1986-01-01</p> <p>Preimplantation mouse embryos were exposed in vitro to /sup 3/H-thymidine (25, 100, or 250 Bq/ml) or ultraviolet (UV) light (1.35 or 4.05 J/m<span class="hlt">2</span>), either alone or in combination with caffeine (1 mM with /sup 3/H-thymidine and 0.5 mM with UV light). Exposure to /sup 3/H-thymidine lasted for <span class="hlt">2</span> days, from the two-cell stage to the late morula/early blastocyst stage, and UV radiation was applied acutely at the late morula/early blastocyst stage. The effects were quantified by the sister <span class="hlt">chromatid</span> exchange (SCE) assay. All three agents <span class="hlt">induced</span> SCEs when used singly. /sup 3/H-thymidine was effective in <span class="hlt">inducing</span> SCEs only at 250more » Bq/ml, whereas UV light was effective at both fluences. Although caffeine did not <span class="hlt">induce</span> SCEs when it was added before exposure to bromodeoxyuridine (BrdUrd), which is used to visualize SCEs, it did <span class="hlt">induce</span> SCEs when present during the entire culture period (/sup 3/H-thymidine experiments) or during incubation in BrdUrd (UV experiments). Caffeine markedly enhanced the SCE-<span class="hlt">inducing</span> effect of UV light but did not influence the effect of /sup 3/H-thymidine.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27288117','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27288117"><span>Asiatic acid uncouples respiration in isolated mouse liver mitochondria and <span class="hlt">induces</span> Hep<span class="hlt">G</span><span class="hlt">2</span> cells death.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lu, Yapeng; Liu, Siyuan; Wang, Ying; Wang, Dang; Gao, Jing; Zhu, Li</p> <p>2016-09-05</p> <p>Asiatic acid, one of the triterpenoid components isolated from Centella asiatica, has received increasing attention due to a wide variety of biological activities. To date, little is known about its mechanisms of action. Here we examined the cytotoxic effect of asiatic acid on Hep<span class="hlt">G</span><span class="hlt">2</span> cells and elucidated some of the underlying mechanisms. Asiatic acid <span class="hlt">induced</span> rapid cell death, as well as mitochondrial membrane potential (MMP) dissipation, ATP depletion and cytochrome c release from mitochondria to the cytosol in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. In mitochondria isolated from mouse liver, asiatic acid treatment significantly stimulated the succinate-supported state 4 respiration rate, dissipated the MMP, increased Ca(<span class="hlt">2</span>+) release from Ca(<span class="hlt">2</span>+)-loaded mitochondria, decreased ATP content and promoted cytochrome c release, indicating the uncoupling effect of asiatic acid. Hydrogen peroxide (H<span class="hlt">2</span>O<span class="hlt">2</span>) produced by succinate-supported mitochondrial respiration was also significantly inhibited by asiatic acid. In addition, asiatic acid inhibited Ca(<span class="hlt">2</span>+)-<span class="hlt">induced</span> mitochondrial swelling but did not <span class="hlt">induce</span> mitochondrial swelling in hyposmotic potassium acetate medium which suggested that asiatic acid may not act as a protonophoric uncoupler. Inhibition of uncoupling proteins (UCPs) or blockade of adenine nucleotide transporter (ANT) attenuated the effect of asiatic acid on MMP dissipation, Ca(<span class="hlt">2</span>+) release, mitochondrial respiration and Hep<span class="hlt">G</span><span class="hlt">2</span> cell death. When combined inhibition of UCPs and ANT, asiatic acid-mediated uncoupling effect was noticeably alleviated. These results suggested that both UCPs and ANT partially contribute to the uncoupling properties of asiatic acid. In conclusion, asiatic acid is a novel mitochondrial uncoupler and this property is potentially involved in its toxicity on Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Copyright © 2016 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28974687','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28974687"><span>Monomeric ephrinB<span class="hlt">2</span> binding <span class="hlt">induces</span> allosteric changes in Nipah virus <span class="hlt">G</span> that precede its full activation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wong, Joyce J W; Young, Tracy A; Zhang, Jiayan; Liu, Shiheng; Leser, George P; Komives, Elizabeth A; Lamb, Robert A; Zhou, Z Hong; Salafsky, Joshua; Jardetzky, Theodore S</p> <p>2017-10-03</p> <p>Nipah virus is an emergent paramyxovirus that causes deadly encephalitis and respiratory infections in humans. Two glycoproteins coordinate the infection of host cells, an attachment protein (<span class="hlt">G</span>), which binds to cell surface receptors, and a fusion (F) protein, which carries out the process of virus-cell membrane fusion. The <span class="hlt">G</span> protein binds to ephrin B<span class="hlt">2</span>/3 receptors, <span class="hlt">inducing</span> <span class="hlt">G</span> conformational changes that trigger F protein refolding. Using an optical approach based on second harmonic generation, we show that monomeric and dimeric receptors activate distinct conformational changes in <span class="hlt">G</span>. The monomeric receptor-<span class="hlt">induced</span> changes are not detected by conformation-sensitive monoclonal antibodies or through electron microscopy analysis of <span class="hlt">G</span>:ephrinB<span class="hlt">2</span> complexes. However, hydrogen/deuterium exchange experiments confirm the second harmonic generation observations and reveal allosteric changes in the <span class="hlt">G</span> receptor binding and F-activating stalk domains, providing insights into the pathway of receptor-activated virus entry.Nipah virus causes encephalitis in humans. Here the authors use a multidisciplinary approach to study the binding of the viral attachment protein <span class="hlt">G</span> to its host receptor ephrinB<span class="hlt">2</span> and show that monomeric and dimeric receptors activate distinct conformational changes in <span class="hlt">G</span> and discuss implications for receptor-activated virus entry.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/86423-sodium-arsenite-induces-chromosome-endoreduplication-inhibits-protein-phosphatase-activity-human-fibroblasts','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/86423-sodium-arsenite-induces-chromosome-endoreduplication-inhibits-protein-phosphatase-activity-human-fibroblasts"><span>Sodium arsenite <span class="hlt">induces</span> chromosome endoreduplication and inhibits protein phosphatase activity in human fibroblasts</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Rong-Nan Huang; I-Ching Ho; Ling-Hui Yih</p> <p></p> <p>Arsenic, strongly associated with increased risks of human cancers, is a potent clastogen in a variety of mammalian cell systems. The effect of sodium arsenite (a trivalent arsenic compound) on <span class="hlt">chromatid</span> separation was studied in human skin fibroblasts (HFW). Human fibroblasts were arrested in S phase by the aid of serum starvation and aphidicolin blocking and then these cells were allowed to synchronously progress into <span class="hlt">G</span><span class="hlt">2</span> phase. Treatment of the <span class="hlt">G</span><span class="hlt">2</span>-enriched HFW cells with sodium arsenite (0-200 {mu}M) resulted in arrest of cells in the <span class="hlt">G</span><span class="hlt">2</span> phase, interference with mitotic division, inhibition of spindle assembly, and induction of chromosome endoreduplicationmore » in their second mitosis. Sodium arsenite treatment also inhibited the activities of serine/threonine protein phosphatases and enhanced phosphorylation levels of a small heat shock protein (HSP27). These results suggest that sodium arsenite may mimic okadaic acid to <span class="hlt">induce</span> chromosome endoreduplication through its inhibitory effect on protein phosphatase activity. 61 refs., 6 figs., <span class="hlt">2</span> tabs.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26547127','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26547127"><span>Electronic cigarettes <span class="hlt">induce</span> DNA strand <span class="hlt">breaks</span> and cell death independently of nicotine in cell lines.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yu, Vicky; Rahimy, Mehran; Korrapati, Avinaash; Xuan, Yinan; Zou, Angela E; Krishnan, Aswini R; Tsui, Tzuhan; Aguilera, Joseph A; Advani, Sunil; Crotty Alexander, Laura E; Brumund, Kevin T; Wang-Rodriguez, Jessica; Ongkeko, Weg M</p> <p>2016-01-01</p> <p>Evaluate the cytotoxicity and genotoxicity of short- and long-term e-cigarette vapor exposure on a panel of normal epithelial and head and neck squamous cell carcinoma (HNSCC) cell lines. HaCaT, UMSCC10B, and HN30 were treated with nicotine-containing and nicotine-free vapor extract from two popular e-cigarette brands for periods ranging from 48 h to 8 weeks. Cytotoxicity was assessed using Annexin V flow cytometric analysis, trypan blue exclusion, and clonogenic assays. Genotoxicity in the form of DNA strand <span class="hlt">breaks</span> was quantified using the neutral comet assay and γ-H<span class="hlt">2</span>AX immunostaining. E-cigarette-exposed cells showed significantly reduced cell viability and clonogenic survival, along with increased rates of apoptosis and necrosis, regardless of e-cigarette vapor nicotine content. They also exhibited significantly increased comet tail length and accumulation of γ-H<span class="hlt">2</span>AX foci, demonstrating increased DNA strand <span class="hlt">breaks</span>. E-cigarette vapor, both with and without nicotine, is cytotoxic to epithelial cell lines and is a DNA strand <span class="hlt">break-inducing</span> agent. Further assessment of the potential carcinogenic effects of e-cigarette vapor is urgently needed. Copyright © 2015 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4891196','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4891196"><span>Electronic cigarettes <span class="hlt">induce</span> DNA strand <span class="hlt">breaks</span> and cell death independently of nicotine in cell lines</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Yu, Vicky; Rahimy, Mehran; Korrapati, Avinaash; Xuan, Yinan; Zou, Angela E.; Krishnan, Aswini R.; Tsui, Tzuhan; Aguilera, Joseph A.; Advani, Sunil; Crotty Alexander, Laura E.; Brumund, Kevin T.; Wang-Rodriguez, Jessica</p> <p>2016-01-01</p> <p>Objectives Evaluate the cytotoxicity and genotoxicity of short- and long-term e-cigarette vapor exposure on a panel of normal epithelial and head and neck squamous cell carcinoma (HNSCC) cell lines. Materials and Methods HaCaT, UMSCC10B, and HN30 were treated with nicotine-containing and nicotine-free vapor extract from two popular e-cigarette brands for periods ranging from 48 hours to 8 weeks. Cytotoxicity was assessed using Annexin V flow cytometric analysis, trypan blue exclusion, and clonogenic assays. Genotoxicity in the form of DNA strand <span class="hlt">breaks</span> was quantified using the neutral comet assay and γ-H<span class="hlt">2</span>AX immunostaining. Results E-cigarette-exposed cells showed significantly reduced cell viability and clonogenic survival, along with increased rates of apoptosis and necrosis, regardless of e-cigarette vapor nicotine content. They also exhibited significantly increased comet tail length and accumulation of γ-H<span class="hlt">2</span>AX foci, demonstrating increased DNA strand <span class="hlt">breaks</span>. Conclusion E-cigarette vapor, both with and without nicotine, is cytotoxic to epithelial cell lines and is a DNA strand <span class="hlt">break-inducing</span> agent. Further assessment of the potential carcinogenic effects of e-cigarette vapor is urgently needed. PMID:26547127</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/8814197','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/8814197"><span>Beta-carotene and lutein protect Hep<span class="hlt">G</span><span class="hlt">2</span> human liver cells against oxidant-<span class="hlt">induced</span> damage.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Martin, K R; Failla, M L; Smith, J C</p> <p>1996-09-01</p> <p>Numerous epidemiological studies support a strong inverse relationship between consumption of carotenoid-rich fruits and vegetables and the incidence of some degenerative diseases. One proposed mechanism of protection by carotenoids centers on their putative antioxidant activity, although direct evidence in support of this contention is limited at the cellular level. The antioxidant potential of beta-carotene (BC) and lutein (LUT), carotenoids with or without provitamin A activity, respectively, was evaluated using the human liver cell line Hep<span class="hlt">G</span><span class="hlt">2</span>. Pilot studies showed that a 90-min exposure of confluent cultures to 500 mumol/L tert-butylhydroperoxide (TBHP) at 37 degrees C significantly (P < 0.05) increased lipid peroxidation and cellular leakage of lactate dehydrogenase (LDH), and decreased the uptake of 3H-alpha-aminoisobutyric acid and 3H-<span class="hlt">2</span>-deoxyglucose. Protein synthesis, mitochondrial activity and glucose oxidation were not affected by TBHP treatment, suggesting that the plasma membrane was the primary site of TBHP-<span class="hlt">induced</span> damage. Overnight incubation of cultures with > or = 1 mumol/L dl-alpha-tocopherol protected cells against oxidant-<span class="hlt">induced</span> changes. In parallel studies, overnight incubation of Hep<span class="hlt">G</span><span class="hlt">2</span> in medium containing micelles with either BC or LUT (final concentrations of 1.1 and 10.9 mumol/L, respectively), the cell content of the carotenoids increased from < 0.04 to 0.32 and 3.39 nmol/mg protein, respectively. Carotenoid-loaded cells were partially or completely protected against oxidant-<span class="hlt">induced</span> changes in lipid peroxidation, LDH release and amino acid and deoxyglucose transport. These data demonstrate that BC and LUT or their metabolites protect Hep<span class="hlt">G</span><span class="hlt">2</span> cells against oxidant-<span class="hlt">induced</span> damage and that the protective effect is independent of provitamin A activity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28794467','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28794467"><span>TDP<span class="hlt">2</span> suppresses chromosomal translocations <span class="hlt">induced</span> by DNA topoisomerase II during gene transcription.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gómez-Herreros, Fernando; Zagnoli-Vieira, Guido; Ntai, Ioanna; Martínez-Macías, María Isabel; Anderson, Rhona M; Herrero-Ruíz, Andrés; Caldecott, Keith W</p> <p>2017-08-10</p> <p>DNA double-strand <span class="hlt">breaks</span> (DSBs) <span class="hlt">induced</span> by abortive topoisomerase II (TOP<span class="hlt">2</span>) activity are a potential source of genome instability and chromosome translocation. TOP<span class="hlt">2</span>-<span class="hlt">induced</span> DNA double-strand <span class="hlt">breaks</span> are rejoined in part by tyrosyl-DNA phosphodiesterase <span class="hlt">2</span> (TDP<span class="hlt">2</span>)-dependent non-homologous end-joining (NHEJ), but whether this process suppresses or promotes TOP<span class="hlt">2</span>-<span class="hlt">induced</span> translocations is unclear. Here, we show that TDP<span class="hlt">2</span> rejoins DSBs <span class="hlt">induced</span> during transcription-dependent TOP<span class="hlt">2</span> activity in breast cancer cells and at the translocation 'hotspot', MLL. Moreover, we find that TDP<span class="hlt">2</span> suppresses chromosome rearrangements <span class="hlt">induced</span> by TOP<span class="hlt">2</span> and reduces TOP<span class="hlt">2</span>-<span class="hlt">induced</span> chromosome translocations that arise during gene transcription. Interestingly, however, we implicate TDP<span class="hlt">2</span>-dependent NHEJ in the formation of a rare subclass of translocations associated previously with therapy-related leukemia and characterized by junction sequences with 4-bp of perfect homology. Collectively, these data highlight the threat posed by TOP<span class="hlt">2</span>-<span class="hlt">induced</span> DSBs during transcription and demonstrate the importance of TDP<span class="hlt">2</span>-dependent non-homologous end-joining in protecting both gene transcription and genome stability.DNA double-strand <span class="hlt">breaks</span> (DSBs) <span class="hlt">induced</span> by topoisomerase II (TOP<span class="hlt">2</span>) are rejoined by TDP<span class="hlt">2</span>-dependent non-homologous end-joining (NHEJ) but whether this promotes or suppresses translocations is not clear. Here the authors show that TDP<span class="hlt">2</span> suppresses chromosome translocations from DSBs introduced during gene transcription.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29601810','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29601810"><span><span class="hlt">2</span>',3-dihydroxy-5-methoxybiphenyl suppresses fMLP-<span class="hlt">induced</span> superoxide anion production and cathepsin <span class="hlt">G</span> release by targeting the β-subunit of <span class="hlt">G</span>-protein in human neutrophils.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liao, Hsiang-Ruei; Chen, Ih-Sheng; Liu, Fu-Chao; Lin, Shinn-Zhi; Tseng, Ching-Ping</p> <p>2018-06-15</p> <p>This study investigates the effect and the underlying mechanism of <span class="hlt">2</span>',3-dihydroxy-5-methoxybiphenyl (RIR-<span class="hlt">2</span>), a lignan extracted from the roots of Rhaphiolepis indica (L.) Lindl. ex Ker var. tashiroi Hayata ex Matsum. & Hayata (Rosaceae), on N-formyl-L-methionyl-L-leucyl-L-phenylalanine (fMLP)-<span class="hlt">induced</span> respiratory burst and cathepsin <span class="hlt">G</span> in human neutrophils. Signaling pathways regulated by RIR-<span class="hlt">2</span> which modulated fMLP-<span class="hlt">induced</span> respiratory burst were evaluated by an interaction between β subunit of <span class="hlt">G</span>-protein (Gβ) with downstream signaling <span class="hlt">induced</span> by fMLP and by immunoblotting analysis of the downstream targets of Gβ-protein. RIR-<span class="hlt">2</span> inhibited fMLP-<span class="hlt">induced</span> superoxide anion production (IC 50 :<span class="hlt">2</span>.57 ± 0.22 μM), cathepsin <span class="hlt">G</span> release (IC 50 :18.72 ± 3.76 μM) and migration in a concentration dependent manner. RIR-<span class="hlt">2</span> specifically suppresses fMLP-<span class="hlt">induced</span> Src family kinases phosphorylation by inhibiting the interaction between Gβ-protein with Src kinases without inhibiting Src kinases activities, therefore, RIR-<span class="hlt">2</span> attenuated the downstream targets of Src kinase, such as phosphorylation of Raf/ERK, AKT, P38, PLCγ<span class="hlt">2</span>, PKC and translocation Tec, p47 ph ° x and P40 ph ° x from the cytosol to the inner leaflet of the plasma membrane. Furthermore, RIR-<span class="hlt">2</span> attenuated fMLP-<span class="hlt">induced</span> intracellular calcium mobilization by inhibiting the interaction between Gβ-protein with PLCβ<span class="hlt">2</span>. RIR-<span class="hlt">2</span> was not a competitive or allosteric antagonist of fMLP. On the contrary, phorbol 12-myristate 13-acetate (PMA)-<span class="hlt">induced</span> phosphorylation of Src, AKT, P38, PKC and membrane localization of p47 ph ° x and P40 ph ° x remained unaffected. RIR-<span class="hlt">2</span> specifically modulates fMLP-mediated neutrophil superoxide anion production and cathepsin <span class="hlt">G</span> release by inhibiting the interaction between Gβ-protein with downstream signaling which subsequently interferes with the activation of intracellular calcium, PLCγ<span class="hlt">2</span>, AKT, p38, PKC, ERK, p47 ph ° x and p40 phox . Copyright © 2018 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26961604','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26961604"><span>Tributyltin <span class="hlt">induces</span> <span class="hlt">G</span><span class="hlt">2</span>/M cell cycle arrest via NAD(+)-dependent isocitrate dehydrogenase in human embryonic carcinoma cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Asanagi, Miki; Yamada, Shigeru; Hirata, Naoya; Itagaki, Hiroshi; Kotake, Yaichiro; Sekino, Yuko; Kanda, Yasunari</p> <p>2016-04-01</p> <p>Organotin compounds, such as tributyltin (TBT), are well-known endocrine-disrupting chemicals (EDCs). We have recently reported that TBT <span class="hlt">induces</span> growth arrest in the human embryonic carcinoma cell line NT<span class="hlt">2</span>/D1 at nanomolar levels by inhibiting NAD(+)-dependent isocitrate dehydrogenase (NAD-IDH), which catalyzes the irreversible conversion of isocitrate to α-ketoglutarate. However, the molecular mechanisms by which NAD-IDH mediates TBT toxicity remain unclear. In the present study, we examined whether TBT at nanomolar levels affects cell cycle progression in NT<span class="hlt">2</span>/D1 cells. Propidium iodide staining revealed that TBT reduced the ratio of cells in the <span class="hlt">G</span>1 phase and increased the ratio of cells in the <span class="hlt">G</span><span class="hlt">2</span>/M phase. TBT also reduced cell division cycle 25C (cdc25C) and cyclin B1, which are key regulators of <span class="hlt">G</span><span class="hlt">2</span>/M progression. Furthermore, apigenin, an inhibitor of NAD-IDH, mimicked the effects of TBT. The <span class="hlt">G</span><span class="hlt">2</span>/M arrest <span class="hlt">induced</span> by TBT was abolished by NAD-IDHα knockdown. Treatment with a cell-permeable α-ketoglutarate analogue recovered the effect of TBT, suggesting the involvement of NAD-IDH. Taken together, our data suggest that TBT at nanomolar levels <span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span>/M cell cycle arrest via NAD-IDH in NT<span class="hlt">2</span>/D1 cells. Thus, cell cycle analysis in embryonic cells could be used to assess cytotoxicity associated with nanomolar level exposure of EDCs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5154725','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5154725"><span>Polyethylenimine-functionalized silver nanoparticle-based co-delivery of paclitaxel to <span class="hlt">induce</span> Hep<span class="hlt">G</span><span class="hlt">2</span> cell apoptosis</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Li, Yinghua; Guo, Min; Lin, Zhengfang; Zhao, Mingqi; Xiao, Misi; Wang, Changbing; Xu, Tiantian; Chen, Tianfeng; Zhu, Bing</p> <p>2016-01-01</p> <p>Hepatocarcinoma is the third leading cause of cancer-related deaths around the world. Recently, a novel emerging nanosystem as anticancer therapeutic agents with intrinsic therapeutic properties has been widely used in various medical applications. In this study, surface decoration of functionalized silver nanoparticles (AgNPs) by polyethylenimine (PEI) and paclitaxel (PTX) was synthesized. The purpose of this study was to evaluate the effect of Ag@ PEI@PTX on cytotoxic and anticancer mechanism on Hep<span class="hlt">G</span><span class="hlt">2</span> cells. The transmission electron microscope image and 3-(4,5-dimethylthiazol-<span class="hlt">2</span>-yl)-<span class="hlt">2</span>,5-diphenyltetrazolium bromide assay showed that Ag@PEI@PTX had satisfactory size distribution and high stability and selectivity between cancer and normal cells. Ag@PEI@PTX-<span class="hlt">induced</span> Hep<span class="hlt">G</span><span class="hlt">2</span> cell apoptosis was confirmed by accumulation of the sub-<span class="hlt">G</span>1 cells population, translocation of phosphatidylserine, depletion of mitochondrial membrane potential, DNA fragmentation, caspase-3 activation, and poly(ADP-ribose) polymerase cleavage. Furthermore, Ag@PEI@PTX enhanced cytotoxic effects on Hep<span class="hlt">G</span><span class="hlt">2</span> cells and triggered intracellular reactive oxygen species; the signaling pathways of AKT, p53, and MAPK were activated to advance cell apoptosis. In conclusion, the results reveal that Ag@ PEI@PTX may provide useful information on Ag@PEI@PTX-<span class="hlt">induced</span> Hep<span class="hlt">G</span><span class="hlt">2</span> cell apoptosis and as appropriate candidate for chemotherapy of cancer. PMID:27994465</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1360053-stress-induced-nematicity-eufe2as2-studied-raman-spectroscopy','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1360053-stress-induced-nematicity-eufe2as2-studied-raman-spectroscopy"><span>Stress-<span class="hlt">induced</span> nematicity in EuFe <span class="hlt">2</span> As <span class="hlt">2</span> studied by Raman spectroscopy</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Zhang, W. -L.; Sefat, Athena S.; Ding, H.; ...</p> <p>2016-07-18</p> <p>Here, we use polarized Raman scattering to study the structural phase transition in EuFe <span class="hlt">2</span> As <span class="hlt">2</span> , the parent compound of the 122-ferropnictide superconductors. The in-plane lattice anisotropy is characterized by measurements of the side surface with different strains <span class="hlt">induced</span> by different preparation methods. We also show that while a fine surface polishing leaves the samples free of residual internal strain, in which case the onset of the C 4 symmetry <span class="hlt">breaking</span> is observed at the nominal structural phase transition temperature T S , cutting the side surface <span class="hlt">induces</span> a permanent fourfold rotational symmetry <span class="hlt">breaking</span> spanning tens ofmore » degrees above T S .« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5312349','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5312349"><span>Poor recognition of O6-isopropyl d<span class="hlt">G</span> by MGMT triggers double strand <span class="hlt">break</span>-mediated cell death and micronucleus induction in FANC-deficient cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Hashimoto, Kiyohiro; Sharma, Vyom; Sasanuma, Hiroyuki; Tian, Xu; Takata, Minoru; Takeda, Shunichi; Swenberg, James A.; Nakamura, Jun</p> <p>2016-01-01</p> <p>Isopropyl methanesulfonate (IPMS) is the most potent genotoxic compound among methanesulfonic acid esters. The genotoxic potential of alkyl sulfonate esters is believed to be due to their alkylating ability of the O6 position of guanine. Understanding the primary repair pathway activated in response to IPMS-<span class="hlt">induced</span> DNA damage is important to profile the genotoxic potential of IPMS. In the present study, both chicken DT40 and human TK6 cell-based DNA damage response (DDR) assays revealed that dysfunction of the FANC pathway resulted in higher sensitivity to IPMS compared to EMS or MMS. O6-alkyl d<span class="hlt">G</span> is primarily repaired by methyl guanine methyltransferase (MGMT), while isopropyl d<span class="hlt">G</span> is less likely to be a substrate for MGMT. Comparison of the cytotoxic potential of IPMS and its isomer n-propyl methanesulfonate (nPMS) revealed that the isopropyl moiety avoids recognition by MGMT and leads to higher cytotoxicity. Next, the micronucleus (MN) assay showed that FANC deficiency increases the sensitivity of DT40 cells to MN induction by IPMS. Pretreatment with O6-benzyl guanine (OBG), an inhibitor of MGMT, increased the MN frequency in DT40 cells treated with nPMS, but not IPMS. Lastly, IPMS <span class="hlt">induced</span> more double strand <span class="hlt">breaks</span> in FANC-deficient cells compared to wild-type cells in a time-dependent manner. All together, these results suggest that IPMS-derived O6-isopropyl d<span class="hlt">G</span> escapes recognition by MGMT, and the unrepaired DNA damage leads to double strand <span class="hlt">breaks</span>, resulting in MN induction. FANC, therefore, plays a pivotal role in preventing MN induction and cell death caused by IPMS. PMID:27486975</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4416797','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4416797"><span>Mechanism of Arctigenin-<span class="hlt">Induced</span> Specific Cytotoxicity against Human Hepatocellular Carcinoma Cell Lines: Hep <span class="hlt">G</span><span class="hlt">2</span> and SMMC7721</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Lu, Zheng; Cao, Shengbo; Zhou, Hongbo; Hua, Ling; Zhang, Shishuo; Cao, Jiyue</p> <p>2015-01-01</p> <p>Arctigenin (ARG) has been previously reported to exert high biological activities including anti-inflammatory, antiviral and anticancer. In this study, the anti-tumor mechanism of ARG towards human hepatocellular carcinoma (HCC) was firstly investigated. We demonstrated that ARG could <span class="hlt">induce</span> apoptosis in Hep <span class="hlt">G</span><span class="hlt">2</span> and SMMC7721 cells but not in normal hepatic cells, and its apoptotic effect on Hep <span class="hlt">G</span><span class="hlt">2</span> was stronger than that on SMMC7721. Furthermore, the following study showed that ARG treatment led to a loss in the mitochondrial out membrane potential, up-regulation of Bax, down-regulation of Bcl-<span class="hlt">2</span>, a release of cytochrome c, caspase-9 and caspase-3 activation and a cleavage of poly (ADP-ribose) polymerase in both Hep <span class="hlt">G</span><span class="hlt">2</span> and SMMC7721 cells, suggesting ARG-<span class="hlt">induced</span> apoptosis was associated with the mitochondria mediated pathway. Moreover, the activation of caspase-8 and the increased expression levels of Fas/FasL and TNF-α revealed that the Fas/FasL-related pathway was also involved in this process. Additionally, ARG <span class="hlt">induced</span> apoptosis was accompanied by a deactivation of PI3K/p-Akt pathway, an accumulation of p53 protein and an inhibition of NF-κB nuclear translocation especially in Hep <span class="hlt">G</span><span class="hlt">2</span> cells, which might be the reason that Hep <span class="hlt">G</span><span class="hlt">2</span> was more sensitive than SMMC7721 cells to ARG treatment. PMID:25933104</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25622747','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25622747"><span>[Ursodeoxycholic acid <span class="hlt">induced</span> apoptosis of human hepatoma cells Hep<span class="hlt">G</span><span class="hlt">2</span> and SMMC-7721 bymitochondrial-mediated pathway].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wu, Duan; Zhou, Jianyin; Yin, Zhenyu; Liu, Pingguo; Zhao, Yilin; Liu, Jianming; Wang, Xiaomin</p> <p>2014-12-02</p> <p>To explore the effects and underlying mechanisms of ursodeoxycholic acid on human hepatoma cells. Hep<span class="hlt">G</span><span class="hlt">2</span> and SMMC-7721 HCC cell lines were respectively treated with ursodeoxycholic acid. And cell proliferation, apoptosis and the expression of Bax/Bcl-<span class="hlt">2</span> gene were detected by methyl thiazolyl tetrazolium (MTT), inverted microscopy, fluorescent microscopy, flow cytometry and Western blot. Ursodeoxycholic acid significantly inhibited the proliferation of human hepatoma cells in a concentration- and time-dependent manner. The half maximal inhibitory concentrations (IC50) of Hep<span class="hlt">G</span><span class="hlt">2</span> and SMMC-7721 were 397.3 and 387.7 µg/ml respectively after a 48-hour treatment of 400 µg /ml ursodeoxycholic acid. And it also <span class="hlt">induced</span> the apoptosis of Hep<span class="hlt">G</span><span class="hlt">2</span> and SMMC-7721 cells, up-regulated Bax gene and down-regulated Bcl-<span class="hlt">2</span> gene. Ursodeoxycholic acid inhibits the proliferation of hepatoma cells and <span class="hlt">induce</span> apoptosis by mitochondrial-mediated pathway.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.T32C..06B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.T32C..06B"><span>Anomalous Lower Crustal and Surface Features as a Result of Plume-<span class="hlt">induced</span> Continental <span class="hlt">Break</span>-up: Inferences from Numerical Models</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Beniest, A.; Koptev, A.; Leroy, S. D.</p> <p>2016-12-01</p> <p>Anomalous features along the South American and African rifted margins at depth and at the surface have been recognised with gravity and magnetic modelling. They include high velocity/high density bodies at lower crustal level and topography variations that are usually interpreted as aborted rifts. We present fully-coupled lithosphere-scale numerical models that permit us to explain both features in a relatively simple framework of an interaction between rheologically stratified continental lithosphere and an active mantle plume. We used <span class="hlt">2</span>D and 3D numerical models to investigate the impact of thermo-rheological structure of the continental lithosphere and initial plume position on continental rifting and breakup processes. Based on the results of our <span class="hlt">2</span>D experiments, three main types of continental <span class="hlt">break</span>-up are revealed: A) mantle plume-<span class="hlt">induced</span> <span class="hlt">break</span>-up, directly located above the centre of the mantle anomaly, B) mantle plume-<span class="hlt">induced</span> <span class="hlt">break</span>-up, 50 to 250 km displaced from the initial plume location and C) self-<span class="hlt">induced</span> <span class="hlt">break</span>-up due to convection and/or slab-subduction/delamination, considerably shifted (300 to 800 km) from the initial plume position. With our 3D, laterally homogenous initial setup, we show that a complex system, with the axis of continental <span class="hlt">break</span>-up 100's of km's shifted from the original plume location, can arise spontaneously from simple and perfectly symmetric preliminary settings. Our modelling demonstrates that fragments of a laterally migrating plume head become glued to the base of the lithosphere and remain at both sides of the newly-formed oceanic basin after continental <span class="hlt">break</span>-up. Underplated plume material soldered into lower parts of lithosphere can be interpreted as the high-velocity/high density magmatic bodies at lower crustal levels. In the very early stages of rifting, first impingement of the vertically upwelled mantle plume to the lithospheric base leads to surface topographic variations. Given the shifted position of the final</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23681662','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23681662"><span>A CO-FISH assay to assess sister <span class="hlt">chromatid</span> segregation patterns in mitosis of mouse embryonic stem cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sauer, Stephan; Burkett, Sandra S; Lewandoski, Mark; Klar, Amar J S</p> <p>2013-05-01</p> <p>Sister <span class="hlt">chromatids</span> contain identical DNA sequence but are chiral with respect to both their helical handedness and their replication history. Emerging evidence from various model organisms suggests that certain stem cells segregate sister <span class="hlt">chromatids</span> nonrandomly to either maintain genome integrity or to bias cellular differentiation in asymmetric cell divisions. Conventional methods for tracing of old vs. newly synthesized DNA strands generally lack resolution for individual chromosomes and employ halogenated thymidine analogs with profound cytotoxic effects on rapidly dividing cells. Here, we present a modified chromosome orientation fluorescence in situ hybridization (CO-FISH) assay, where identification of individual chromosomes and their replication history is achieved in subsequent hybridization steps with chromosome-specific DNA probes and PNA telomere probes. Importantly, we tackle the issue of BrdU cytotoxicity and show that our method is compatible with normal mouse ES cell biology, unlike a recently published related protocol. Results from our CO-FISH assay show that mitotic segregation of mouse chromosome 7 is random in ES cells, which contrasts previously published results from our laboratory and settles a controversy. Our straightforward protocol represents a useful resource for future studies on <span class="hlt">chromatid</span> segregation patterns of in vitro-cultured cells from distinct model organisms.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18179804','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18179804"><span>Comparison of the induction and disappearance of DNA double strand <span class="hlt">breaks</span> and gamma-H<span class="hlt">2</span>AX foci after irradiation of chromosomes in <span class="hlt">G</span>1-phase or in condensed metaphase cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kato, Takamitsu A; Okayasu, Ryuichi; Bedford, Joel S</p> <p>2008-03-01</p> <p>The induction and disappearance of DNA double strand <span class="hlt">breaks</span> (DSBs) after irradiation of <span class="hlt">G</span>1 and mitotic cells were compared with the gamma-H<span class="hlt">2</span>AX foci assay and a gel electrophoresis assay. This is to determine whether cell cycle related changes in chromatin structure might influence the gamma-H<span class="hlt">2</span>AX assay which depends on extensive phosphorylation and dephosphorylation of the H<span class="hlt">2</span>AX histone variant surrounding DSBs. The disappearance of gamma-H<span class="hlt">2</span>AX foci after irradiation was much slower for mitotic than for <span class="hlt">G</span>1 cells. On the other hand, no difference was seen for the gel electrophoresis assay. Our data may suggest the limited accessibility of dephosphorylation enzyme in irradiated metaphase cells or trapped gamma-H<span class="hlt">2</span>AX in condensed chromatin.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/10390512','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/10390512"><span>Induction and prevention of micronuclei and chromosomal aberrations in cultured human lymphocytes exposed to the light of halogen tungsten lamps.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>D'Agostini, F; Caimo, A; De Filippi, S; De Flora, S</p> <p>1999-07-01</p> <p>Previous studies have shown that the light emitted by halogen tungsten lamps contains UV radiation in the UV-A, UV-B and UV-C regions, <span class="hlt">induces</span> mutations and irreparable DNA damage in bacteria, enhances the frequency of micronuclei in cultured human lymphocytes and is potently carcinogenic to the skin of hairless mice. The present study showed that the light emitted by an uncovered, traditional halogen lamp <span class="hlt">induces</span> a significant, dose-related and time-related increase not only in micronuclei but also in chromosome-type aberrations, such as <span class="hlt">breaks</span>, and even more in <span class="hlt">chromatid</span>-type aberrations, such as isochromatid <span class="hlt">breaks</span>, exchanges and isochromatid/<span class="hlt">chromatid</span> interchanges, all including gaps or not, in cultured human lymphocytes. All these genotoxic effects were completely prevented by shielding the same lamp with a silica glass cover, blocking UV radiation. A new model of halogen lamp, having the quartz bulb treated in order to reduce the output of UV radiation, was considerably less genotoxic than the uncovered halogen lamp, yet induction of chromosomal alterations was observed at high illuminance levels.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li class="active"><span>10</span></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_10 --> <div id="page_11" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li class="active"><span>11</span></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="201"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006PhLB..633..512M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006PhLB..633..512M"><span>Constraints on the I = 1 hadronic τ decay and e+e- →hadrons data sets and implications for (<span class="hlt">g</span> - <span class="hlt">2</span>) μ</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Maltman, Kim</p> <p>2006-02-01</p> <p>Sum rule tests are performed on the spectral data for (i) flavor ud vector-current-<span class="hlt">induced</span> hadronic τ decays and (ii) e+e- hadroproduction, in the region below s ∼ 3- 4 GeV<span class="hlt">2</span>, where discrepancies exist between the isospin-<span class="hlt">breaking</span>-corrected charged and neutral current I = 1 spectral functions. The τ data is found to be compatible with expectations based on high-scale αs (MZ) determinations, while the electroproduction data displays two problems. The results favor determinations of the leading order hadronic contribution to (<span class="hlt">g</span> - <span class="hlt">2</span>) μ which incorporate hadronic τ decay data over those employing electroproduction data only, and hence a reduced discrepancy between experiment and the Standard Model prediction for (<span class="hlt">g</span> - <span class="hlt">2</span>) μ.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012Nanos...4.7168P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012Nanos...4.7168P"><span>Copper(ii) oxide nanoparticles penetrate into Hep<span class="hlt">G</span><span class="hlt">2</span> cells, exert cytotoxicity via oxidative stress and <span class="hlt">induce</span> pro-inflammatory response</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Piret, Jean-Pascal; Jacques, Diane; Audinot, Jean-Nicolas; Mejia, Jorge; Boilan, Emmanuelle; Noël, Florence; Fransolet, Maude; Demazy, Catherine; Lucas, Stéphane; Saout, Christelle; Toussaint, Olivier</p> <p>2012-10-01</p> <p>The potential toxic effects of two types of copper(ii) oxide (CuO) nanoparticles (NPs) with different specific surface areas, different shapes (rod or spheric), different sizes as raw materials and similar hydrodynamic diameter in suspension were studied on human hepatocarcinoma Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Both CuO NPs were shown to be able to enter into Hep<span class="hlt">G</span><span class="hlt">2</span> cells and <span class="hlt">induce</span> cellular toxicity by generating reactive oxygen species. CuO NPs increased the abundance of several transcripts coding for pro-inflammatory interleukins and chemokines. Transcriptomic data, siRNA knockdown and DNA binding activities suggested that Nrf<span class="hlt">2</span>, NF-κB and AP-1 were implicated in the response of Hep<span class="hlt">G</span><span class="hlt">2</span> cells to CuO NPs. CuO NP incubation also <span class="hlt">induced</span> activation of MAPK pathways, ERKs and JNK/SAPK, playing a major role in the activation of AP-1. In addition, cytotoxicity, inflammatory and antioxidative responses and activation of intracellular transduction pathways <span class="hlt">induced</span> by rod-shaped CuO NPs were more important than spherical CuO NPs. Measurement of Cu<span class="hlt">2</span>+ released in cell culture medium suggested that Cu<span class="hlt">2</span>+ cations released from CuO NPs were involved only to a small extent in the toxicity <span class="hlt">induced</span> by these NPs on Hep<span class="hlt">G</span><span class="hlt">2</span> cells.The potential toxic effects of two types of copper(ii) oxide (CuO) nanoparticles (NPs) with different specific surface areas, different shapes (rod or spheric), different sizes as raw materials and similar hydrodynamic diameter in suspension were studied on human hepatocarcinoma Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Both CuO NPs were shown to be able to enter into Hep<span class="hlt">G</span><span class="hlt">2</span> cells and <span class="hlt">induce</span> cellular toxicity by generating reactive oxygen species. CuO NPs increased the abundance of several transcripts coding for pro-inflammatory interleukins and chemokines. Transcriptomic data, siRNA knockdown and DNA binding activities suggested that Nrf<span class="hlt">2</span>, NF-κB and AP-1 were implicated in the response of Hep<span class="hlt">G</span><span class="hlt">2</span> cells to CuO NPs. CuO NP incubation also <span class="hlt">induced</span> activation of MAPK pathways, ERKs and JNK/SAPK, playing a major</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26446980','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26446980"><span>ML-7 amplifies the quinocetone-<span class="hlt">induced</span> cell death through akt and MAPK-mediated apoptosis on Hep<span class="hlt">G</span><span class="hlt">2</span> cell line.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhou, Yan; Zhang, Shen; Deng, Sijun; Dai, Chongshan; Tang, Shusheng; Yang, Xiayun; Li, Daowen; Zhao, Kena; Xiao, Xilong</p> <p>2016-01-01</p> <p>The study aims at evaluating the combination of the quinocetone and the ML-7 in preclinical hepatocellular carcinoma models. To this end, the effect of quinocetone and ML-7 on apoptosis induction and signaling pathways was analyzed on Hep<span class="hlt">G</span><span class="hlt">2</span> cell lines. Here, we report that ML-7, in a nontoxic concentration, sensitized the Hep<span class="hlt">G</span><span class="hlt">2</span> cells to quinocetone-<span class="hlt">induced</span> cytotoxicity. Also, ML-7 profoundly enhances quinocetone-<span class="hlt">induced</span> apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cell line. Mechanistic investigations revealed that ML-7 and quinocetone act in concert to trigger the cleavage of caspase-8 as well as Bax/Bcl-<span class="hlt">2</span> ratio up-regulation and subsequent cleavage of Bid, capsases-9 and -3. Importantly, ML-7 weakened the quinocetone-<span class="hlt">induced</span> Akt pathway activation, but strengthened the phosphorylation of p-38, ERK and JNK. Further treatment of Akt activator and p-38 inhibitor almost completely abolished the ML-7/quinocetone-<span class="hlt">induced</span> apoptosis. In contrast, the ERK and JNK inhibitor aggravated the ML-7/quinocetone-<span class="hlt">induced</span> apoptosis, indicating that the synergism critically depended on p-38 phosphorylation and Hep<span class="hlt">G</span><span class="hlt">2</span> cells provoke Akt, ERK and JNK signaling pathways to against apoptosis. In conclusion, the rational combination of quinocetone and ML-7 presents a promising approach to trigger apoptosis in hepatocellular carcinoma, which warrants further investigation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19112901','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19112901"><span>[Apoptosis and activity changes of telomerase <span class="hlt">induced</span> by essential oil from pine needles in Hep<span class="hlt">G</span><span class="hlt">2</span> cell line].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wei, Feng-xiang; Li, Mei-yu; Song, Yu-hong; Li, Hong-zhi</p> <p>2008-08-01</p> <p>To study the effects of essential oil extracted from pine needles on Hep<span class="hlt">G</span><span class="hlt">2</span> cell line. Hep<span class="hlt">G</span><span class="hlt">2</span> cells were treated with essential oil extracted from pine needles. Cell growth rate was determined with MTF assay, cell morphologic changes were examined under transmission electromicroscope and HE straining. Flow cytometry was used to exmine apoptotic cells. Bcl-<span class="hlt">2</span> gene expression was determined by flow cytometry and telomerase activity by TRAP assay. Essential oils from pine needles could not only repress the growth of Hep<span class="hlt">G</span><span class="hlt">2</span> cells significantly, but also <span class="hlt">induce</span> apoptosis to them. Both dose-effect and time-effect relationship could be confirmed. Typical morphology changes of apoptosis such as nuclear enrichment and karyorrhexis were observed through transmission electromicroscope and HE straining. Telomerase activity was down regulated in the essential oil extracted from pine needles <span class="hlt">induced</span> apoptotic cells. The expression of bcl-<span class="hlt">2</span> gene was suppressed after the essential oil from pine needles treatement. The essential oil extracted from pine needles can inhibit cell growth of Hep<span class="hlt">G</span><span class="hlt">2</span> cell line and <span class="hlt">induce</span> apoptosis, which may associate with inhibition of telomerase activity and bcl-<span class="hlt">2</span> may be involved in the regulation of telomerase activity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017Sci...356..295H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017Sci...356..295H"><span>A parity-<span class="hlt">breaking</span> electronic nematic phase transition in the spin-orbit coupled metal Cd<span class="hlt">2</span>Re<span class="hlt">2</span>O7</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Harter, J. W.; Zhao, Z. Y.; Yan, J.-Q.; Mandrus, D. G.; Hsieh, D.</p> <p>2017-04-01</p> <p>Strong electron interactions can drive metallic systems toward a variety of well-known symmetry-broken phases, but the instabilities of correlated metals with strong spin-orbit coupling have only recently begun to be explored. We uncovered a multipolar nematic phase of matter in the metallic pyrochlore Cd<span class="hlt">2</span>Re<span class="hlt">2</span>O7 using spatially resolved second-harmonic optical anisotropy measurements. Like previously discovered electronic nematic phases, this multipolar phase spontaneously <span class="hlt">breaks</span> rotational symmetry while preserving translational invariance. However, it has the distinguishing property of being odd under spatial inversion, which is allowed only in the presence of spin-orbit coupling. By examining the critical behavior of the multipolar nematic order parameter, we show that it drives the thermal phase transition near 200 kelvin in Cd<span class="hlt">2</span>Re<span class="hlt">2</span>O7 and <span class="hlt">induces</span> a parity-<span class="hlt">breaking</span> lattice distortion as a secondary order.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/10827941','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/10827941"><span>Splitting the chromosome: cutting the ties that bind sister <span class="hlt">chromatids</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nasmyth, K; Peters, J M; Uhlmann, F</p> <p>2000-05-26</p> <p>In eukaryotic cells, sister DNA molecules remain physically connected from their production at S phase until their separation during anaphase. This cohesion is essential for the separation of sister <span class="hlt">chromatids</span> to opposite poles of the cell at mitosis. It also permits chromosome segregation to take place long after duplication has been completed. Recent work has identified a multisubunit complex called cohesin that is essential for connecting sisters. Proteolytic cleavage of one of cohesin's subunits may trigger sister separation at the onset of anaphase.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22389213','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22389213"><span>6-Gingerol <span class="hlt">induces</span> apoptosis through lysosomal-mitochondrial axis in human hepatoma <span class="hlt">G</span><span class="hlt">2</span> cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yang, Guang; Wang, Shaopeng; Zhong, Laifu; Dong, Xu; Zhang, Wenli; Jiang, Liping; Geng, Chengyan; Sun, Xiance; Liu, Xiaofang; Chen, Min; Ma, Yufang</p> <p>2012-11-01</p> <p>6-Gingerol, a major phenolic compound derived from ginger, has been known to possess anticarcinogenic activities. However, the mechanisms are not well understood. In our previous study, it was demonstrated that lysosome and mitochondria may be the primary targets for 6-gingerol in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Therefore, the aim was to evaluate lysosome-mitochondria cross-signaling in 6-gingerol-<span class="hlt">induced</span> apoptosis. Apoptosis was detected by Hoechst 33342 and TUNEL assay after 24 h treatment, and the destabilization of lysosome and mitochondria were early upstream initiating events. This study showed that cathepsin D played a crucial role in the process of apoptosis. The release of cathepsin D to the cytosol appeared to be an early event that preceded the release of cytochrome c from mitochondria. Moreover, inhibition of cathepsin D activity resulted in suppressed release of cytochrome c. To further determine the involvement of oxidative stress in 6-gingerol-<span class="hlt">induced</span> apoptosis, the intracellular generation of reactive oxygen species (ROS) and reduced glutathione (GSH) were examined. Taken together, these results suggest that cathepsin D may be a positive mediator of 6-gingerol <span class="hlt">induced</span> apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells, acting upstream of cytochrome c release, and the apoptosis may be associated with oxidative stress. Copyright © 2012 John Wiley & Sons, Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22215326-cyclosporine-palmitic-acid-treatment-synergistically-induce-cytotoxicity-hepg2-cells','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22215326-cyclosporine-palmitic-acid-treatment-synergistically-induce-cytotoxicity-hepg2-cells"><span>Cyclosporine A and palmitic acid treatment synergistically <span class="hlt">induce</span> cytotoxicity in Hep<span class="hlt">G</span><span class="hlt">2</span> cells</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Luo, Yi, E-mail: yi.luo@pfizer.com; Rana, Payal; Will, Yvonne</p> <p></p> <p>Immunosuppressant cyclosporine A (CsA) treatment can cause severe side effects. Patients taking immunosuppressant after organ transplantation often display hyperlipidemia and obesity. Elevated levels of free fatty acids have been linked to the etiology of metabolic syndromes, nonalcoholic fatty liver and steatohepatitis. The contribution of free fatty acids to CsA-<span class="hlt">induced</span> toxicity is not known. In this study we explored the effect of palmitic acid on CsA-<span class="hlt">induced</span> toxicity in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. CsA by itself at therapeutic exposure levels did not <span class="hlt">induce</span> detectible cytotoxicity in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Co-treatment of palmitic acid and CsA resulted in a dose dependent increase in cytotoxicity, suggesting thatmore » fatty acid could sensitize cells to CsA-<span class="hlt">induced</span> cytotoxicity at the therapeutic doses of CsA. A synergized induction of caspase-3/7 activity was also observed, indicating that apoptosis may contribute to the cytotoxicity. We demonstrated that CsA reduced cellular oxygen consumption which was further exacerbated by palmitic acid, implicating that impaired mitochondrial respiration might be an underlying mechanism for the enhanced toxicity. Inhibition of c-Jun N-terminal kinase (JNK) attenuated palmitic acid and CsA <span class="hlt">induced</span> toxicity, suggesting that JNK activation plays an important role in mediating the enhanced palmitic acid/CsA-<span class="hlt">induced</span> toxicity. Our data suggest that elevated FFA levels, especially saturated FFA such as palmitic acid, may be predisposing factors for CsA toxicity, and patients with underlying diseases that would elevate free fatty acids may be susceptible to CsA-<span class="hlt">induced</span> toxicity. Furthermore, hyperlipidemia/obesity resulting from immunosuppressive therapy may aggravate CsA-<span class="hlt">induced</span> toxicity and worsen the outcome in transplant patients. -- Highlights: ► Palmitic acid and cyclosporine (CsA) synergistically increased cytotoxicity. ► The impairment of mitochondrial functions may contribute to the enhanced toxicity. ► Inhibition of JNK activity</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4326275','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4326275"><span>RAD18 Activates the <span class="hlt">G</span><span class="hlt">2</span>/M Checkpoint through DNA Damage Signaling to Maintain Genome Integrity after Ionizing Radiation Exposure</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Sasatani, Megumi; Xu, Yanbin; Kawai, Hidehiko; Cao, Lili; Tateishi, Satoshi; Shimura, Tsutomu; Li, Jianxiang; Iizuka, Daisuke; Noda, Asao; Hamasaki, Kanya; Kusunoki, Yoichiro; Kamiya, Kenji</p> <p>2015-01-01</p> <p>The ubiquitin ligase RAD18 is involved in post replication repair pathways via its recruitment to stalled replication forks, and its role in the ubiquitylation of proliferating cell nuclear antigen (PCNA). Recently, it has been reported that RAD18 is also recruited to DNA double strand <span class="hlt">break</span> (DSB) sites, where it plays novel functions in the DNA damage response <span class="hlt">induced</span> by ionizing radiation (IR). This new role is independent of PCNA ubiquitylation, but little is known about how RAD18 functions after IR exposure. Here, we describe a role for RAD18 in the IR-<span class="hlt">induced</span> DNA damage signaling pathway at <span class="hlt">G</span><span class="hlt">2</span>/M phase in the cell cycle. Depleting cells of RAD18 reduced the recruitment of the DNA damage signaling factors ATM, γH<span class="hlt">2</span>AX, and 53BP1 to foci in cells at the <span class="hlt">G</span><span class="hlt">2</span>/M phase after IR exposure, and attenuated activation of the <span class="hlt">G</span><span class="hlt">2</span>/M checkpoint. Furthermore, depletion of RAD18 increased micronuclei formation and cell death following IR exposure, both in vitro and in vivo. Our data suggest that RAD18 can function as a mediator for DNA damage response signals to activate the <span class="hlt">G</span><span class="hlt">2</span>/M checkpoint in order to maintain genome integrity and cell survival after IR exposure. PMID:25675240</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004PhRvB..69b4509T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004PhRvB..69b4509T"><span>Magnetic and superconducting phase diagrams of single-crystal Er0.8R0.<span class="hlt">2</span>Ni<span class="hlt">2</span>B<span class="hlt">2</span>C (R=Tb,Lu) and ErNi1.9Co0.1B<span class="hlt">2</span>C: Identification of pair-<span class="hlt">breaking</span> mechanisms</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Takeya, H.; El Massalami, M.</p> <p>2004-01-01</p> <p>We investigated the magnetism, superconductivity and their interplay in single crystals Er0.8R0.<span class="hlt">2</span>Ni<span class="hlt">2</span>B<span class="hlt">2</span>C (R=Tb,Lu) and ErNi1.9Co0.1B<span class="hlt">2</span>C. In contrast to Co substitution, R substitutions <span class="hlt">induce</span> considerable modifications in the magnetism of Er sublattice: e.<span class="hlt">g</span>., Tb (Lu) substitution enhances (reduces) TN and critical fields. Both R substitutions introduce size effects and pinning centers; the former modifies the magnon specific heat while the latter hinders the formation of a weak ferromagnetism. The superconductivity, on the other hand, is strongly (weakly) influenced by Tb and Co (Lu) substitution. Taking LuNi<span class="hlt">2</span>B<span class="hlt">2</span>C as a nonmagnetic superconducting limit, we analyzed their superconductivities, as well as that of ErNi<span class="hlt">2</span>B<span class="hlt">2</span>C, in terms of multiple pair <span class="hlt">breaking</span> theory on dirty superconductors. Based on this analysis, many of their superconducting features can be explained: The breakdown of de Gennes scaling is due to the presence of multiple pair breakers, the anisotropy of Hc<span class="hlt">2</span>(T) is related to the magnetic anisotropy, the absence of a structure in Hc<span class="hlt">2</span>(T) at TN of Lu substitution (TN<Tc) is attributed to an alloying-<span class="hlt">induced</span> destruction of phase space truncation, and the quasi parabolic temperature dependence of Hc<span class="hlt">2</span>(T) of Tb and Co substitutions is in part due to a saturation of antiferromagnetic correlations. For Lu substitution, the strength of magnon mediated pair <span class="hlt">breaking</span> process(es) is substantially reduced.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25745980','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25745980"><span>Sodium valproate <span class="hlt">induces</span> mitochondrial respiration dysfunction in Hep<span class="hlt">G</span><span class="hlt">2</span> in vitro cell model.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Komulainen, Tuomas; Lodge, Tiffany; Hinttala, Reetta; Bolszak, Maija; Pietilä, Mika; Koivunen, Peppi; Hakkola, Jukka; Poulton, Joanna; Morten, Karl J; Uusimaa, Johanna</p> <p>2015-05-04</p> <p>Sodium valproate (VPA) is a potentially hepatotoxic antiepileptic drug. Risk of VPA-<span class="hlt">induced</span> hepatotoxicity is increased in patients with mitochondrial diseases and especially in patients with POLG1 gene mutations. We used a Hep<span class="hlt">G</span><span class="hlt">2</span> cell in vitro model to investigate the effect of VPA on mitochondrial activity. Cells were incubated in glucose medium and mitochondrial respiration-<span class="hlt">inducing</span> medium supplemented with galactose and pyruvate. VPA treatments were carried out at concentrations of 0-<span class="hlt">2</span>.0mM for 24-72 h. In both media, VPA caused decrease in oxygen consumption rates and mitochondrial membrane potential. VPA exposure led to depleted ATP levels in Hep<span class="hlt">G</span><span class="hlt">2</span> cells incubated in galactose medium suggesting dysfunction in mitochondrial ATP production. In addition, VPA exposure for 72 h increased levels of mitochondrial reactive oxygen species (ROS), but adversely decreased protein levels of mitochondrial superoxide dismutase SOD<span class="hlt">2</span>, suggesting oxidative stress caused by impaired elimination of mitochondrial ROS and a novel pathomechanism related to VPA toxicity. Increased cell death and decrease in cell number was detected under both metabolic conditions. However, immunoblotting did not show any changes in the protein levels of the catalytic subunit A of mitochondrial DNA polymerase γ, the mitochondrial respiratory chain complexes I, II and IV, ATP synthase, E3 subunit dihydrolipoyl dehydrogenase of pyruvate dehydrogenase, <span class="hlt">2</span>-oxoglutarate dehydrogenase and glutathione peroxidase. Our results show that VPA inhibits mitochondrial respiration and leads to mitochondrial dysfunction, oxidative stress and increased cell death, thus suggesting an essential role of mitochondria in VPA-<span class="hlt">induced</span> hepatotoxicity. Copyright © 2015 Elsevier Ireland Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27934874','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27934874"><span>Raman signatures of inversion symmetry <span class="hlt">breaking</span> and structural phase transition in type-II Weyl semimetal MoTe<span class="hlt">2</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Kenan; Bao, Changhua; Gu, Qiangqiang; Ren, Xiao; Zhang, Haoxiong; Deng, Ke; Wu, Yang; Li, Yuan; Feng, Ji; Zhou, Shuyun</p> <p>2016-12-09</p> <p>Transition metal dichalcogenide MoTe <span class="hlt">2</span> is an important candidate for realizing the newly predicted type-II Weyl fermions, for which the <span class="hlt">breaking</span> of the inversion symmetry is a prerequisite. Here we present direct spectroscopic evidence for the inversion symmetry <span class="hlt">breaking</span> in the low-temperature phase of MoTe <span class="hlt">2</span> by systematic Raman experiments and first-principles calculations. We identify five lattice vibrational modes that are Raman-active only in the low-temperature noncentrosymmetric structure. A hysteresis is also observed in the peak intensity of inversion symmetry-activated Raman modes, confirming a temperature-<span class="hlt">induced</span> structural phase transition with a concomitant change in the inversion symmetry. Our results provide definitive evidence for the low-temperature noncentrosymmetric T d phase from vibrational spectroscopy, and suggest MoTe <span class="hlt">2</span> as an ideal candidate for investigating the temperature-<span class="hlt">induced</span> topological phase transition.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5155143','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5155143"><span>Raman signatures of inversion symmetry <span class="hlt">breaking</span> and structural phase transition in type-II Weyl semimetal MoTe<span class="hlt">2</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Zhang, Kenan; Bao, Changhua; Gu, Qiangqiang; Ren, Xiao; Zhang, Haoxiong; Deng, Ke; Wu, Yang; Li, Yuan; Feng, Ji; Zhou, Shuyun</p> <p>2016-01-01</p> <p>Transition metal dichalcogenide MoTe<span class="hlt">2</span> is an important candidate for realizing the newly predicted type-II Weyl fermions, for which the <span class="hlt">breaking</span> of the inversion symmetry is a prerequisite. Here we present direct spectroscopic evidence for the inversion symmetry <span class="hlt">breaking</span> in the low-temperature phase of MoTe<span class="hlt">2</span> by systematic Raman experiments and first-principles calculations. We identify five lattice vibrational modes that are Raman-active only in the low-temperature noncentrosymmetric structure. A hysteresis is also observed in the peak intensity of inversion symmetry-activated Raman modes, confirming a temperature-<span class="hlt">induced</span> structural phase transition with a concomitant change in the inversion symmetry. Our results provide definitive evidence for the low-temperature noncentrosymmetric Td phase from vibrational spectroscopy, and suggest MoTe<span class="hlt">2</span> as an ideal candidate for investigating the temperature-<span class="hlt">induced</span> topological phase transition. PMID:27934874</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016NatCo...713552Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016NatCo...713552Z"><span>Raman signatures of inversion symmetry <span class="hlt">breaking</span> and structural phase transition in type-II Weyl semimetal MoTe<span class="hlt">2</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhang, Kenan; Bao, Changhua; Gu, Qiangqiang; Ren, Xiao; Zhang, Haoxiong; Deng, Ke; Wu, Yang; Li, Yuan; Feng, Ji; Zhou, Shuyun</p> <p>2016-12-01</p> <p>Transition metal dichalcogenide MoTe<span class="hlt">2</span> is an important candidate for realizing the newly predicted type-II Weyl fermions, for which the <span class="hlt">breaking</span> of the inversion symmetry is a prerequisite. Here we present direct spectroscopic evidence for the inversion symmetry <span class="hlt">breaking</span> in the low-temperature phase of MoTe<span class="hlt">2</span> by systematic Raman experiments and first-principles calculations. We identify five lattice vibrational modes that are Raman-active only in the low-temperature noncentrosymmetric structure. A hysteresis is also observed in the peak intensity of inversion symmetry-activated Raman modes, confirming a temperature-<span class="hlt">induced</span> structural phase transition with a concomitant change in the inversion symmetry. Our results provide definitive evidence for the low-temperature noncentrosymmetric Td phase from vibrational spectroscopy, and suggest MoTe<span class="hlt">2</span> as an ideal candidate for investigating the temperature-<span class="hlt">induced</span> topological phase transition.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22278179-brown-adipocytes-adrenergically-induced-sub-sub-sub-sub-sub-sub-sub-signalling-erk1-activation-mediated-via-egf-receptor-transactivation','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22278179-brown-adipocytes-adrenergically-induced-sub-sub-sub-sub-sub-sub-sub-signalling-erk1-activation-mediated-via-egf-receptor-transactivation"><span>In brown adipocytes, adrenergically <span class="hlt">induced</span> β{sub 1}-/β{sub 3}-(<span class="hlt">G</span>{sub s})-, α{sub <span class="hlt">2</span>}-(<span class="hlt">G</span>{sub i})- and α{sub 1}-(<span class="hlt">G</span>{sub q})-signalling to Erk1/<span class="hlt">2</span> activation is not mediated via EGF receptor transactivation</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Wang, Yanling; Fälting, Johanna M.; Mattsson, Charlotte L.</p> <p>2013-10-15</p> <p>Brown adipose tissue is unusual in that the neurotransmitter norepinephrine influences cell destiny in ways generally associated with effects of classical growth factors: regulation of cell proliferation, of apoptosis, and progression of differentiation. The norepinephrine effects are mediated through <span class="hlt">G</span>-protein-coupled receptors; further mediation of such stimulation to e.<span class="hlt">g</span>. Erk1/<span class="hlt">2</span> activation is in cell biology in general accepted to occur through transactivation of the EGF receptor (by external or internal pathways). We have examined here the significance of such transactivation in brown adipocytes. Stimulation of mature brown adipocytes with cirazoline (α{sub 1}-adrenoceptor coupled via <span class="hlt">G</span>{sub q}), clonidine (α{sub <span class="hlt">2</span>} via <span class="hlt">G</span>{submore » i}) or CL316243 (β{sub 3} via <span class="hlt">G</span>{sub s}) or via β{sub 1}-receptors significantly activated Erk1/<span class="hlt">2</span>. Pretreatment with the EGF receptor kinase inhibitor AG1478 had, remarkably, no significant effect on Erk1/<span class="hlt">2</span> activation <span class="hlt">induced</span> by any of these adrenergic agonists (although it fully abolished EGF-<span class="hlt">induced</span> Erk1/<span class="hlt">2</span> activation), demonstrating absence of EGF receptor-mediated transactivation. Results with brown preadipocytes (cells in more proliferative states) were not qualitatively different. Joint stimulation of all adrenoceptors with norepinephrine did not result in synergism on Erk1/<span class="hlt">2</span> activation. AG1478 action on EGF-stimulated Erk1/<span class="hlt">2</span> phosphorylation showed a sharp concentration–response relationship (IC{sub 50} 0.3 µM); a minor apparent effect of AG1478 on norepinephrine-stimulated Erk1/<span class="hlt">2</span> phosphorylation showed nonspecific kinetics, implying caution in interpretation of partial effects of AG1478 as reported in other systems. Transactivation of the EGF receptor is clearly not a universal prerequisite for coupling of <span class="hlt">G</span>-protein coupled receptors to Erk1/<span class="hlt">2</span> signalling cascades. - Highlights: • In brown adipocytes, norepinephrine regulates proliferation, apoptosis, differentiation. • EGF receptor transactivation is supposed to</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23546397','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23546397"><span>Hepatoprotective potential of Lavandula coronopifolia extracts against ethanol <span class="hlt">induced</span> oxidative stress-mediated cytotoxicity in Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Farshori, Nida Nayyar; Al-Sheddi, Ebtsam S; Al-Oqail, Mai M; Hassan, Wafaa H B; Al-Khedhairy, Abdulaziz A; Musarrat, Javed; Siddiqui, Maqsood A</p> <p>2015-08-01</p> <p>The present investigations were carried out to study the protective potential of four extracts (namely petroleum ether extract (LCR), chloroform extract (LCM), ethyl acetate extract (LCE), and alcoholic extract (LCL)) of Lavandula coronopifolia on oxidative stress-mediated cell death <span class="hlt">induced</span> by ethanol, a known hepatotoxin in human hapatocellular carcinoma (Hep<span class="hlt">G</span><span class="hlt">2</span>) cells. Cells were pretreated with LCR, LCM, LCE, and LCL extracts (10-50 μ<span class="hlt">g</span>/ml) of L. coronopifolia for 24 h and then ethanol was added and incubated further for 24 h. After the exposure, cell viability using (3-(4,5-dimethylthiazol-<span class="hlt">2</span>-yl)-<span class="hlt">2</span>,5-diphenyltetrazolium bromide) and neutral red uptake assays and morphological changes in Hep<span class="hlt">G</span><span class="hlt">2</span> cells were studied. Pretreatment with various extracts of L. coronpifolia was found to be significantly effective in countering the cytotoxic responses of ethanol. Antioxidant properties of these L. coronopifolia extracts against reactive oxygen species (ROS) generation, lipid peroxidation (LPO), and glutathione (GSH) levels <span class="hlt">induced</span> by ethanol were investigated. Results show that pretreatment with these extracts for 24 h significantly inhibited ROS generation and LPO <span class="hlt">induced</span> and increased the GSH levels reduced by ethanol. The data from the study suggests that LCR, LCM, LCE, and LCL extracts of L. coronopifolia showed hepatoprotective activity against ethanol-<span class="hlt">induced</span> damage in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. However, a comparative study revealed that the LCE extract was found to be the most effective and LCL the least effective. The hepatoprotective effects observed in the study could be associated with the antioxidant properties of these extracts of L. coronopifolia. © The Author(s) 2013.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28262510','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28262510"><span>Regeneration of glutathione by α-lipoic acid via Nrf<span class="hlt">2</span>/ARE signaling pathway alleviates cadmium-<span class="hlt">induced</span> Hep<span class="hlt">G</span><span class="hlt">2</span> cell toxicity.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Jiayu; Zhou, Xue; Wu, Wenbo; Wang, Jiachun; Xie, Hong; Wu, Zhigang</p> <p>2017-04-01</p> <p>Alpha-lipoic acid (α-LA) is an important antioxidant that is capable of regenerating other antioxidants, such as glutathione (GSH). However, the underlying molecular mechanism by which α-LA regenerates GSH remains poorly understood. The current study aimed to investigate whether α-LA regenerates GSH by activation of Nrf<span class="hlt">2</span> to alleviate cadmium-<span class="hlt">induced</span> cytotoxicity in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. In the present study, we found that cadmium <span class="hlt">induced</span> cell death by depletion of GSH through inactivation of Nrf<span class="hlt">2</span>. Addition of α-LA to cadmium-treated cells reactivated Nrf<span class="hlt">2</span> and regenerated GSH through elevating the Nrf<span class="hlt">2</span>-downstream genes γ-glutamate-cysteine ligase (γ-GCL) and GR, both of which are key enzymes for GSH synthesis. However, blocking Nrf<span class="hlt">2</span> with brusatol in the cells co-treated with α-LA and cadmium reduced the mRNA and the protein levels of γ-GCL and GR, thus suppressed GSH regeneration by α-LA. Our results indicated that α-LA activated Nrf<span class="hlt">2</span> signaling pathway, which upregulated the transcription of the enzymes for GSH synthesis and therefore GSH contents to alleviate cadmium-<span class="hlt">induced</span> cytotoxicity in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Copyright © 2017. Published by Elsevier B.V.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/9635591','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/9635591"><span>The <span class="hlt">G</span><span class="hlt">2</span> block <span class="hlt">induced</span> by DNA damage: a caffeine-resistant component independent of Cdc25C, MPM-<span class="hlt">2</span> phosphorylation, and H1 kinase activity.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Barratt, R A; Kao, G; McKenna, W G; Kuang, J; Muschel, R J</p> <p>1998-06-15</p> <p>Treatment of cells with agents that cause DNA damage often results in a delay in <span class="hlt">G</span><span class="hlt">2</span>. There is convincing evidence showing that inhibition of p34cdc<span class="hlt">2</span> kinase activation is involved in the DNA damage-<span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span> delay. In this study, we have demonstrated the existence of an additional pathway, independent of the p34cdc<span class="hlt">2</span> kinase activation pathway, that leads to a <span class="hlt">G</span><span class="hlt">2</span> arrest in etoposide-treated cells. Both the X-ray-<span class="hlt">induced</span> and the etoposide-<span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span> arrest were associated with inhibition of the p34cdc<span class="hlt">2</span> H1 kinase activation pathway as judged by p34cdc<span class="hlt">2</span> H1 kinase activity and phosphorylation of cdc25C. Caffeine treatment restored these activities after either of the treatments. However, the etoposide-treated cells did not resume cycling, revealing the presence of an alternative pathway leading to a <span class="hlt">G</span><span class="hlt">2</span> arrest. To explore the possibility that this additional pathway involved phosphorylation of the MPM-<span class="hlt">2</span> epitope that is shared by a large family of mitotic phosphoproteins, we monitored the phosphorylation status of the MPM-<span class="hlt">2</span> epitope after DNA damage and after treatment with caffeine. Phosphorylation of the MPM-<span class="hlt">2</span> epitope was depressed in both X-ray and etoposide-treated cells, and the depression was reversed by caffeine in both cases. The results indicate that the pathway affecting MPM-<span class="hlt">2</span> epitope phosphorylation is involved in the <span class="hlt">G</span><span class="hlt">2</span> delay caused by DNA damage. However, it is not part of the caffeine-insensitive pathway leading to a <span class="hlt">G</span><span class="hlt">2</span> block seen in etoposide-treated cells.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1376525','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1376525"><span>A parity-<span class="hlt">breaking</span> electronic nematic phase transition in the spin-orbit coupled metal Cd <span class="hlt">2</span>Re <span class="hlt">2</span>O 7</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Harter, J. W.; Zhao, Z. Y.; Yan, J. -Q.</p> <p></p> <p>Strong electron interactions can drive metallic systems toward a variety of well-known symmetry-broken phases, but the instabilities of correlated metals with strong spin-orbit coupling have only recently begun to be explored. We uncovered a multipolar nematic phase of matter in the metallic pyrochlore Cd <span class="hlt">2</span>Re <span class="hlt">2</span>O 7 using spatially resolved second-harmonic optical anisotropy measurements. Like previously discovered electronic nematic phases, this multipolar phase spontaneously <span class="hlt">breaks</span> rotational symmetry while preserving translational invariance. However, it has the distinguishing property of being odd under spatial inversion, which is allowed only in the presence of spin-orbit coupling. By examining the critical behavior of themore » multipolar nematic order parameter, we show that it drives the thermal phase transition near 200 kelvin in Cd <span class="hlt">2</span>Re <span class="hlt">2</span>O 7 and <span class="hlt">induces</span> a parity-<span class="hlt">breaking</span> lattice distortion as a secondary order.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1376525-parity-breaking-electronic-nematic-phase-transition-spin-orbit-coupled-metal-cd2re2o7','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1376525-parity-breaking-electronic-nematic-phase-transition-spin-orbit-coupled-metal-cd2re2o7"><span>A parity-<span class="hlt">breaking</span> electronic nematic phase transition in the spin-orbit coupled metal Cd <span class="hlt">2</span>Re <span class="hlt">2</span>O 7</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Harter, J. W.; Zhao, Z. Y.; Yan, J. -Q.; ...</p> <p>2017-04-21</p> <p>Strong electron interactions can drive metallic systems toward a variety of well-known symmetry-broken phases, but the instabilities of correlated metals with strong spin-orbit coupling have only recently begun to be explored. We uncovered a multipolar nematic phase of matter in the metallic pyrochlore Cd <span class="hlt">2</span>Re <span class="hlt">2</span>O 7 using spatially resolved second-harmonic optical anisotropy measurements. Like previously discovered electronic nematic phases, this multipolar phase spontaneously <span class="hlt">breaks</span> rotational symmetry while preserving translational invariance. However, it has the distinguishing property of being odd under spatial inversion, which is allowed only in the presence of spin-orbit coupling. By examining the critical behavior of themore » multipolar nematic order parameter, we show that it drives the thermal phase transition near 200 kelvin in Cd <span class="hlt">2</span>Re <span class="hlt">2</span>O 7 and <span class="hlt">induces</span> a parity-<span class="hlt">breaking</span> lattice distortion as a secondary order.« less</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li class="active"><span>11</span></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_11 --> <div id="page_12" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li class="active"><span>12</span></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="221"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23291741','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23291741"><span>Sphingoid bases from sea cucumber <span class="hlt">induce</span> apoptosis in human hepatoma Hep<span class="hlt">G</span><span class="hlt">2</span> cells through p-AKT and DR5.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hossain, Zakir; Sugawara, Tatsuya; Hirata, Takashi</p> <p>2013-03-01</p> <p>Biofunctional marine compounds have recently received substantial attention for their nutraceutical characteristics. In this study, we investigated the apoptosis-<span class="hlt">inducing</span> effects of sphingoid bases prepared from sea cucumber using human hepatoma Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Apoptotic effects were determined by cell viability assay, DNA fragmentation assay, caspase-3 and caspase-8 activities. The expression levels of apoptosis-<span class="hlt">inducing</span> death receptor-5 (DR5) and p-AKT were assayed by western blot analysis, and mRNA expression of bax, GADD45 and PPARγ was assayed by quantitative RT-PCR analysis. Sphingoid bases from sea cucumber markedly reduced the cell viability of Hep<span class="hlt">G</span><span class="hlt">2</span> cells. DNA fragmentation indicative of apoptosis was observed in a dose-dependent manner. The expression levels of the apoptosis <span class="hlt">inducer</span> protein Bax were increased by the sphingoid bases from sea cucumber. GADD45, which plays an important role in apoptosis-<span class="hlt">inducing</span> pathways, was markedly upregulated by sphingoid bases from sea cucumber. Upregulation of PPARγ mRNA was also observed during apoptosis <span class="hlt">induced</span> by the sphingoid bases. The expression levels of DR5 and p-AKT proteins were increased and decreased, respectively, as a result of the effects of sphingoid bases from sea cucumber. The results indicate that sphingoid bases from sea cucumber <span class="hlt">induce</span> apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells through upregulation of DR5, Bax, GADD45 and PPARγ and downregulation of p-AKT. Our results show for the first time the functional properties of marine sphingoid bases as <span class="hlt">inducers</span> of apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16309949','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16309949"><span>Evaluation of genotoxic effects of Apitol (cymiazole hydrochloride) in vitro by measurement of sister <span class="hlt">chromatid</span> exchange.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Stanimirovic, Zoran; Stevanovic, Jevrosima; Jovanovic, Slobodan; Andjelkovic, Marko</p> <p>2005-12-30</p> <p>Apitol, with cymiazole hydrochloride as the active ingredient, is used in bee-keeping against the ectoparasitic mite Varroa destructor. The preparation was evaluated for genotoxicity in cultured human peripheral blood lymphocytes. Sister <span class="hlt">chromatid</span> exchange, the mitotic index and the cell proliferation index were determined for three experimental concentrations of Apitol (0.001, 0.01 and 0.1 mg/ml). All concentrations significantly (p < 0.001) increased the mitotic index (MI = 7.35+/-0.18%, 8.31+/-0.20% and 12.33+/-0.25%, respectively), the proliferative index (PI = 1.83+/-0.01, 1.84+/-0.01 and 1.88+/-0.02, respectively) and the frequency of sister <span class="hlt">chromatid</span> exchange (SCE = 8.19+/-1.81, 8.78+/-1.80 and 13.46+/-1.88, respectively), suggesting that cymiazole hydrochloride has genotoxic potential.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23370383','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23370383"><span>Involvement of enniatins-<span class="hlt">induced</span> cytotoxicity in human Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Juan-García, Ana; Manyes, Lara; Ruiz, María-José; Font, Guillermina</p> <p>2013-04-12</p> <p>Enniatins (ENNs) are mycotoxins found in Fusarium fungi and they appear in nature as mixtures of cyclic depsipeptides. The ability to form ionophores in the cell membrane is related to their cytotoxicity. Changes in ion distribution between inner and outer phases of the mitochondria affect to their metabolism, proton gradient, and chemiosmotic coupling, so a mitochondrial toxicity analysis of enniatins is highly recommended because they host the homeostasis required for cellular survival. Two ENNs, ENN A and ENN B on hepatocarcinoma cells (Hep<span class="hlt">G</span><span class="hlt">2</span>) at 1.5 and 3 μM and three exposure times (24, 48 and 72 h) were studied. Flow cytometry was used to examine their effects on cell proliferation, to characterize at which phase of the cell cycle progression the cells were blocked and to study the role of the mitochondrial in ENNs-<span class="hlt">induced</span> apoptosis. In conclusion, apoptosis induction on Hep<span class="hlt">G</span><span class="hlt">2</span> cells allowed to compare cytotoxic effects caused by both ENNs, A and B. It is reported the possible mechanism observed in MMP changes, cell cycle analysis and apoptosis/necrosis, identifying ENN B more toxic than ENN A. Copyright © 2013 Elsevier Ireland Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5376537','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5376537"><span>Protective effects of an ethanol extract of Angelica keiskei against acetaminophen-<span class="hlt">induced</span> hepatotoxicity in Hep<span class="hlt">G</span><span class="hlt">2</span> and HepaRG cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Choi, Yoon-Hee; Lee, Hyun Sook; Chung, Cha-Kwon</p> <p>2017-01-01</p> <p>BACKGROUND/OBJECTIVE Although Angelica keiskei (AK) has widely been utilized for the purpose of general health improvement among Asian, its functionality and mechanism of action. The aim of this study was to determine the protective effect of ethanol extract of AK (AK-Ex) on acute hepatotoxicity <span class="hlt">induced</span> by acetaminophen (AAP) in Hep<span class="hlt">G</span><span class="hlt">2</span> human hepatocellular liver carcinoma cells and HepaRG human hepatic progenitor cells. MATERIALS/METHODS AK-Ex was prepared Hep<span class="hlt">G</span><span class="hlt">2</span> and HepaRG cells were cultured with various concentrations and 30 mM AAP. The protective effects of AK-Ex against AAP-<span class="hlt">induced</span> hepatotoxicity in Hep<span class="hlt">G</span><span class="hlt">2</span> and HepaRG cells were evaluated using 3-[4,5-dimethylthiazol-<span class="hlt">2</span>-yl]-<span class="hlt">2</span>,5-diphenyltetrazolium bromide, lactate dehydrogenase (LDH) assay, flow cytometry, and Western blotting. RESULTS AK-Ex, when administered prior to AAP, increased cell growth and decreased leakage of LDH in a dose-dependent manner in Hep<span class="hlt">G</span><span class="hlt">2</span> and HepaRG cells against AAP-<span class="hlt">induced</span> hepatotoxicity. AK-Ex increased the level of Bcl-<span class="hlt">2</span> and decreased the levels of Bax, Bok and Bik decreased the permeability of the mitochondrial membrane in Hep<span class="hlt">G</span><span class="hlt">2</span> cells intoxicated with AAP. AK-Ex decreased the cleavage of poly (ADP-ribose) polymerase (PARP) and the activation of caspase-9, -7, and -3. CONCLUSIONS These results demonstrate that AK-Ex downregulates apoptosis via intrinsic and extrinsic pathways against AAP-<span class="hlt">induced</span> hepatotoxicity. We suggest that AK could be a useful preventive agent against AAP-<span class="hlt">induced</span> apoptosis in hepatocytes. PMID:28386382</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5615184','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5615184"><span>Alternative Lengthening of Telomeres Mediated by Mitotic DNA Synthesis Engages <span class="hlt">Break-Induced</span> Replication Processes</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Min, Jaewon; Wright, Woodring E.</p> <p>2017-01-01</p> <p>ABSTRACT Alternative lengthening of telomeres (ALT) is a telomerase-independent telomere maintenance mechanism that occurs in a subset of cancers. By analyzing telomerase-positive cells and their human TERC knockout-derived ALT human cell lines, we show that ALT cells harbor more fragile telomeres representing telomere replication problems. ALT-associated replication defects trigger mitotic DNA synthesis (MiDAS) at telomeres in a RAD52-dependent, but RAD51-independent, manner. Telomeric MiDAS is a conservative DNA synthesis process, potentially mediated by <span class="hlt">break-induced</span> replication, similar to type II ALT survivors in Saccharomyces cerevisiae. Replication stresses <span class="hlt">induced</span> by ectopic oncogenic expression of cyclin E, <span class="hlt">G</span>-quadruplexes, or R-loop formation facilitate the ALT pathway and lead to telomere clustering, a hallmark of ALT cancers. The TIMELESS/TIPIN complex suppresses telomere clustering and telomeric MiDAS, whereas the SMC5/6 complex promotes them. In summary, ALT cells exhibit more telomere replication defects that result in persistent DNA damage responses at telomeres, leading to the engagement of telomeric MiDAS (spontaneous mitotic telomere synthesis) that is triggered by DNA replication stress, a potential driver of genomic duplications in cancer. PMID:28760773</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29693750','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29693750"><span>Protein phosphatase <span class="hlt">2</span>A mediates JS-K-<span class="hlt">induced</span> apoptosis by affecting Bcl-<span class="hlt">2</span> family proteins in human hepatocellular carcinoma Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liu, Ling; Huang, Zile; Chen, Jingjing; Wang, Jiangang; Wang, Shuying</p> <p>2018-04-25</p> <p>Protein phosphatase <span class="hlt">2</span>A (PP<span class="hlt">2</span>A) is an important enzyme within various signal transduction pathways. The present study was investigated PP<span class="hlt">2</span>A mediates JS-K-<span class="hlt">induced</span> apoptosis by affecting Bcl-<span class="hlt">2</span> family protein. JS-K showed diverse inhibitory effects in five HCC cell lines, especially Hep<span class="hlt">G</span><span class="hlt">2</span> cells. JS-K caused a dose- and time-dependent reduction in cell viability and increased in levels of LDH release. Meanwhile, JS-K- <span class="hlt">induced</span> apoptosis was characterized by mitochondrial membrane potential reduction, Hoechst 33342 + /PI + dual staining, release of cytochrome c (Cyt c), and activation of cleaved caspase-9/3. Moreover, JS-K-treatment could lead to the activation of protein phosphatase <span class="hlt">2</span>A-C (PP<span class="hlt">2</span>A-C), decrease of anti-apoptotic Bcl-<span class="hlt">2</span> family-protein expression including p-Bcl-<span class="hlt">2</span> (Ser70), Bcl-<span class="hlt">2</span>, Bcl-xL, and Mcl-1 as well as the increase of pro-apoptosis Bcl-<span class="hlt">2</span> family-protein including Bim, Bad, Bax, and Bak. Furthermore, JS-K caused a marked increase of intracellular NO levels while pre-treatment with Carboxy-PTIO (a NO scavenger) reduced the cytotoxicity effects and the apoptosis rate. Meanwhile, pre-treatment with Carboxy-PTIO attenuated the JS-K-<span class="hlt">induced</span> up-regulation of PP<span class="hlt">2</span>A, Cyt c, and cleaved-caspase-9/3 activation. The silencing PP<span class="hlt">2</span>A-C by siRNA could abolish the activation of PP<span class="hlt">2</span>A-C, down-regulation of anti-apoptotic Bcl-<span class="hlt">2</span> family-protein (p-Bcl-<span class="hlt">2</span>, Bcl-<span class="hlt">2</span>, Bcl-xL, and Mcl-1), increase of pro-apoptosis Bcl-<span class="hlt">2</span> family-protein (Bim, Bad, Bax, and Bak) and apoptotic-related protein (Cyt c, cleaved caspase-9/3) that were caused by JS-K in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. In addition, pre-treatment with OA (a PP<span class="hlt">2</span>A inhibitor) also attenuated the above effects <span class="hlt">induced</span> by JS-K. In summary, NO release from JS-K <span class="hlt">induces</span> apoptosis through PP<span class="hlt">2</span>A activation, which contributed to the regulation of Bcl-<span class="hlt">2</span> family proteins. © 2018 Wiley Periodicals, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29595459','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29595459"><span>[Pseudolaric acid B <span class="hlt">induces</span> <span class="hlt">G</span><span class="hlt">2</span>/M arrest and inhibits invasion and migration in Hep<span class="hlt">G</span><span class="hlt">2</span> hepatoma cells].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Li, Shuai; Guo, Lianyi</p> <p>2018-01-01</p> <p>Objective To investigate the mechanisms of pseudolaric acid B (PAB) blocks cell cycle and inhibits invasion and migration in human hepatoma Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Methods The proliferation effect of PAB on Hep<span class="hlt">G</span><span class="hlt">2</span> cells was evaluated by MTT assay. The effect of PAB on the cell cycle of Hep<span class="hlt">G</span><span class="hlt">2</span> cells was analyzed by flow cytometry. Immunofluorescence cytochemical staining was applied to observe the effect of PAB on the α-tubulin polymerization and expression in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Transwell TM chamber invasion assay and wound healing assay were performed to detect the influence of PAB on the migration and invasion ability of Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Western blotting was used to determine the expressions of α-tubulin, E-cadherin and MMP-9 in Hep<span class="hlt">G</span><span class="hlt">2</span> cells after treated with PAB. Results PAB inhibited the proliferation of Hep<span class="hlt">G</span><span class="hlt">2</span> cells in a dose-dependent manner and blocked the cell cycle in <span class="hlt">G</span><span class="hlt">2</span>/M phase. PAB significantly changed the polymerization and decreased the expression of α-tubulin. The capacities of invasion and migration of Hep<span class="hlt">G</span><span class="hlt">2</span> cells treated by PAB were significantly depressed. The protein levels of α-tubulin and MMP-9 decreased while the E-cadherin protein level increased. Conclusion PAB can inhibits the proliferation of Hep<span class="hlt">G</span><span class="hlt">2</span> cells by down-regulating the expression of α-tubulin and influencing its polymerization, arresting Hep<span class="hlt">G</span><span class="hlt">2</span> cells in <span class="hlt">G</span><span class="hlt">2</span>/M phase. Meanwhile, PAB also can inhibit the invasion and migration of Hep<span class="hlt">G</span><span class="hlt">2</span> cells by lowering cytoskeleton α-tubulin and MMP-9, and increasing E-cadherin.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3483531','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3483531"><span>Mangiferin, a Dietary Xanthone Protects Against Mercury-<span class="hlt">Induced</span> Toxicity in Hep<span class="hlt">G</span><span class="hlt">2</span> Cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Agarwala, Sobhika; Rao, B. Nageshwar; Mudholkar, Kaivalya; Bhuwania, Ridhirama; Rao, B. S. Satish</p> <p>2012-01-01</p> <p>Mercury is one of the noxious heavy metal environmental toxicants and is a cause of concern for human exposure. Mangiferin (MGN), a glucosylxanthone found in Mangifera indica, reported to have a wide range of pharmacological properties. The objective of this study was to evaluate the cytoprotective potential of MGN, against mercury chloride (HgCl<span class="hlt">2</span>) <span class="hlt">induced</span> toxicity in Hep<span class="hlt">G</span><span class="hlt">2</span> cell line. The cytoprotective effect of MGN on HgCl<span class="hlt">2</span> <span class="hlt">induced</span> toxicity was assessed by colony formation assay, while antiapoptotic effect by fluorescence microscopy, flow cytometric DNA analysis, and DNA fragmentation pattern assays. Further, the cytoprotective effect of MGN against HgCl<span class="hlt">2</span> toxicity was assessed by using biochemical parameters like reduced glutathione (GSH), glutathione-S-transferase (GST), superoxide dismutase (SOD), catalase (CAT) by spectrophotometrically, mitochondrial membrane potential by flowcytometry and the changes in reactive oxygen species levels by DCFH-DA spectrofluoremetric analysis. A significant increase in the surviving fraction was observed with 50 µM of MGN administered two hours prior to various concentrations of HgCl<span class="hlt">2</span>. Further, pretreatment of MGN significantly decreased the percentage of HgCl<span class="hlt">2</span> <span class="hlt">induced</span> apoptotic cells. Similarly, the levels of ROS generated by the HgCl<span class="hlt">2</span> treatment were inhibited significantly (P < 0.01) by MGN. MGN also significantly (P < 0.01) inhibited the HgCl<span class="hlt">2</span> <span class="hlt">induced</span> decrease in GSH, GST, SOD, and CAT levels at all the post incubation intervals. Our study demonstrated the cytoprotective potential of MGN, which may be attributed to quenching of the ROS generated in the cells due to oxidative stress <span class="hlt">induced</span> by HgCl<span class="hlt">2</span>, restoration of mitochondrial membrane potential and normalization of cellular antioxidant levels. PMID:20629087</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2002PNAS...99.8944Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2002PNAS...99.8944Z"><span>DNA containing Cp<span class="hlt">G</span> motifs <span class="hlt">induces</span> angiogenesis</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zheng, Mei; Klinman, Dennis M.; Gierynska, Malgorzata; Rouse, Barry T.</p> <p>2002-06-01</p> <p>New blood vessel formation in the cornea is an essential step in the pathogenesis of a blinding immunoinflammatory reaction caused by ocular infection with herpes simplex virus (HSV). By using a murine corneal micropocket assay, we found that HSV DNA (which contains a significant excess of potentially bioactive "Cp<span class="hlt">G</span>" motifs when compared with mammalian DNA) <span class="hlt">induces</span> angiogenesis. Moreover, synthetic oligodeoxynucleotides containing Cp<span class="hlt">G</span> motifs attract inflammatory cells and stimulate the release of vascular endothelial growth factor (VEGF), which in turn triggers new blood vessel formation. In vitro, Cp<span class="hlt">G</span> DNA <span class="hlt">induces</span> the J774A.1 murine macrophage cell line to produce VEGF. In vivo Cp<span class="hlt">G-induced</span> angiogenesis was blocked by the administration of anti-mVEGF Ab or the inclusion of "neutralizing" oligodeoxynucleotides that specifically oppose the stimulatory activity of Cp<span class="hlt">G</span> DNA. These findings establish that DNA containing bioactive Cp<span class="hlt">G</span> motifs <span class="hlt">induces</span> angiogenesis, and suggest that Cp<span class="hlt">G</span> motifs in HSV DNA may contribute to the blinding lesions of stromal keratitis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3483857','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3483857"><span>5-AED enhances survival of irradiated mice in a <span class="hlt">G</span>-CSF-dependent manner, stimulates innate immune cell function, reduces radiation-<span class="hlt">induced</span> DNA damage and <span class="hlt">induces</span> genes that modulate cell cycle progression and apoptosis</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Grace, Marcy B.; Singh, Vijay K.; Rhee, Juong G.; Jackson, William E.; Kao, Tzu-Cheg; Whitnall, Mark H.</p> <p>2012-01-01</p> <p>The steroid androst-5-ene-3ß,17ß-diol (5-androstenediol, 5-AED) elevates circulating granulocytes and platelets in animals and humans, and enhances survival during the acute radiation syndrome (ARS) in mice and non-human primates. 5-AED promotes survival of irradiated human hematopoietic progenitors in vitro through induction of Nuclear Factor-κB (NFκB)-dependent Granulocyte Colony-Stimulating Factor (<span class="hlt">G</span>-CSF) expression, and causes elevations of circulating <span class="hlt">G</span>-CSF and interleukin-6 (IL-6). However, the in vivo cellular and molecular effects of 5-AED are not well understood. The aim of this study was to investigate the mechanisms of action of 5-AED administered subcutaneously (s.c.) to mice 24 h before total body γ- or X-irradiation (TBI). We used neutralizing antibodies, flow cytometric functional assays of circulating innate immune cells, analysis of expression of genes related to cell cycle progression, DNA repair and apoptosis, and assessment of DNA strand <span class="hlt">breaks</span> with halo-comet assays. Neutralization experiments indicated endogenous <span class="hlt">G</span>-CSF but not IL-6 was involved in survival enhancement by 5-AED. In keeping with known effects of <span class="hlt">G</span>-CSF on the innate immune system, s.c. 5-AED stimulated phagocytosis in circulating granulocytes and oxidative burst in monocytes. 5-AED <span class="hlt">induced</span> expression of both bax and bcl-<span class="hlt">2</span> in irradiated animals. Cdkn1a and ddb1, but not gadd45a expression, were upregulated by 5-AED in irradiated mice. S.c. 5-AED administration caused decreased DNA strand <span class="hlt">breaks</span> in splenocytes from irradiated mice. Our results suggest 5-AED survival enhancement is <span class="hlt">G</span>-CSF-dependent, and that it stimulates innate immune cell function and reduces radiation-<span class="hlt">induced</span> DNA damage via induction of genes that modulate cell cycle progression and apoptosis. PMID:22843381</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22843381','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22843381"><span>5-AED enhances survival of irradiated mice in a <span class="hlt">G</span>-CSF-dependent manner, stimulates innate immune cell function, reduces radiation-<span class="hlt">induced</span> DNA damage and <span class="hlt">induces</span> genes that modulate cell cycle progression and apoptosis.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Grace, Marcy B; Singh, Vijay K; Rhee, Juong G; Jackson, William E; Kao, Tzu-Cheg; Whitnall, Mark H</p> <p>2012-11-01</p> <p>The steroid androst-5-ene-3ß,17ß-diol (5-androstenediol, 5-AED) elevates circulating granulocytes and platelets in animals and humans, and enhances survival during the acute radiation syndrome (ARS) in mice and non-human primates. 5-AED promotes survival of irradiated human hematopoietic progenitors in vitro through induction of Nuclear Factor-κB (NFκB)-dependent Granulocyte Colony-Stimulating Factor (<span class="hlt">G</span>-CSF) expression, and causes elevations of circulating <span class="hlt">G</span>-CSF and interleukin-6 (IL-6). However, the in vivo cellular and molecular effects of 5-AED are not well understood. The aim of this study was to investigate the mechanisms of action of 5-AED administered subcutaneously (s.c.) to mice 24 h before total body γ- or X-irradiation (TBI). We used neutralizing antibodies, flow cytometric functional assays of circulating innate immune cells, analysis of expression of genes related to cell cycle progression, DNA repair and apoptosis, and assessment of DNA strand <span class="hlt">breaks</span> with halo-comet assays. Neutralization experiments indicated endogenous <span class="hlt">G</span>-CSF but not IL-6 was involved in survival enhancement by 5-AED. In keeping with known effects of <span class="hlt">G</span>-CSF on the innate immune system, s.c. 5-AED stimulated phagocytosis in circulating granulocytes and oxidative burst in monocytes. 5-AED <span class="hlt">induced</span> expression of both bax and bcl-<span class="hlt">2</span> in irradiated animals. Cdkn1a and ddb1, but not gadd45a expression, were upregulated by 5-AED in irradiated mice. S.c. 5-AED administration caused decreased DNA strand <span class="hlt">breaks</span> in splenocytes from irradiated mice. Our results suggest 5-AED survival enhancement is <span class="hlt">G</span>-CSF-dependent, and that it stimulates innate immune cell function and reduces radiation-<span class="hlt">induced</span> DNA damage via induction of genes that modulate cell cycle progression and apoptosis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4317954','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4317954"><span>Double-strand <span class="hlt">break</span> repair-adox: Restoration of suppressed double-strand <span class="hlt">break</span> repair during mitosis <span class="hlt">induces</span> genomic instability</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Terasawa, Masahiro; Shinohara, Akira; Shinohara, Miki</p> <p>2014-01-01</p> <p>Double-strand <span class="hlt">breaks</span> (DSBs) are one of the severest types of DNA damage. Unrepaired DSBs easily <span class="hlt">induce</span> cell death and chromosome aberrations. To maintain genomic stability, cells have checkpoint and DSB repair systems to respond to DNA damage throughout most of the cell cycle. The failure of this process often results in apoptosis or genomic instability, such as aneuploidy, deletion, or translocation. Therefore, DSB repair is essential for maintenance of genomic stability. During mitosis, however, cells seem to suppress the DNA damage response and proceed to the next <span class="hlt">G</span>1 phase, even if there are unrepaired DSBs. The biological significance of this suppression is not known. In this review, we summarize recent studies of mitotic DSB repair and discuss the mechanisms of suppression of DSB repair during mitosis. DSB repair, which maintains genomic integrity in other phases of the cell cycle, is rather toxic to cells during mitosis, often resulting in chromosome missegregation and aberration. Cells have multiple safeguards to prevent genomic instability during mitosis: inhibition of 53BP1 or BRCA1 localization to DSB sites, which is important to promote non-homologous end joining or homologous recombination, respectively, and also modulation of the non-homologous end joining core complex to inhibit DSB repair. We discuss how DSBs during mitosis are toxic and the multiple safeguard systems that suppress genomic instability. PMID:25287622</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5173359','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5173359"><span>Absolute cross-sections for DNA strand <span class="hlt">breaks</span> and crosslinks <span class="hlt">induced</span> by low energy electrons</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Chen, Wenzhuang; Chen, Shiliang; Dong, Yanfang; Cloutier, Pierre; Sanche, Léon</p> <p>2016-01-01</p> <p>Absolute cross sections (CSs) for the interaction of low energy electrons with condensed macromolecules are essential parameters to accurately model ionizing radiation <span class="hlt">induced</span> reactions. To determine CSs for various conformational DNA damage <span class="hlt">induced</span> by <span class="hlt">2</span>–20 eV electrons, we investigated the influence of the attenuation length (AL) and penetration factor (f) using a mathematical model. Solid films of super-coiled plasmid DNA with thicknesses of 10, 15 and 20 nm were irradiated with 4.6, 5.6, 9.6 and 14.6 eV electrons. DNA conformational changes were quantified by gel electrophoresis, and the respective yields were extrapolated from exposure–response curves. The absolute CS, AL and f values were generated by applying the model developed by Rezaee et al. The values of AL were found to lie between 11 and 16 nm with the maximum at 14.6 eV. The absolute CSs for the loss of the supercoiled (LS) configuration and production of crosslinks (CL), single strand <span class="hlt">breaks</span> (SSB) and double strand <span class="hlt">breaks</span> (DSB) <span class="hlt">induced</span> by 4.6, 5.6, 9.6 and 14.6 eV electrons are obtained. The CSs for SSB are smaller, but similar to those for LS, indicating that SSB are the main conformational damage. The CSs for DSB and CL are about one order of magnitude smaller than those of LS and SSB. The value of f is found to be independent of electron energy, which allows extending the absolute CSs for these types of damage within the range <span class="hlt">2</span>–20 eV, from previous measurements of effective CSs. When comparison is possible, the absolute CSs are found to be in good agreement with those obtained from previous similar studies with double-stranded DNA. The high values of the absolute CSs of 4.6 and 9.6 eV provide quantitative evidence for the high efficiency of low energy electrons to <span class="hlt">induce</span> DNA damage via the formation of transient anions. PMID:27878170</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27878170','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27878170"><span>Absolute cross-sections for DNA strand <span class="hlt">breaks</span> and crosslinks <span class="hlt">induced</span> by low energy electrons.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chen, Wenzhuang; Chen, Shiliang; Dong, Yanfang; Cloutier, Pierre; Zheng, Yi; Sanche, Léon</p> <p>2016-12-07</p> <p>Absolute cross sections (CSs) for the interaction of low energy electrons with condensed macromolecules are essential parameters to accurately model ionizing radiation <span class="hlt">induced</span> reactions. To determine CSs for various conformational DNA damage <span class="hlt">induced</span> by <span class="hlt">2</span>-20 eV electrons, we investigated the influence of the attenuation length (AL) and penetration factor (f) using a mathematical model. Solid films of supercoiled plasmid DNA with thicknesses of 10, 15 and 20 nm were irradiated with 4.6, 5.6, 9.6 and 14.6 eV electrons. DNA conformational changes were quantified by gel electrophoresis, and the respective yields were extrapolated from exposure-response curves. The absolute CS, AL and f values were generated by applying the model developed by Rezaee et al. The values of AL were found to lie between 11 and 16 nm with the maximum at 14.6 eV. The absolute CSs for the loss of the supercoiled (LS) configuration and production of crosslinks (CL), single strand <span class="hlt">breaks</span> (SSB) and double strand <span class="hlt">breaks</span> (DSB) <span class="hlt">induced</span> by 4.6, 5.6, 9.6 and 14.6 eV electrons are obtained. The CSs for SSB are smaller, but similar to those for LS, indicating that SSB are the main conformational damage. The CSs for DSB and CL are about one order of magnitude smaller than those of LS and SSB. The value of f is found to be independent of electron energy, which allows extending the absolute CSs for these types of damage within the range <span class="hlt">2</span>-20 eV, from previous measurements of effective CSs. When comparison is possible, the absolute CSs are found to be in good agreement with those obtained from previous similar studies with double-stranded DNA. The high values of the absolute CSs of 4.6 and 9.6 eV provide quantitative evidence for the high efficiency of low energy electrons to <span class="hlt">induce</span> DNA damage via the formation of transient anions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21768208','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21768208"><span>Evaluation of differential representative values between Chinese hamster cells and human lymphocytes in mitomycin C-<span class="hlt">induced</span> cytogenetic assays and caspase-3 activity.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liao, Pei-Hu; Lin, Ruey-Hseng; Yang, Ming-Ling; Li, Yi-Ching; Kuan, Yu-Hsiang</p> <p>2012-03-01</p> <p>Chinese hamster ovary (CHO) cells, its lung fibroblasts (V79), and human lymphocytes are routinely used in in vitro cytogenetic assays, which include micronuclei (MN), sister <span class="hlt">chromatid</span> exchange (SCE), and chromosome aberration (CA) assays. Mitomycin C (MMC), a DNA cross-link alkylating agent, is both an anticancer medicine and a carcinogen. To study the differential representative values of cell types in MMC-treated cytogenetic assays and its upstream factor, cysteine aspartic acid-specific protease (caspase)-3. Among the chosen cell types, lymphocytes expressed the highest sensitivity in all three MMC-<span class="hlt">induced</span> assays, whereas CHO and V79 showed varied sensitivity in different assays. In MN assay, the sensitivity of CHO is higher than or equal to V79; in SCE assay, the sensitivity of CHO is the same as V79; and in CA assay, the sensitivity of CHO is higher than V79. In-depth analysis of CA revealed that in <span class="hlt">chromatid</span> <span class="hlt">breaks</span> and dicentrics formation, lymphocyte was the most sensitive of all and CHO was more sensitive than V79; and in acentrics and interchanges formation, lymphocyte was much more sensitive than the others. Furthermore, we found caspase-3 activity plays an important role in MMC-<span class="hlt">induced</span> cytogenetic assays, with MMC-<span class="hlt">induced</span> caspase-3 activity resulting in more sensitivity in lymphocytes than in CHO and V79. Based on these findings, lymphocyte will make a suitable predictive or representative control reference in cytogenetic assays and caspase-3 activity with its high specificity, positive predictive value, and sensitivity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23238018','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23238018"><span>Colorimetric detection of UV light-<span class="hlt">induced</span> single-strand DNA <span class="hlt">breaks</span> using gold nanoparticles.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kim, Joong Hyun; Chung, Chan Ho; Chung, Bong Hyun</p> <p>2013-02-21</p> <p>We developed a colorimetric method to specifically detect single-strand DNA <span class="hlt">breaks</span> using gold nanoparticles. In our assay, broken DNA cannot stabilize gold nanoparticles to prevent salt-<span class="hlt">induced</span> aggregation as good as intact DNA can, and this effect can be easily observed with the naked eye as a red-to-purple color change.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27837826','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27837826"><span>Novel extraction <span class="hlt">induced</span> by microemulsion <span class="hlt">breaking</span>: a model study for Hg extraction from Brazilian gasoline.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Vicentino, Priscila O; Cassella, Ricardo J</p> <p>2017-01-01</p> <p>This paper proposes a novel approach for the extraction of Hg from Brazilian gasoline samples: extraction <span class="hlt">induced</span> by microemulsion <span class="hlt">breaking</span> (EIMB). In this approach, a microemulsion is formed by mixing the sample with n-propanol and HCl. Afterwards, the microemulsion is destabilized by the addition of water and the two phases are separated: (i) the top phase, containing the residual gasoline and (ii) the bottom phase, containing the extracted analyte in a medium containing water, n-propanol and the ethanol originally present in the gasoline sample. The bottom phase is then collected and the Hg is measured by cold vapor atomic absorption spectrometry (CV-AAS). This model study used Brazilian gasoline samples spiked with Hg (organometallic compound) to optimize the process. Under the optimum extraction conditions, the microemulsion was prepared by mixing 8.7mL of sample with 1.<span class="hlt">2</span>mL of n-propanol and 0.1mL of a 10molL -1 HCl solution. Emulsion <span class="hlt">breaking</span> was <span class="hlt">induced</span> by adding 300µL of deionized water and the bottom phase was collected for the measurement of Hg. Six samples of Brazilian gasoline were spiked with Hg in the organometallic form and recovery percentages in the range of 88-109% were observed. Copyright © 2016 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017E%26PSL.461..176B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017E%26PSL.461..176B"><span>Numerical models for continental <span class="hlt">break</span>-up: Implications for the South Atlantic</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Beniest, A.; Koptev, A.; Burov, E.</p> <p>2017-03-01</p> <p>We propose a mechanism that explains in one unified framework the presence of continental <span class="hlt">break</span>-up features such as failed rift arms and high-velocity and high-density bodies that occur along the South Atlantic rifted continental margins. We used <span class="hlt">2</span>D and 3D numerical models to investigate the impact of thermo-rheological structure of the continental lithosphere and initial plume position on continental rifting and <span class="hlt">break</span>-up processes. <span class="hlt">2</span>D experiments show that <span class="hlt">break</span>-up can be 1) "central", mantle plume-<span class="hlt">induced</span> and directly located above the centre of the mantle anomaly, <span class="hlt">2</span>) "shifted", mantle plume-<span class="hlt">induced</span> and 50 to 200 km shifted from the initial plume location or 3) "distant", self-<span class="hlt">induced</span> due to convection and/or slab-subduction/delamination and 300 to 800 km off-set from the original plume location. With a 3D, perfectly symmetrical and laterally homogeneous setup, the location of continental <span class="hlt">break</span>-up can be shifted hundreds of kilometres from the initial position of the mantle anomaly. We demonstrate that in case of shifted or distant continental <span class="hlt">break</span>-up with respect to the original plume location, multiple features can be explained. Its deep-seated source can remain below the continent at one or both sides of the newly-formed ocean. This mantle material, glued underneath the margins at lower crustal levels, resembles the geometry and location of high velocity/high density bodies observed along the South Atlantic conjugate margins. Impingement of vertically up-welled plume material on the base of the lithosphere results in pre-<span class="hlt">break</span>-up topography variations that are located just above this initial anomaly impingement. This can be interpreted as aborted rift features that are also observed along the rifted margins. When extension continues after continental <span class="hlt">break</span>-up, high strain rates can relocalize. This relocation has been so far attributed to rift jumps. Most importantly, this study shows that there is not one, single rift mode for plume-<span class="hlt">induced</span> crustal <span class="hlt">break</span>-up.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15158121','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15158121"><span>Ca(<span class="hlt">2</span>+)-sensitive tyrosine kinase Pyk<span class="hlt">2</span>/CAK beta-dependent signaling is essential for <span class="hlt">G</span>-protein-coupled receptor agonist-<span class="hlt">induced</span> hypertrophy.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hirotani, Shinichi; Higuchi, Yoshiharu; Nishida, Kazuhiko; Nakayama, Hiroyuki; Yamaguchi, Osamu; Hikoso, Shungo; Takeda, Toshihiro; Kashiwase, Kazunori; Watanabe, Tetsuya; Asahi, Michio; Taniike, Masayuki; Tsujimoto, Ikuko; Matsumura, Yasushi; Sasaki, Terukatsu; Hori, Masatsugu; Otsu, Kinya</p> <p>2004-06-01</p> <p><span class="hlt">G</span>-protein-coupled receptor agonists including endothelin-1 (ET-1) and phenylephrine (PE) <span class="hlt">induce</span> hypertrophy in neonatal ventricular cardiomyocytes. Others and we previously reported that Rac1 signaling pathway plays an important role in this agonist-<span class="hlt">induced</span> cardiomyocyte hypertrophy. In this study reported here, we found that a Ca(<span class="hlt">2</span>+)-sensitive non-receptor tyrosine kinase, proline-rich tyrosine kinase <span class="hlt">2</span> (Pyk<span class="hlt">2</span>)/cell adhesion kinase beta (CAKbeta), is involved in ET-1- and PE-<span class="hlt">induced</span> cardiomyocyte hypertrophy medicated through Rac1 activation. ET-1, PE or the Ca(<span class="hlt">2</span>+) inophore, ionomycin, stimulated a rapid increase in tyrosine phosphorylation of Pyk<span class="hlt">2</span>. The tyrosine phosphorylation of Pyk<span class="hlt">2</span> was suppressed by the Ca(<span class="hlt">2</span>+) chelator, BAPTA. ET-1- or PE-<span class="hlt">induced</span> increases in [(3)H]-leucine incorporation and expression of atrial natriuretic factor and the enhancement of sarcomere organization. Infection of cardiomyocytes with an adenovirus expressing a mutant Pyk<span class="hlt">2</span> which lacked its kinase domain or its ability to bind to c-Src, eliminated ET-1- and PE-<span class="hlt">induced</span> hypertrophic responses. Inhibition of Pyk<span class="hlt">2</span> activation also suppressed Rac1 activation and reactive oxygen species (ROS) production. These findings suggest that the signal transduction pathway leading to hypertrophy involves Ca(<span class="hlt">2</span>+)-<span class="hlt">induced</span> Pyk<span class="hlt">2</span> activation followed by Rac1-dependent ROS production.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28670830','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28670830"><span>Resonant Formation of Strand <span class="hlt">Breaks</span> in Sensitized Oligonucleotides <span class="hlt">Induced</span> by Low-Energy Electrons (0.5-9 eV).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Schürmann, Robin; Tsering, Thupten; Tanzer, Katrin; Denifl, Stephan; Kumar, S V K; Bald, Ilko</p> <p>2017-08-28</p> <p>Halogenated nucleobases are used as radiosensitizers in cancer radiation therapy, enhancing the reactivity of DNA to secondary low-energy electrons (LEEs). LEEs <span class="hlt">induce</span> DNA strand <span class="hlt">breaks</span> at specific energies (resonances) by dissociative electron attachment (DEA). Although halogenated nucleobases show intense DEA resonances at various electron energies in the gas phase, it is inherently difficult to investigate the influence of halogenated nucleobases on the actual DNA strand breakage over the broad range of electron energies at which DEA can take place (<12 eV). By using DNA origami nanostructures, we determined the energy dependence of the strand <span class="hlt">break</span> cross-section for oligonucleotides modified with 8-bromoadenine ( 8Br A). These results were evaluated against DEA measurements with isolated 8Br A in the gas phase. Contrary to expectations, the major contribution to strand <span class="hlt">breaks</span> is from resonances at around 7 eV while resonances at very low energy (<<span class="hlt">2</span> eV) have little influence on strand <span class="hlt">breaks</span>. © 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li class="active"><span>12</span></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_12 --> <div id="page_13" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li class="active"><span>13</span></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="241"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3988499','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3988499"><span>Black rice extract protected Hep<span class="hlt">G</span><span class="hlt">2</span> cells from oxidative stress-<span class="hlt">induced</span> cell death via ERK1/<span class="hlt">2</span> and Akt activation</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Yoon, Jaemin; Ham, Hyeonmi; Sung, Jeehye; Kim, Younghwa; Choi, Youngmin; Lee, Jeom-Sig; Jeong, Heon-Sang; Lee, Junsoo</p> <p>2014-01-01</p> <p>BACKGROUND/OBJECTIVES The objective of this study was to evaluate the protective effect of black rice extract (BRE) on tert-butyl hydroperoxide (TBHP)-<span class="hlt">induced</span> oxidative injury in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. MATERIALS/METHODS Methanolic extract from black rice was evaluated for the protective effect on TBHP-<span class="hlt">induced</span> oxidative injury in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Several biomarkers that modulate cell survival and death including reactive oxygen species (ROS), caspase-3 activity, and related cellular kinases were determined. RESULTS TBHP <span class="hlt">induced</span> cell death and apoptosis by a rapid increase in ROS generation and caspase-3 activity. Moreover, TBHP-<span class="hlt">induced</span> oxidative stress resulted in a transient ERK1/<span class="hlt">2</span> activation and a sustained increase of JNK1/<span class="hlt">2</span> activation. While, BRE pretreatment protects the cells against oxidative stress by reducing cell death, caspase-3 activity, and ROS generation and also by preventing ERKs deactivation and the prolonged JNKs activation. Moreover, pretreatment of BRE increased the activation of ERKs and Akt which are pro-survival signal proteins. However, this effect was blunted in the presence of ERKs and Akt inhibitors. CONCLUSIONS These results suggest that activation of ERKs and Akt pathway might be involved in the cytoprotective effect of BRE against oxidative stress. Our findings provide new insights into the cytoprotective effects and its possible mechanism of black rice against oxidative stress. PMID:24741394</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/391300-induction-sister-chromatid-exchange-presence-gadolinium-dtpa-its-reduction-dimethyl-sulfoxide','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/391300-induction-sister-chromatid-exchange-presence-gadolinium-dtpa-its-reduction-dimethyl-sulfoxide"><span>Induction of sister <span class="hlt">chromatid</span> exchange in the presence of gadolinium-DTPA and its reduction by dimethyl sulfoxide</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Yamazaki, Etsuo; Fukuda, Hozumi; Shibuya, Hitoshi</p> <p></p> <p>The authors investigate the frequency of sister <span class="hlt">chromatid</span> exchange (SCE) after the addition of gadolinium (Gd)-DTPA to venous blood samples. Venous blood was obtained from nonsmokers. Samples were incubated with Gd-DTPA alone or in combination with mitomycin C, cytarabine, and dimethyl sulfoxide (DMSO), and then evaluated for SCEs. The frequency of SCE increased with the concentration of Gd-DTPA and as each chemotherapeutic agent was added. Sister <span class="hlt">chromatid</span> exchange frequencies were lower when the blood was treated with a combination of Gd-DTPA and DMSO compared with Gd-DTPA alone. The increase in frequency of SCE seen after the addition of Gd-DTPA wasmore » decreased by the addition of DMSO, indicating the production of hydroxyl radicals. The effect likely is dissociation-related. 14 refs., 6 tabs.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5752527','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5752527"><span>3D-structured illumination microscopy reveals clustered DNA double-strand <span class="hlt">break</span> formation in widespread γH<span class="hlt">2</span>AX foci after high LET heavy-ion particle radiation</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Hagiwara, Yoshihiko; Niimi, Atsuko; Isono, Mayu; Yamauchi, Motohiro; Yasuhara, Takaaki; Limsirichaikul, Siripan; Oike, Takahiro; Sato, Hiro; Held, Kathryn D.; Nakano, Takashi; Shibata, Atsushi</p> <p>2017-01-01</p> <p>DNA double-strand <span class="hlt">breaks</span> (DSBs) <span class="hlt">induced</span> by ionising radiation are considered the major cause of genotoxic mutations and cell death. While DSBs are dispersed throughout chromatin after X-rays or γ-irradiation, multiple types of DNA damage including DSBs, single-strand <span class="hlt">breaks</span> and base damage can be generated within 1–<span class="hlt">2</span> helical DNA turns, defined as a complex DNA lesion, after high Linear Energy Transfer (LET) particle irradiation. In addition to the formation of complex DNA lesions, recent evidence suggests that multiple DSBs can be closely generated along the tracks of high LET particle irradiation. Herein, by using three dimensional (3D)-structured illumination microscopy, we identified the formation of 3D widespread γH<span class="hlt">2</span>AX foci after high LET carbon-ion irradiation. The large γH<span class="hlt">2</span>AX foci in <span class="hlt">G</span><span class="hlt">2</span>-phase cells encompassed multiple foci of replication protein A (RPA), a marker of DSBs undergoing resection during homologous recombination. Furthermore, we demonstrated by 3D analysis that the distance between two individual RPA foci within γH<span class="hlt">2</span>AX foci was approximately 700 nm. Together, our findings suggest that high LET heavy-ion particles <span class="hlt">induce</span> clustered DSB formation on a scale of approximately 1 μm3. These closely localised DSBs are considered to be a risk for the formation of chromosomal rearrangement after heavy-ion irradiation. PMID:29312614</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29312614','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29312614"><span>3D-structured illumination microscopy reveals clustered DNA double-strand <span class="hlt">break</span> formation in widespread γH<span class="hlt">2</span>AX foci after high LET heavy-ion particle radiation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hagiwara, Yoshihiko; Niimi, Atsuko; Isono, Mayu; Yamauchi, Motohiro; Yasuhara, Takaaki; Limsirichaikul, Siripan; Oike, Takahiro; Sato, Hiro; Held, Kathryn D; Nakano, Takashi; Shibata, Atsushi</p> <p>2017-12-12</p> <p>DNA double-strand <span class="hlt">breaks</span> (DSBs) <span class="hlt">induced</span> by ionising radiation are considered the major cause of genotoxic mutations and cell death. While DSBs are dispersed throughout chromatin after X-rays or γ-irradiation, multiple types of DNA damage including DSBs, single-strand <span class="hlt">breaks</span> and base damage can be generated within 1-<span class="hlt">2</span> helical DNA turns, defined as a complex DNA lesion, after high Linear Energy Transfer (LET) particle irradiation. In addition to the formation of complex DNA lesions, recent evidence suggests that multiple DSBs can be closely generated along the tracks of high LET particle irradiation. Herein, by using three dimensional (3D)-structured illumination microscopy, we identified the formation of 3D widespread γH<span class="hlt">2</span>AX foci after high LET carbon-ion irradiation. The large γH<span class="hlt">2</span>AX foci in <span class="hlt">G</span> <span class="hlt">2</span> -phase cells encompassed multiple foci of replication protein A (RPA), a marker of DSBs undergoing resection during homologous recombination. Furthermore, we demonstrated by 3D analysis that the distance between two individual RPA foci within γH<span class="hlt">2</span>AX foci was approximately 700 nm. Together, our findings suggest that high LET heavy-ion particles <span class="hlt">induce</span> clustered DSB formation on a scale of approximately 1 μm 3 . These closely localised DSBs are considered to be a risk for the formation of chromosomal rearrangement after heavy-ion irradiation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27698266','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27698266"><span>Houttuynia cordata Thunb Promotes Activation of HIF-1A-FOXO3 and MEF<span class="hlt">2</span>A Pathways to <span class="hlt">Induce</span> Apoptosis in Human Hep<span class="hlt">G</span><span class="hlt">2</span> Hepatocellular Carcinoma Cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kim, Jung Min; Hwang, In-Hu; Jang, Ik-Soon; Kim, Min; Bang, In Seok; Park, Soo Jung; Chung, Yun-Jo; Joo, Jong-Cheon; Lee, Min-Goo</p> <p>2017-09-01</p> <p>Houttuynia cordata Thunb ( H cordata), a medicinal plant, has anticancer activity, as it inhibits cell growth and <span class="hlt">induces</span> cell apoptosis in cancer. However, the potential anti-cancer activity and mechanism of H cordata for human liver cancer cells is not well understood. Recently, we identified hypoxia-<span class="hlt">inducible</span> factor (HIF)-1A, Forkhead box (FOX)O3, and MEF<span class="hlt">2</span>A as proapoptotic factors <span class="hlt">induced</span> by H cordata, suggesting that HIF-1A, FOXO3, and MEF<span class="hlt">2</span>A contribute to the apoptosis of Hep<span class="hlt">G</span><span class="hlt">2</span> hepatocellular carcinoma cells. FOXO3 transcription factors regulate target genes involved in apoptosis. H cordata significantly increased the mRNA and protein expression of HIF-1A and FOXO3 and stimulated MEF<span class="hlt">2</span>A expression in addition to increased apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells within 24 hours. Therefore, we determined the potential role of FOXO3 on apoptosis and on H cordata-<span class="hlt">induced</span> MEF<span class="hlt">2</span>A in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. HIF-1A silencing by siRNA attenuated MEF<span class="hlt">2</span>A and H cordata-mediated FOXO3 upregulation in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Furthermore, H cordata-mediated MEF<span class="hlt">2</span>A expression enhanced caspase-3 and caspase-7, which were abolished on silencing FOXO3 with siRNA. In addition, H cordata inhibited growth of human hepatocellular carcinoma xenografts in nude mice. Taken together, our results demonstrate that H cordata enhances HIF-1A/FOXO3 signaling, leading to MEF<span class="hlt">2</span>A upregulation in Hep<span class="hlt">G</span><span class="hlt">2</span> cells, and in parallel, it disturbs the expression of Bcl-<span class="hlt">2</span> family proteins (Bax, Bcl-<span class="hlt">2</span>, and Bcl-xL), which results in apoptosis. Taken together, these findings demonstrate that H cordata promotes the activation of HIF-1A-FOXO3 and MEF<span class="hlt">2</span>A pathways to <span class="hlt">induce</span> apoptosis in human Hep<span class="hlt">G</span><span class="hlt">2</span> hepatocellular carcinoma cells and is, therefore, a promising candidate for antitumor drug development.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5759946','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5759946"><span>Houttuynia cordata Thunb Promotes Activation of HIF-1A–FOXO3 and MEF<span class="hlt">2</span>A Pathways to <span class="hlt">Induce</span> Apoptosis in Human Hep<span class="hlt">G</span><span class="hlt">2</span> Hepatocellular Carcinoma Cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Kim, Jung Min; Hwang, In-Hu; Jang, Ik-Soon; Kim, Min; Bang, In Seok; Park, Soo Jung; Chung, Yun-Jo; Joo, Jong-Cheon; Lee, Min-Goo</p> <p>2016-01-01</p> <p>Houttuynia cordata Thunb (H cordata), a medicinal plant, has anticancer activity, as it inhibits cell growth and <span class="hlt">induces</span> cell apoptosis in cancer. However, the potential anti-cancer activity and mechanism of H cordata for human liver cancer cells is not well understood. Recently, we identified hypoxia-<span class="hlt">inducible</span> factor (HIF)-1A, Forkhead box (FOX)O3, and MEF<span class="hlt">2</span>A as proapoptotic factors <span class="hlt">induced</span> by H cordata, suggesting that HIF-1A, FOXO3, and MEF<span class="hlt">2</span>A contribute to the apoptosis of Hep<span class="hlt">G</span><span class="hlt">2</span> hepatocellular carcinoma cells. FOXO3 transcription factors regulate target genes involved in apoptosis. H cordata significantly increased the mRNA and protein expression of HIF-1A and FOXO3 and stimulated MEF<span class="hlt">2</span>A expression in addition to increased apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells within 24 hours. Therefore, we determined the potential role of FOXO3 on apoptosis and on H cordata–<span class="hlt">induced</span> MEF<span class="hlt">2</span>A in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. HIF-1A silencing by siRNA attenuated MEF<span class="hlt">2</span>A and H cordata–mediated FOXO3 upregulation in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Furthermore, H cordata–mediated MEF<span class="hlt">2</span>A expression enhanced caspase-3 and caspase-7, which were abolished on silencing FOXO3 with siRNA. In addition, H cordata inhibited growth of human hepatocellular carcinoma xenografts in nude mice. Taken together, our results demonstrate that H cordata enhances HIF-1A/FOXO3 signaling, leading to MEF<span class="hlt">2</span>A upregulation in Hep<span class="hlt">G</span><span class="hlt">2</span> cells, and in parallel, it disturbs the expression of Bcl-<span class="hlt">2</span> family proteins (Bax, Bcl-<span class="hlt">2</span>, and Bcl-xL), which results in apoptosis. Taken together, these findings demonstrate that H cordata promotes the activation of HIF-1A–FOXO3 and MEF<span class="hlt">2</span>A pathways to <span class="hlt">induce</span> apoptosis in human Hep<span class="hlt">G</span><span class="hlt">2</span> hepatocellular carcinoma cells and is, therefore, a promising candidate for antitumor drug development. PMID:27698266</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5888797','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5888797"><span>DNA mismatch repair and oligonucleotide end-protection promote base-pair substitution distal from a CRISPR/Cas9-<span class="hlt">induced</span> DNA <span class="hlt">break</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Harmsen, Tim; Klaasen, Sjoerd; van de Vrugt, Henri; te Riele, Hein</p> <p>2018-01-01</p> <p>Abstract Single-stranded oligodeoxyribonucleotide (ssODN)-mediated repair of CRISPR/Cas9-<span class="hlt">induced</span> DNA double-strand <span class="hlt">breaks</span> (DSB) can effectively be used to introduce small genomic alterations in a defined locus. Here, we reveal DNA mismatch repair (MMR) activity is crucial for efficient nucleotide substitution distal from the Cas9-<span class="hlt">induced</span> DNA <span class="hlt">break</span> when the substitution is instructed by the 3′ half of the ssODN. Furthermore, protecting the ssODN 3′ end with phosphorothioate linkages enhances MMR-dependent gene editing events. Our findings can be exploited to optimize efficiencies of nucleotide substitutions distal from the DSB and imply that oligonucleotide-mediated gene editing is effectuated by templated <span class="hlt">break</span> repair. PMID:29447381</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/3365681','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/3365681"><span>Effect of betel chewing on the frequency of sister <span class="hlt">chromatid</span> exchanges in pregnant women and women using oral contraceptives.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ghosh, P K; Ghosh, R</p> <p>1988-06-01</p> <p>The incidence of sister <span class="hlt">chromatid</span> exchange (SCE) was investigated in the lymphocyte chromosomes of betel chewing and non-chewing normal women, pregnant women, and women using oral contraceptives. The frequency of SCE was found to be 7.82 +/- 0.24 and 8.27 +/- 0.27 in non-chewing pregnant women and women using oral contraceptives respectively, which were significantly higher than the mean value of 5.21 +/- 0.18 observed in non-chewing normal women. Betel chewing <span class="hlt">induced</span> higher SCE in pregnant women and women using oral contraceptives, the frequencies being 11.79 +/- 0.38 and 12.51 +/- 0.44, respectively, which were significantly higher than the SCE frequency of 6.28 +/- 0.21 found in normal betel chewing females.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApSS..419..875T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApSS..419..875T"><span>Photo-<span class="hlt">induced</span> CO<span class="hlt">2</span> reduction by CH4/H<span class="hlt">2</span>O to fuels over Cu-modified <span class="hlt">g</span>-C3N4 nanorods under simulated solar energy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tahir, Beenish; Tahir, Muhammad; Amin, Nor Aishah Saidina</p> <p>2017-10-01</p> <p>Copper modified polymeric graphitic carbon nitride (Cu/<span class="hlt">g</span>-C3N4) nanorods for photo-<span class="hlt">induced</span> CO<span class="hlt">2</span> conversion with methane (CH4) and water (H<span class="hlt">2</span>O) as reducing system under simulated solar energy has been investigated. The nanocatalysts, synthesized by pyrolysis and sonication, were characterized by XRD, FTIR, Raman analysis, XPS, SEM, N<span class="hlt">2</span> adsorption-desorption and PL spectroscopy. The presence of Cu<span class="hlt">2</span>+ ions over the <span class="hlt">g</span>-C3N4 structure inhibited charge carriers recombination process. The results indicated that photo-activity and selectivity of Cu/<span class="hlt">g</span>-C3N4 photo-catalyst for CO<span class="hlt">2</span> reduction greatly dependent on the type of CO<span class="hlt">2</span>-reduction system. CO<span class="hlt">2</span> was efficiently converted to CH4 and CH3OH with traces of C<span class="hlt">2</span>H4 and C<span class="hlt">2</span>H6 hydrocarbons in the CO<span class="hlt">2</span>-water system. The yield of the main product, CH4 over 3 wt.% Cu/<span class="hlt">g</span>-C3N4 was 109 μmole <span class="hlt">g</span>-cata.-1 h-1 under visible light irradiation, significantly higher than the pure <span class="hlt">g</span>-C3N4 catalyst (60 μmole/<span class="hlt">g</span>.cat). In photo-<span class="hlt">induced</span> CO<span class="hlt">2</span>-CH4 reaction, CO and H<span class="hlt">2</span> were detected as the main products with smaller amount of hydrocarbons. The highest efficiency was detected over 3 wt.%Cu-loading of <span class="hlt">g</span>-C3N4 and at optimal CH4/CO<span class="hlt">2</span> feed ratio of 1.0. The maximum yield of CO and H<span class="hlt">2</span> detected were 142 and 76 μmole <span class="hlt">g</span>-catal.-1 h-1, respectively at selectivity 66.6% and 32.5%, respectively. Significantly enhanced CO<span class="hlt">2</span>/CH4 reduction over Cu/<span class="hlt">g</span>-C3N4 was attributed to its polymeric structure with efficient charge transfer property and inhibited charges recombination rate. A proposed photo-<span class="hlt">induced</span> reaction mechanism, corroborated with the experimental data, was also deliberated.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22977515','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22977515"><span>Trifluorothymidine exhibits potent antitumor activity via the induction of DNA double-strand <span class="hlt">breaks</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Suzuki, Norihiko; Nakagawa, Fumio; Nukatsuka, Mamoru; Fukushima, Masakazu</p> <p>2011-05-01</p> <p>TAS-102 is an oral anticancer drug composed of trifluorothymidine (TFT) and TPI (an inhibitor of thymidine phosphorylase that strongly inhibits the biodegradation of TFT). Similar to 5-fluorouracil (5FU) and 5-fluoro-<span class="hlt">2</span>'-deoxyuridine (FdUrd), TFT also inhibits thymidylate synthase (TS), a rate-limiting enzyme of DNA biosynthesis, and is incorporated into DNA. TFT exhibits an anticancer effect on colorectal cancer cells that have acquired 5FU and/or FdUrd resistance as a result of the overexpression of TS. Therefore, we examined the mode of action of TFT-<span class="hlt">induced</span> DNA damage after its incorporation into DNA. When HeLa cells were treated with TFT, the number of ring-open aldehyde forms at apurinic/apyrimidinic sites increased in a dose-dependent manner, although we previously reported that no detectable excisions of TFT paired to adenine were observed using uracil DNA glycosylases, thymine DNA glycosylase or methyl-Cp<span class="hlt">G</span> binding domain 4 and HeLa whole cell extracts. To investigate the functional mechanism of TFT-<span class="hlt">induced</span> DNA damage, we measured the phosphorylation of ATR, ATM, BRCA<span class="hlt">2</span>, chk1 and chk<span class="hlt">2</span> in nuclear extracts of HeLa cells after 0, 24, 48 or 72 h of exposure to an IC(50) concentration of TFT, FdUrd or 5FU using Western blot analysis or an enzyme-linked immunosorbent assay (ELISA). Unlike FdUrd and 5FU, TFT resulted in an earlier phosphorylation of ATR and chk1 proteins after only 24 h of exposure, while phosphorylated ATM, BRCA<span class="hlt">2</span> and chk<span class="hlt">2</span> proteins were detected after more than 48 h of exposure to TFT. These results suggest that TFT causes single-strand <span class="hlt">breaks</span> followed by double-strand <span class="hlt">breaks</span> in the DNA of TFT-treated cells. TFT (as TAS-102) showed a more potent antitumor activity than oral 5FU on CO-3 colon cancer xenografts in mice, and such antitumor potency was supported by the increased number of double-strand <span class="hlt">breaks</span> occurring after single-strand <span class="hlt">breaks</span> in the DNA of the TFT-treated tumors. These results suggest that TFT causes single-strand <span class="hlt">breaks</span> after its</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22118662','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22118662"><span>β<span class="hlt">2</span>-adrenoceptor blockage <span class="hlt">induces</span> <span class="hlt">G</span>1/S phase arrest and apoptosis in pancreatic cancer cells via Ras/Akt/NFκB pathway.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Dong; Ma, Qingyong; Wang, Zheng; Zhang, Min; Guo, Kun; Wang, Fengfei; Wu, Erxi</p> <p>2011-11-26</p> <p>Smoking and stress, pancreatic cancer (PanCa) risk factors, stimulate nitrosamine 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone (NNK) and catecholamines production respectively. NNK and catecholamine bind the β-adrenoceptors and <span class="hlt">induce</span> PanCa cell proliferation; and we have previously suggested that β-adrenergic antagonists may suppress proliferation and invasion and stimulate apoptosis in PanCa. To clarify the mechanism of apoptosis <span class="hlt">induced</span> by β<span class="hlt">2</span>-adrenergic antagonist, we hypothesize that blockage of the β<span class="hlt">2</span>-adrenoceptor could <span class="hlt">induce</span> <span class="hlt">G</span>1/S phase arrest and apoptosis and Ras may be a key player in PanCa cells. The β1 and β<span class="hlt">2</span>-adrenoceptor proteins were detected on the cell surface of PanCa cells from pancreatic carcinoma specimen samples by immunohistochemistry. The β<span class="hlt">2</span>-adrenergic antagonist ICI118,551 significantly <span class="hlt">induced</span> <span class="hlt">G</span>1/S phase arrest and apoptosis compared with the β1-adrenergic antagonist metoprolol, which was determined by the flow cytometry assay. β<span class="hlt">2</span>-adrenergic antagonist therapy significantly suppressed the expression of extracellular signal-regulated kinase, Akt, Bcl-<span class="hlt">2</span>, cyclin D1, and cyclin E and <span class="hlt">induced</span> the activation of caspase-3, caspase-9 and Bax by Western blotting. Additionally, the β<span class="hlt">2</span>-adrenergic antagonist reduced the activation of NFκB in vitro cultured PanCa cells. The blockage of β<span class="hlt">2</span>-adrenoceptor markedly <span class="hlt">induced</span> PanCa cells to arrest at <span class="hlt">G</span>1/S phase and consequently resulted in cell death, which is possibly due to that the blockage of β<span class="hlt">2</span>-adrenoceptor inhibited NFκB, extracellular signal-regulated kinase, and Akt pathways. Therefore, their upstream molecule Ras may be a key factor in the β<span class="hlt">2</span>-adrenoceptor antagonist <span class="hlt">induced</span> <span class="hlt">G</span>1/S phase arrest and apoptosis in PanCa cells. The new pathway discovered in this study may provide an effective therapeutic strategy for PanCa.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16467335','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16467335"><span>Role of CYP<span class="hlt">2</span>E1 immunoglobulin <span class="hlt">G</span>4 subclass antibodies and complement in pathogenesis of idiosyncratic drug-<span class="hlt">induced</span> hepatitis.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Njoku, Dolores B; Mellerson, Jenelle L; Talor, Monica V; Kerr, Douglas R; Faraday, Nauder R; Outschoorn, Ingrid; Rose, Noel R</p> <p>2006-02-01</p> <p>Idiosyncratic drug-<span class="hlt">induced</span> hepatitis (IDDIH) is the third most common cause for acute liver failure in the United States. Previous studies have attempted to identify susceptible patients or early stages of disease with various degrees of success. To determine if total serum immunoglobulin subclasses, CYP<span class="hlt">2</span>E1-specific subclass autoantibodies, complement components, or immune complexes could distinguish persons with IDDIH from others exposed to drugs, we studied persons exposed to halogenated volatile anesthetics, which have been associated with IDDIH and CYP<span class="hlt">2</span>E1 autoantibodies. We found that patients with anesthetic-<span class="hlt">induced</span> IDDIH had significantly elevated levels of CYP<span class="hlt">2</span>E1-specific immunoglobulin <span class="hlt">G</span>4 (Ig<span class="hlt">G</span>4) autoantibodies, while anesthetic-exposed healthy persons had significantly elevated levels of CYP<span class="hlt">2</span>E1-specific Ig<span class="hlt">G</span>1 autoantibodies. Anesthetic IDDIH patients had significantly lower levels of C4a, C3a, and C5a compared to anesthetic-exposed healthy persons. C1q- and C3d-containing immune complexes were significantly elevated in anesthetic-exposed persons. In conclusion, our data suggest that anesthetic-exposed persons develop CYP<span class="hlt">2</span>E1-specific Ig<span class="hlt">G</span>1 autoantibodies which may form detectable circulating immune complexes subsequently cleared by classical pathway activation of the complement system. Persons susceptible to anesthetic-<span class="hlt">induced</span> IDDIH develop CYP<span class="hlt">2</span>E1-specific Ig<span class="hlt">G</span>4 autoantibodies which form small, nonprecipitating immune complexes that escape clearance because of their size or by direct inhibition of complement activation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1391926','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1391926"><span>Role of CYP<span class="hlt">2</span>E1 Immunoglobulin <span class="hlt">G</span>4 Subclass Antibodies and Complement in Pathogenesis of Idiosyncratic Drug-<span class="hlt">Induced</span> Hepatitis</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Njoku, Dolores B.; Mellerson, Jenelle L.; Talor, Monica V.; Kerr, Douglas R.; Faraday, Nauder R.; Outschoorn, Ingrid; Rose, Noel R.</p> <p>2006-01-01</p> <p>Idiosyncratic drug-<span class="hlt">induced</span> hepatitis (IDDIH) is the third most common cause for acute liver failure in the United States. Previous studies have attempted to identify susceptible patients or early stages of disease with various degrees of success. To determine if total serum immunoglobulin subclasses, CYP<span class="hlt">2</span>E1-specific subclass autoantibodies, complement components, or immune complexes could distinguish persons with IDDIH from others exposed to drugs, we studied persons exposed to halogenated volatile anesthetics, which have been associated with IDDIH and CYP<span class="hlt">2</span>E1 autoantibodies. We found that patients with anesthetic-<span class="hlt">induced</span> IDDIH had significantly elevated levels of CYP<span class="hlt">2</span>E1-specific immunoglobulin <span class="hlt">G</span>4 (Ig<span class="hlt">G</span>4) autoantibodies, while anesthetic-exposed healthy persons had significantly elevated levels of CYP<span class="hlt">2</span>E1-specific Ig<span class="hlt">G</span>1 autoantibodies. Anesthetic IDDIH patients had significantly lower levels of C4a, C3a, and C5a compared to anesthetic-exposed healthy persons. C1q- and C3d-containing immune complexes were significantly elevated in anesthetic-exposed persons. In conclusion, our data suggest that anesthetic-exposed persons develop CYP<span class="hlt">2</span>E1-specific Ig<span class="hlt">G</span>1 autoantibodies which may form detectable circulating immune complexes subsequently cleared by classical pathway activation of the complement system. Persons susceptible to anesthetic-<span class="hlt">induced</span> IDDIH develop CYP<span class="hlt">2</span>E1-specific Ig<span class="hlt">G</span>4 autoantibodies which form small, nonprecipitating immune complexes that escape clearance because of their size or by direct inhibition of complement activation. PMID:16467335</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016cosp...41E.477D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016cosp...41E.477D"><span>How do accretion discs <span class="hlt">break</span>?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dogan, Suzan</p> <p>2016-07-01</p> <p>Accretion discs are common in binary systems, and they are often found to be misaligned with respect to the binary orbit. The gravitational torque from a companion <span class="hlt">induces</span> nodal precession in misaligned disc orbits. In this study, we first calculate whether this precession is strong enough to overcome the internal disc torques communicating angular momentum. We compare the disc precession torque with the disc viscous torque to determine whether the disc should warp or <span class="hlt">break</span>. For typical parameters precession wins: the disc <span class="hlt">breaks</span> into distinct planes that precess effectively independently. To check our analytical findings, we perform 3D hydrodynamical numerical simulations using the PHANTOM smoothed particle hydrodynamics code, and confirm that disc <span class="hlt">breaking</span> is widespread and enhances accretion on to the central object. For some inclinations, the disc goes through strong Kozai cycles. Disc <span class="hlt">breaking</span> promotes markedly enhanced and variable accretion and potentially produces high-energy particles or radiation through shocks. This would have significant implications for all binary systems: e.<span class="hlt">g</span>. accretion outbursts in X-ray binaries and fuelling supermassive black hole (SMBH) binaries. The behaviour we have discussed in this work is relevant to a variety of astrophysical systems, for example X-ray binaries, where the disc plane may be tilted by radiation warping, SMBH binaries, where accretion of misaligned gas can create effectively random inclinations and protostellar binaries, where a disc may be misaligned by a variety of effects such as binary capture/exchange, accretion after binary formation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28343997','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28343997"><span>DNA double-strand <span class="hlt">breaks</span> and Aurora B mislocalization <span class="hlt">induced</span> by exposure of early mitotic cells to H<span class="hlt">2</span>O<span class="hlt">2</span> appear to increase chromatin bridges and resultant cytokinesis failure.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cho, Min-Guk; Ahn, Ju-Hyun; Choi, Hee-Song; Lee, Jae-Ho</p> <p>2017-07-01</p> <p>Aneuploidy, an abnormal number of chromosomes that is a hallmark of cancer cells, can arise from tetraploid/binucleated cells through a failure of cytokinesis. Reactive oxygen species (ROS) have been implicated in various diseases, including cancer. However, the nature and role of ROS in cytokinesis progression and related mechanisms has not been clearly elucidated. Here, using time-lapse analysis of asynchronously growing cells and immunocytochemical analyses of synchronized cells, we found that hydrogen peroxide (H <span class="hlt">2</span> O <span class="hlt">2</span> ) treatment at early mitosis (primarily prometaphase) significantly <span class="hlt">induced</span> cytokinesis failure. Cytokinesis failure and the resultant formation of binucleated cells containing nucleoplasmic bridges (NPBs) seemed to be caused by increases in DNA double-strand <span class="hlt">breaks</span> (DSBs) and subsequent unresolved chromatin bridges. We further found that H <span class="hlt">2</span> O <span class="hlt">2</span> <span class="hlt">induced</span> mislocalization of Aurora B during mitosis. All of these effects were attenuated by pretreatment with N-acetyl-L-cysteine (NAC) or overexpression of Catalase. Surprisingly, the PARP inhibitor PJ34 also reduced H <span class="hlt">2</span> O <span class="hlt">2</span> -<span class="hlt">induced</span> Aurora B mislocalization and binucleated cell formation. Results of parallel experiments with etoposide, a topoisomerase IIα inhibitor that triggers DNA DSBs, suggested that both DNA DSBs and Aurora B mislocalization contribute to chromatin bridge formation. Aurora B mislocalization also appeared to weaken the "abscission checkpoint". Finally, we showed that KRAS-<span class="hlt">induced</span> binucleated cell formation appeared to be also H <span class="hlt">2</span> O <span class="hlt">2</span> -dependent. In conclusion, we propose that a ROS, mainly H <span class="hlt">2</span> O <span class="hlt">2</span> increases binucleation through unresolved chromatin bridges caused by DNA damage and mislocalization of Aurora B, the latter of which appears to augment the effect of DNA damage on chromatin bridge formation. Copyright © 2017 Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28877783','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28877783"><span><span class="hlt">G</span>-Protein-Coupled Estrogen Receptor Antagonist <span class="hlt">G</span>15 Decreases Estrogen-<span class="hlt">Induced</span> Development of Non-Small Cell Lung Cancer.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liu, Changyu; Liao, Yongde; Fan, Sheng; Fu, Xiangning; Xiong, Jing; Zhou, Sheng; Zou, Man; Wang, Jianmiao</p> <p>2017-08-25</p> <p><span class="hlt">G</span>-protein-coupled estrogen receptor (GPER) was found to promote Non-small cell lung cancer (NSCLC) by estrogen, indicating the potential necessity of inhibiting GPER by selective antagonist. This study was performed to elucidate the function of GPER selective inhibitor <span class="hlt">G</span>15 in NSCLC development. Cytoplasmic GPER (cGPER) and nuclear GPER (nGPER) were detected by immunohistochemical analysis in NSCLC samples. The relation of GPER and estrogen receptor β (ERβ) expression and correlation between GPER, ERβ and clinical factors were analyzed. The effects of activating GPER and function of <span class="hlt">G</span>15 were analyzed in proliferation of A549, H1793 cell lines and development of urethane-<span class="hlt">induced</span> adenocarcinoma. Overexpression of cGPER and nGPER was detected in 80.49% (120/150) and 52.00% (78/150) of the NSCLC samples. High expression of GPER related with higher stages, poorer differentiation and high expression of ERβ. Protein level of GPER in A549 and H1793 cell lines increased by treatment of E<span class="hlt">2</span>, <span class="hlt">G</span>1 (GPER agonist) or Ful (fulvestrant, ERβ antagonist), and decreased by <span class="hlt">G</span>15. Administration with <span class="hlt">G</span>15 reversed the E<span class="hlt">2</span>- or <span class="hlt">G</span>1-<span class="hlt">induced</span> cell growth by inhibiting GPER. In urethane-<span class="hlt">induced</span> adenocarcinoma mice, number of tumor nodules and tumor index increased in E<span class="hlt">2</span> or <span class="hlt">G</span>1 group and decreased by treatment of <span class="hlt">G</span>15. These findings deomonstrate that using of <span class="hlt">G</span>15 to block GPER signaling may be considered as a new therapeutic target in NSCLC.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28300666','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28300666"><span>Curcumin attenuates BPA-<span class="hlt">induced</span> insulin resistance in Hep<span class="hlt">G</span><span class="hlt">2</span> cells through suppression of JNK/p38 pathways.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Geng, Shanshan; Wang, Shijia; Zhu, Weiwei; Xie, Chunfeng; Li, Xiaoting; Wu, Jieshu; Zhu, Jianyun; Jiang, Ye; Yang, Xue; Li, Yuan; Chen, Yue; Wang, Xiaoqian; Meng, Yu; Zhu, Mingming; Wu, Rui; Huang, Cong; Zhong, Caiyun</p> <p>2017-04-15</p> <p>Bisphenol A (BPA) is an artificial environmental endocrine disrupting chemicals. Accumulating evidence indicates that exposure to BPA contributes to insulin resistance through diverse mechanism including inflammation and oxidative stress. Previous studies have suggested curcumin as a safe phytochemical which can improve obesity-related insulin resistance, inflammation and oxidative stress. The present study aimed to investigate the ability of curcumin to prevent BPA-<span class="hlt">induced</span> insulin resistance in vitro and the underlying mechanism. Following the establishmet of in vitro insulin resistance via BPA treatment in human liver Hep<span class="hlt">G</span><span class="hlt">2</span> cells, the protective effects of curcumin were determiend. We showed that treatment of Hep<span class="hlt">G</span><span class="hlt">2</span> cells with 100nM BPA for 5days <span class="hlt">induced</span> significantly decreased glucose consumption, impaired insulin signaling, elevation of pro-inflammatory cytokines and oxidative stress, and activation of signaling pathways; inhibition of JNK and p38 pathways, but not ERK nor NF-κB pathways, improved glucose consumption and insulin signaling in BPA-treated Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Moreover, we revealed that curcumin effectively attenuated the spectrum of effects of BPA-triggered insulin resistance, whereas pretreatment with JNK and p38 agonist anisomycin could significantly compensate the effects caused by curcumin. These data illustrated the role of JNK/p38 activation in BPA-<span class="hlt">induced</span> insulin resistance and suggested curcumin as a promising candidate for the intervention of BPA-<span class="hlt">induced</span> insulin resistance. Copyright © 2017 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29466339','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29466339"><span>Intragenic origins due to short <span class="hlt">G</span>1 phases underlie oncogene-<span class="hlt">induced</span> DNA replication stress.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Macheret, Morgane; Halazonetis, Thanos D</p> <p>2018-03-01</p> <p>Oncogene-<span class="hlt">induced</span> DNA replication stress contributes critically to the genomic instability that is present in cancer. However, elucidating how oncogenes deregulate DNA replication has been impeded by difficulty in mapping replication initiation sites on the human genome. Here, using a sensitive assay to monitor nascent DNA synthesis in early S phase, we identified thousands of replication initiation sites in cells before and after induction of the oncogenes CCNE1 and MYC. Remarkably, both oncogenes <span class="hlt">induced</span> firing of a novel set of DNA replication origins that mapped within highly transcribed genes. These ectopic origins were normally suppressed by transcription during <span class="hlt">G</span>1, but precocious entry into S phase, before all genic regions had been transcribed, allowed firing of origins within genes in cells with activated oncogenes. Forks from oncogene-<span class="hlt">induced</span> origins were prone to collapse, as a result of conflicts between replication and transcription, and were associated with DNA double-stranded <span class="hlt">break</span> formation and chromosomal rearrangement breakpoints both in our experimental system and in a large cohort of human cancers. Thus, firing of intragenic origins caused by premature S phase entry represents a mechanism of oncogene-<span class="hlt">induced</span> DNA replication stress that is relevant for genomic instability in human cancer.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JMMM..449..440S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JMMM..449..440S"><span>Field-<span class="hlt">induced</span> cluster spin glass and inverse symmetry <span class="hlt">breaking</span> enhanced by frustration</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schmidt, M.; Zimmer, F. M.; Magalhaes, S. G.</p> <p>2018-03-01</p> <p>We consider a cluster disordered model to study the interplay between short- and long-range interactions in geometrically frustrated spin systems under an external magnetic field (h). In our approach, the intercluster long-range disorder (J) is analytically treated to get an effective cluster model that is computed exactly. The clusters follow a checkerboard lattice with first-neighbor (J1) and second-neighbor (J<span class="hlt">2</span>) interactions. We find a reentrant transition from the cluster spin-glass (CSG) state to a paramagnetic (PM) phase as the temperature decreases for a certain range of h. This inverse symmetry <span class="hlt">breaking</span> (ISB) appears as a consequence of both quenched disorder with frustration and h, that introduce a CSG state with higher entropy than the polarized PM phase. The competitive scenario introduced by antiferromagnetic (AF) short-range interactions increases the CSG state entropy, leading to continuous ISB transitions and enhancing the ISB regions, mainly in the geometrically frustrated case (J1 =J<span class="hlt">2</span>). Remarkably, when strong AF intracluster couplings are present, field-<span class="hlt">induced</span> CSG phases can be found. These CSG regions are strongly related to the magnetization plateaus observed in this cluster disordered system. In fact, it is found that each field-<span class="hlt">induced</span> magnetization jump brings a CSG region. We notice that geometrical frustration, as well as cluster size, play an important role in the magnetization plateaus and, therefore, are also relevant in the field-<span class="hlt">induced</span> glassy states. Our findings suggest that competing interactions support ISB and field-<span class="hlt">induced</span> CSG phases in disordered cluster systems under an external magnetic field.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=62118&keyword=Colorectal+AND+Cancer&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50','EPA-EIMS'); return false;" href="https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=62118&keyword=Colorectal+AND+Cancer&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50"><span>ANALYSIS OF IN VITRO AND IN VIVO DNA STRAND <span class="hlt">BREAKS</span> <span class="hlt">INDUCED</span> BY TRIHALOMETHANES (THMS)</span></a></p> <p><a target="_blank" href="http://oaspub.epa.gov/eims/query.page">EPA Science Inventory</a></p> <p></p> <p></p> <p>Analysis of In Vitro and In Vivo DNA Strand <span class="hlt">Breaks</span> <span class="hlt">Induced</span> by Trihalomethanes (TRMs)<br> <br>The THMs are the most widely distributed and the most concentrated of the cWorine disinfection by-products (D BPs) found in finished drinking water. All of the THMs, cWoroform (CHCI3), br...</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li class="active"><span>13</span></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_13 --> <div id="page_14" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li class="active"><span>14</span></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="261"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3400450','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3400450"><span>The autumn effect: timing of physical dormancy <span class="hlt">break</span> in seeds of two winter annual species of Geraniaceae by a stepwise process</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Gama-Arachchige, N. S.; Baskin, J. M.; Geneve, R. L.; Baskin, C. C.</p> <p>2012-01-01</p> <p>Background and Aims The involvement of two steps in the physical dormancy (PY)-<span class="hlt">breaking</span> process previously has been demonstrated in seeds of Fabaceae and Convolvulaceae. Even though there is a claim for a moisture-controlled stepwise PY-<span class="hlt">breaking</span> in some species of Geraniaceae, no study has evaluated the role of temperature in the PY-<span class="hlt">breaking</span> process in this family. The aim of this study was to determine whether a temperature-controlled stepwise PY-<span class="hlt">breaking</span> process occurs in seeds of the winter annuals Geranium carolinianum and <span class="hlt">G</span>. dissectum. Methods Seeds of <span class="hlt">G</span>. carolinianum and <span class="hlt">G</span>. dissectum were stored under different temperature regimes to test the effect of storage temperature on PY-<span class="hlt">break</span>. The role of temperature and moisture regimes in regulating PY-<span class="hlt">break</span> was investigated by treatments simulating natural conditions. Greenhouse (non-heated) experiments on seed germination and burial experiments (outdoors) were carried out to determine the PY-<span class="hlt">breaking</span> behaviour in the natural habitat. Key Results Irrespective of moisture conditions, sensitivity to the PY-<span class="hlt">breaking</span> step in seeds of <span class="hlt">G</span>. carolinianum was <span class="hlt">induced</span> at temperatures ≥20 °C, and exposure to temperatures ≤20 °C made the sensitive seeds permeable. Sensitivity of seeds increased with time. In <span class="hlt">G</span>. dissectum, PY-<span class="hlt">break</span> occurred at temperatures ≥20 °C in a single step under constant wet or dry conditions and in two steps under alternate wet–dry conditions if seeds were initially kept wet. Conclusions Timing of seed germination with the onset of autumn can be explained by PY-<span class="hlt">breaking</span> processes involving (a) two temperature-dependent steps in <span class="hlt">G</span>. carolinianum and (b) one or two moisture-dependent step(s) along with the inability to germinate under high temperatures in <span class="hlt">G</span>. dissectum. Geraniaceae is the third of 18 families with PY in which a two-step PY-<span class="hlt">breaking</span> process has been demonstrated. PMID:22684684</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15916875','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15916875"><span>116 kDa glycoprotein isolated from Ulmus davidiana Nakai (UDN) inhibits glucose/glucose oxidase (<span class="hlt">G/GO)-induced</span> apoptosis in BNL CL.<span class="hlt">2</span> cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ko, Jeong-Hyeon; Lee, Sei-Jung; Lim, Kye-Taek</p> <p>2005-09-14</p> <p>Ulmus davidiana Nakai (UDN) has been used in folk medicine for its anti-inflammatory activity. In the present study, we investigated the antiapoptotic effect of UDN glycoprotein in glucose/glucose oxidase (<span class="hlt">G/GO)-induced</span> BNL CL.<span class="hlt">2</span> cells. To evaluate the antiapoptotic effect of UDN glycoprotein, experiments were carried out using Western blot analysis for nuclear factor-kappa B (NF-kappaB), caspase-3, and poly(ADP-ribose) polymerase (PARP). We also examined nitric oxide (NO) production and nuclear staining. When BNL CL.<span class="hlt">2</span> cells were treated with <span class="hlt">G</span>/GO (50 mU/ml), viability of the cells was 54.1%. However, the number of living cells after the addition of UDN glycoprotein in the presence of <span class="hlt">G</span>/GO increased. UDN glycoprotein protected from cell damage caused by <span class="hlt">G</span>/GO. Interestingly, UDN glycoprotein decreased NF-kappaB activation and stimulated NO production in <span class="hlt">G/GO-induced</span> BNL CL.<span class="hlt">2</span> cells. In apoptotic parameters, UDN glycoprotein inhibited activations of caspase-3 and PARP cleavage in <span class="hlt">G/GO-induced</span> BNL CL.<span class="hlt">2</span> cells. The results of nuclear staining indicated that UDN glycoprotein (50 microg/ml) has a protective ability from apoptotic cell death caused <span class="hlt">G</span>/GO (50 mU/ml). In conclusion, UDN glycoprotein has a protective effect on apoptosis <span class="hlt">induced</span> by <span class="hlt">G</span>/GO through the inhibition of NF-kappaB, caspase-3, and PARP activity, and the stimulation of NO production in BNL CL.<span class="hlt">2</span> cells.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70118552','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70118552"><span>SIRT1 attenuates palmitate-<span class="hlt">induced</span> endoplasmic reticulum stress and insulin resistance in Hep<span class="hlt">G</span><span class="hlt">2</span> cells via induction of oxygen-regulated protein 150</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Jung, T.W.; Lee, K.T.; Lee, M.W.; Ka, K.H.</p> <p>2012-01-01</p> <p>Endoplasmic reticulum (ER) stress has been implicated in the pathology of type <span class="hlt">2</span> diabetes mellitus (T<span class="hlt">2</span>DM). Although SIRT1 has a therapeutic effect on T<span class="hlt">2</span>DM, the mechanisms by which SIRT1 ameliorates insulin resistance (IR) remain unclear. In this study, we investigated the impact of SIRT1 on palmitate-<span class="hlt">induced</span> ER stress in Hep<span class="hlt">G</span><span class="hlt">2</span> cells and its underlying signal pathway. Treatment with resveratrol, a SIRT1 activator significantly inhibited palmitate-<span class="hlt">induced</span> ER stress, leading to the protection against palmitate-<span class="hlt">induced</span> ER stress and insulin resistance. Resveratrol and SIRT1 overexpression <span class="hlt">induced</span> the expression of oxygen-regulated protein (ORP) 150 in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Forkhead box O1 (FOXO1) was involved in the regulation of ORP150 expression because suppression of FOXO1 inhibited the induction of ORP150 by SIRT1. Our results indicate a novel mechanism by which SIRT1 regulates ER stress by overexpression of ORP150, and suggest that SIRT1 ameliorates palmitate-<span class="hlt">induced</span> insulin resistance in Hep<span class="hlt">G</span><span class="hlt">2</span> cells via regulation of ER stress.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28206769','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28206769"><span>Extensive Chemical Modifications in the Primary Protein Structure of Ig<span class="hlt">G</span>1 Subvisible Particles Are Necessary for <span class="hlt">Breaking</span> Immune Tolerance.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Boll, Björn; Bessa, Juliana; Folzer, Emilien; Ríos Quiroz, Anacelia; Schmidt, Roland; Bulau, Patrick; Finkler, Christof; Mahler, Hanns-Christian; Huwyler, Jörg; Iglesias, Antonio; Koulov, Atanas V</p> <p>2017-04-03</p> <p>A current concern with the use of therapeutic proteins is the likely presence of aggregates and submicrometer, subvisible, and visible particles. It has been proposed that aggregates and particles may lead to unwanted increases in the immune response with a possible impact on safety or efficacy. The aim of this study was thus to evaluate the ability of subvisible particles of a therapeutic antibody to <span class="hlt">break</span> immune tolerance in an Ig<span class="hlt">G</span>1 transgenic mouse model and to understand the particle attributes that might play a role in this process. We investigated the immunogenic properties of subvisible particles (unfractionated, mixed populations, and well-defined particle size fractions) using a transgenic mouse model expressing a mini-repertoire of human Ig<span class="hlt">G</span>1 (hIg<span class="hlt">G</span>1 tg). Immunization with proteinaceous subvisible particles generated by artificial stress conditions demonstrated that only subvisible particles bearing very extensive chemical modifications within the primary amino acid structure could <span class="hlt">break</span> immune tolerance in the hIg<span class="hlt">G</span>1 transgenic mouse model. Protein particles exhibiting low levels of chemical modification were not immunogenic in this model.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3117017','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3117017"><span>Quantitative, non-invasive imaging of radiation-<span class="hlt">induced</span> DNA double strand <span class="hlt">breaks</span> in vivo</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Li, Wenrong; Li, Fang; Huang, Qian; Shen, Jingping; Wolf, Frank; He, Yujun; Liu, Xinjian; Hu, Y. Angela; Bedford, Joel. S.; Li, Chuan-Yuan</p> <p>2011-01-01</p> <p>DNA double strand <span class="hlt">breaks</span> is a major form of DNA damage and a key mechanism through which radiotherapy and some chemotherapeutic agents kill cancer cells. Despite its importance, measuring DNA double strand <span class="hlt">breaks</span> is still a tedious task that is normally carried out by gel electrophoresis or immunofluorescence staining. Here we report a novel approach to image and quantify DNA double strand <span class="hlt">breaks</span> in live mammalian cells through bi-fragment luciferase reconstitution. N- and C- terminal fragments of firefly luciferase gene were fused with H<span class="hlt">2</span>AX and MDC1 genes, respectively. Our strategy was based on the established fact that at the sites of DNA double strand <span class="hlt">breaks</span>, H<span class="hlt">2</span>AX protein is phosphoryated and physically associates with the MDC1 protein, thus bringing together N- and C- luciferase fragments and reconstituting luciferase activity. Our strategy allowed serial, non-invasive quantification of DNA double strand <span class="hlt">breaks</span> in cells irradiated with x-rays and 56Fe ions. Furthermore, it allowed for the evaluation of DNA double strand <span class="hlt">breaks</span> (DSBs) non-invasively in vivo in irradiated tumors over two weeks. Surprisingly, we detected a second wave of DSB induction in irradiated tumor cells days after radiation exposure in addition to the initial rapid induction of DSBs. We conclude that our new split-luciferase based method for imaging γ-H<span class="hlt">2</span>AX-MDC1 interaction is a powerful new tool to study DNA double strand <span class="hlt">break</span> repair kinetics in vivo with considerable advantage for experiments requiring observations over an extended period of time. PMID:21527553</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25287622','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25287622"><span>Double-strand <span class="hlt">break</span> repair-adox: Restoration of suppressed double-strand <span class="hlt">break</span> repair during mitosis <span class="hlt">induces</span> genomic instability.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Terasawa, Masahiro; Shinohara, Akira; Shinohara, Miki</p> <p>2014-12-01</p> <p>Double-strand <span class="hlt">breaks</span> (DSBs) are one of the severest types of DNA damage. Unrepaired DSBs easily <span class="hlt">induce</span> cell death and chromosome aberrations. To maintain genomic stability, cells have checkpoint and DSB repair systems to respond to DNA damage throughout most of the cell cycle. The failure of this process often results in apoptosis or genomic instability, such as aneuploidy, deletion, or translocation. Therefore, DSB repair is essential for maintenance of genomic stability. During mitosis, however, cells seem to suppress the DNA damage response and proceed to the next <span class="hlt">G</span>1 phase, even if there are unrepaired DSBs. The biological significance of this suppression is not known. In this review, we summarize recent studies of mitotic DSB repair and discuss the mechanisms of suppression of DSB repair during mitosis. DSB repair, which maintains genomic integrity in other phases of the cell cycle, is rather toxic to cells during mitosis, often resulting in chromosome missegregation and aberration. Cells have multiple safeguards to prevent genomic instability during mitosis: inhibition of 53BP1 or BRCA1 localization to DSB sites, which is important to promote non-homologous end joining or homologous recombination, respectively, and also modulation of the non-homologous end joining core complex to inhibit DSB repair. We discuss how DSBs during mitosis are toxic and the multiple safeguard systems that suppress genomic instability. © 2014 The Authors. Cancer Science published by Wiley Publishing Asia Pty Ltd on behalf of Japanese Cancer Association.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25075717','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25075717"><span>Nicotine, cotinine, and β-nicotyrine inhibit NNK-<span class="hlt">induced</span> DNA-strand <span class="hlt">break</span> in the hepatic cell line HepaRG.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ordonez, Patricia; Sierra, Ana Belen; Camacho, Oscar M; Baxter, Andrew; Banerjee, Anisha; Waters, David; Minet, Emmanuel</p> <p>2014-07-15</p> <p>Recent in vitro work using purified enzymes demonstrated that nicotine and/or a nicotine metabolite could inhibit CYPs (CYP<span class="hlt">2</span>A6, <span class="hlt">2</span>A13, <span class="hlt">2</span>E1) involved in the metabolism of the genotoxic tobacco nitrosamine NNK. This observation raises the possibility of nicotine interaction with the mechanism of NNK bioactivation. Therefore, we hypothesized that nicotine or a nicotine metabolite such as cotinine might contribute to the inhibition of NNK-<span class="hlt">induced</span> DNA strand <span class="hlt">breaks</span> by interfering with CYP enzymes. The effect of nicotine and cotinine on DNA strand <span class="hlt">breaks</span> was evaluated using the COMET assay in CYP competent HepaRG cells incubated with bioactive CYP-dependent NNK and CYP-independent NNKOAc (4-(acetoxymethylnitrosoamino)-1-(3-pyridyl)-1-butanone). We report a dose-dependent reduction in DNA damage in hepatic-derived cell lines in the presence of nicotine and cotinine. Those results are discussed in the context of the in vitro model selected. Copyright © 2014. Published by Elsevier Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22462199-inhibition-sh2-domain-containing-inositol-phosphatase-ship2-ameliorates-palmitate-induced-apoptosis-through-regulating-akt-foxo1-pathway-ros-production-hepg2-cells','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22462199-inhibition-sh2-domain-containing-inositol-phosphatase-ship2-ameliorates-palmitate-induced-apoptosis-through-regulating-akt-foxo1-pathway-ros-production-hepg2-cells"><span>Inhibition of SH<span class="hlt">2</span>-domain-containing inositol 5-phosphatase (SHIP<span class="hlt">2</span>) ameliorates palmitate <span class="hlt">induced</span>-apoptosis through regulating Akt/FOXO1 pathway and ROS production in Hep<span class="hlt">G</span><span class="hlt">2</span> cells</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Gorgani-Firuzjaee, Sattar; Adeli, Khosrow; Meshkani, Reza, E-mail: rmeshkani@tums.ac.ir</p> <p></p> <p>The serine–threonine kinase Akt regulates proliferation and survival by phosphorylating a network of protein substrates; however, the role of a negative regulator of the Akt pathway, the SH<span class="hlt">2</span>-domain-containing inositol 5-phosphatase (SHIP<span class="hlt">2</span>) in apoptosis of the hepatocytes, remains unknown. In the present study, we studied the molecular mechanisms linking SHIP<span class="hlt">2</span> expression to apoptosis using overexpression or suppression of SHIP<span class="hlt">2</span> gene in Hep<span class="hlt">G</span><span class="hlt">2</span> cells exposed to palmitate (0.5 mM). Overexpression of the dominant negative mutant SHIP<span class="hlt">2</span> (SHIP<span class="hlt">2</span>-DN) significantly reduced palmitate-<span class="hlt">induced</span> apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells, as these cells had increased cell viability, decreased apoptotic cell death and reduced the activity of caspase-3, cytochrome cmore » and poly (ADP-ribose) polymerase. Overexpression of the wild-type SHIP<span class="hlt">2</span> gene led to a massive apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. The protection from palmitate-<span class="hlt">induced</span> apoptosis by SHIP<span class="hlt">2</span> inhibition was accompanied by a decrease in the generation of reactive oxygen species (ROS). In addition, SHIP<span class="hlt">2</span> inhibition was accompanied by an increased Akt and FOXO-1 phosphorylation, whereas overexpression of the wild-type SHIP<span class="hlt">2</span> gene had the opposite effects. Taken together, these findings suggest that SHIP<span class="hlt">2</span> expression level is an important determinant of hepatic lipoapotosis and its inhibition can potentially be a target in treatment of hepatic lipoapoptosis in diabetic patients. - Highlights: • Lipoapoptosis is the major contributor to the development of NAFLD. • The PI3-K/Akt pathway regulates apoptosis in different cells. • The role of negative regulator of this pathway, SHIP<span class="hlt">2</span> in lipoapoptosis is unknown. • SHIP<span class="hlt">2</span> inhibition significantly reduces palmitate-<span class="hlt">induced</span> apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. • SHIP<span class="hlt">2</span> inhibition prevents palmitate <span class="hlt">induced</span>-apoptosis by regulating Akt/FOXO1 pathway.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27110813','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27110813"><span>Vaccination with the Secreted Glycoprotein <span class="hlt">G</span> of Herpes Simplex Virus <span class="hlt">2</span> <span class="hlt">Induces</span> Protective Immunity after Genital Infection.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Önnheim, Karin; Ekblad, Maria; Görander, Staffan; Bergström, Tomas; Liljeqvist, Jan-Åke</p> <p>2016-04-22</p> <p>Herpes simplex virus <span class="hlt">2</span> (HSV-<span class="hlt">2</span>) infects the genital mucosa and establishes a life-long infection in sensory ganglia. After primary infection HSV-<span class="hlt">2</span> may reactivate causing recurrent genital ulcerations. HSV-<span class="hlt">2</span> infection is prevalent, and globally more than 400 million individuals are infected. As clinical trials have failed to show protection against HSV-<span class="hlt">2</span> infection, new vaccine candidates are warranted. The secreted glycoprotein <span class="hlt">G</span> (sg<span class="hlt">G</span>-<span class="hlt">2</span>) of HSV-<span class="hlt">2</span> was evaluated as a prophylactic vaccine in mice using two different immunization and adjuvant protocols. The protocol with three intramuscular immunizations combining sg<span class="hlt">G</span>-<span class="hlt">2</span> with cytosine-phosphate-guanine dinucleotide (Cp<span class="hlt">G</span>) motifs and alum <span class="hlt">induced</span> almost complete protection from genital and systemic disease after intra-vaginal challenge with HSV-<span class="hlt">2</span>. Robust immunoglobulin <span class="hlt">G</span> (Ig<span class="hlt">G</span>) antibody titers were detected with no neutralization activity. Purified splenic CD4+ T cells proliferated and produced interferon-γ (IFN-γ) when re-stimulated with the antigen in vitro. sg<span class="hlt">G</span>-<span class="hlt">2</span> + adjuvant intra-muscularly immunized mice showed a significant reduction of infectious HSV-<span class="hlt">2</span> and increased IFN-γ levels in vaginal washes. The HSV-<span class="hlt">2</span> DNA copy numbers were significantly reduced in dorsal root ganglia, spinal cord, and in serum at day six or day 21 post challenge. We show that a sg<span class="hlt">G</span>-<span class="hlt">2</span> based vaccine is highly effective and can be considered as a novel candidate in the development of a prophylactic vaccine against HSV-<span class="hlt">2</span> infection.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20170009422&hterms=drug&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Ddrug','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20170009422&hterms=drug&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Ddrug"><span>Adrenergic Receptor Stimulation Prevents Radiation-<span class="hlt">Induced</span> DNA Strand <span class="hlt">Breaks</span>, Apoptosis and Gene Expression in Simulated Microgravity</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Moreno-Villanueva, Maria; Krieger, Stephanie; Feiveson, Alan; Kovach, Annie Marie; Buerkle, Alexander; Wu, Honglu</p> <p>2017-01-01</p> <p>Under Earth gravity conditions cellular damage can be counteracted by activation of the physiological defense mechanisms or through medical interventions. The mode of action of both, physiological response and medical interventions can be affected by microgravity leading to failure in repairing the damage. There are many studies reporting the effects of microgravity and/or radiation on cellular functions. However, little is known about the synergistic effects on cellular response to radiation when other endogenous cellular stress-response pathways are previously activated. Here, we investigated whether previous stimulation of the adrenergic receptor, which modulates immune response, affects radiation-<span class="hlt">induced</span> apoptosis in immune cells under simulated microgravity conditions. Peripheral blood mononuclear cells (PBMCs) were stimulated with isoproterenol (a sympathomimetic drug) and exposed to 0.8 or <span class="hlt">2</span>Gy gamma-radiation in simulated microgravity versus Earth gravity. Expression of genes involved in adrenergic receptor pathways, DNA repair and apoptosis as well as the number of apoptotic cells and DNA strand <span class="hlt">breaks</span> were determined. Our results showed that, under simulated microgravity conditions, previous treatment with isoproterenol prevented radiation-<span class="hlt">induced</span> i) gene down regulation, ii) DNA strand <span class="hlt">breaks</span> formation and iii) apoptosis induction. Interestedly, we found a radiation-<span class="hlt">induced</span> increase of adrenergic receptor gene expression, which was also abolished in simulated microgravity. Understanding the mechanisms of isoproterenol-mediated radioprotection in simulated microgravity can help to develop countermeasures for space-associated health risks as well as radio-sensitizers for cancer therapy.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005PhDT........52C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005PhDT........52C"><span>Correlations between radiation-<span class="hlt">induced</span> double strand <span class="hlt">breaks</span>, cell division delay, and cyclin-dependent signaling in x-irradiated NIH3T3 fibroblasts</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cariveau, Mickael J.</p> <p>2005-07-01</p> <p>Molecular responses to radiation-<span class="hlt">induced</span> DNA double strand <span class="hlt">breaks</span> (DSB) are mediated by the phosphorylation of the histone variant H<span class="hlt">2</span>AX which forms identifiable gamma-H<span class="hlt">2</span>AX foci at the site of the DSB. This event is thought to be linked with the down-regulation of signaling proteins contributing to the checkpoints regulating cell cycle progression and, vis-a-vis , the induction of cell division delay. However, it is unclear whether this division delay is directly related to the number of DSB (gamma-H<span class="hlt">2</span>AX foci) sustained by an irradiated cell and, if so, whether this number drives cells into cell cycle delay or apoptosis. For this reason, studies were conducted in the immortalized NIH/3T3 fibroblast cell in order to establish correlations between the temporal appearance of the gamma-H<span class="hlt">2</span>AX foci (a DSB) and the expression of the cell cycle regulatory proteins, cyclin E, A, B1, and their cyclin kinase inhibitor, p21. Cell cycle kinetics and flow cytometry were used to establish radiation-<span class="hlt">induced</span> division delay over a dose range of 1--6 Gy where a mitotic delay of <span class="hlt">2</span>.65 min/cGy was established. Correlations between the expression of cyclin E, A, B1, p21, and the generation of DSB were established in NIH/3T3 cells exposed to <span class="hlt">2</span> or 4 Gy x-irradiation. The data suggest that the <span class="hlt">G</span>1/S and S phase delay (cyclin E and cyclin A protein levels) are dependent on the dose of radiation while the <span class="hlt">G</span><span class="hlt">2</span>/M (cyclin B1 protein levels) delay is dependent on the quantity of DSB sustained by the irradiated cell.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22215977-isoorientin-induces-apoptosis-through-mitochondrial-dysfunction-inhibition-pi3k-akt-signaling-pathway-hepg2-cancer-cells','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22215977-isoorientin-induces-apoptosis-through-mitochondrial-dysfunction-inhibition-pi3k-akt-signaling-pathway-hepg2-cancer-cells"><span>Isoorientin <span class="hlt">induces</span> apoptosis through mitochondrial dysfunction and inhibition of PI3K/Akt signaling pathway in Hep<span class="hlt">G</span><span class="hlt">2</span> cancer cells</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Yuan, Li; Wang, Jing; Xiao, Haifang</p> <p></p> <p>Isoorientin (ISO) is a flavonoid compound that can be extracted from several plant species, such as Phyllostachys pubescens, Patrinia, and Drosophyllum lusitanicum; however, its biological activity remains poorly understood. The present study investigated the effects and putative mechanism of apoptosis <span class="hlt">induced</span> by ISO in human hepatoblastoma cancer (Hep<span class="hlt">G</span><span class="hlt">2</span>) cells. The results showed that ISO <span class="hlt">induced</span> cell death in a dose-dependent manner in Hep<span class="hlt">G</span><span class="hlt">2</span> cells, but no toxicity in human liver cells (HL-7702) and buffalo rat liver cells (BRL-3A) treated with ISO at the indicated concentrations. ISO-<span class="hlt">induced</span> cell death included apoptosis which characterized by the appearance of nuclear shrinkage, the cleavagemore » of poly (ADP-ribose) polymerase (PARP) and DNA fragmentation. ISO significantly (p < 0.01) increased the Bax/Bcl-<span class="hlt">2</span> ratio, disrupted the mitochondrial membrane potential (MMP), increased the release of cytochrome c, activated caspase-3, and enhanced intracellular levels of reactive oxygen species (ROS) and nitric oxide (NO). In addition, ISO effectively inhibited the phosphorylation of Akt and increased FoxO4 expression. The PI3K/Akt inhibitor LY294002 enhanced the apoptosis-<span class="hlt">inducing</span> effect of ISO. However, LY294002 markedly quenched ROS and NO generation and diminished the protein expression of heme peroxidase enzyme (HO-1) and <span class="hlt">inducible</span> nitric oxide synthase (iNOS). Furthermore, the addition of a ROS inhibitor (N-acetyl cysteine, NAC) or iNOS inhibitor (N-[3-(aminomethyl) benzyl] acetamidine, dihydrochloride, 1400W) significantly diminished the apoptosis <span class="hlt">induced</span> by ISO and also blocked the phosphorylation of Akt. These results demonstrated for the first time that ISO <span class="hlt">induces</span> apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells and indicate that this apoptosis might be mediated through mitochondrial dysfunction and PI3K/Akt signaling pathway, and has no toxicity in normal liver cells, suggesting that ISO may have good potential as a therapeutic and chemopreventive agent for liver cancer</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5278675','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5278675"><span>Epicutaneous immunization with ovalbumin and Cp<span class="hlt">G</span> <span class="hlt">induces</span> TH1/TH17 cytokines, which regulate IgE and Ig<span class="hlt">G</span><span class="hlt">2</span>a production</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Majewska-Szczepanik, Monika; Askenase, Philip W.; Lobo, Francis M.; Marcińska, Katarzyna; Wen, Li; Szczepanik, Marian</p> <p>2017-01-01</p> <p>Background Subcutaneous allergen-specific immunotherapy is a standard route for the immunotherapy of allergic diseases. It modulates the course of allergy and can generate long-term remission. However, subcutaneous allergen-specific immunotherapy can also <span class="hlt">induce</span> anaphylaxis in some patients, and therefore additional routes of administration should be investigated to improve the safety and tolerability of immunotherapy. Objective We sought to determine whether epicutaneous treatment with antigen in the presence of a Toll-like receptor 9 agonist can suppress TH<span class="hlt">2</span>-mediated responses in an antigen-specific manner. Methods Epicutaneous immunization was performed by applying a skin patch soaked with ovalbumin (OVA) plus Cp<span class="hlt">G</span>, and its suppressor activity was determined by using the mouse model of atopic dermatitis. Finally, adoptive cell transfers were implemented to characterize the regulatory cells that are <span class="hlt">induced</span> by epicutaneous immunization. Results Epicutaneous immunization with OVA and Cp<span class="hlt">G</span> reduces the production of OVA-specific IgE and increases the synthesis of OVA-specific Ig<span class="hlt">G</span><span class="hlt">2</span>a antibodies in an antigen-specific manner. Moreover, eosinophil peroxidase activity in the skin and production of IL-4, IL-5, IL-10, and IL-13 are suppressed. The observed reduction of IgE synthesis is transferable with T-cell receptor (TCR) αβ+CD4+CD25− cells, whereas Ig<span class="hlt">G</span><span class="hlt">2</span>a production is dependent on both TCRαβ+ and TCRγδ+ T cells. Further experiments show that the described phenomenon is myeloid differentiation primary response 88, IFN-γ, and IL-17A dependent. Finally, the results suggest that epicutaneous immunization with OVA and Cp<span class="hlt">G</span> decreases the synthesis of OVA-specific IgE and skin eosinophil peroxidase activity in mice with ongoing skin allergy. Conclusion Epicutaneous application of protein antigen in the presence of adjuvant could be an attractive needle-free and self-administered immunotherapy for allergic diseases. PMID:26810716</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27391338','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27391338"><span>Low level phosphorylation of histone H<span class="hlt">2</span>AX on serine 139 (γH<span class="hlt">2</span>AX) is not associated with DNA double-strand <span class="hlt">breaks</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Rybak, Paulina; Hoang, Agnieszka; Bujnowicz, Lukasz; Bernas, Tytus; Berniak, Krzysztof; Zarębski, Mirosław; Darzynkiewicz, Zbigniew; Dobrucki, Jerzy</p> <p>2016-08-02</p> <p>Phosphorylation of histone H<span class="hlt">2</span>AX on serine 139 (γH<span class="hlt">2</span>AX) is an early step in cellular response to a DNA double-strand <span class="hlt">break</span> (DSB). γH<span class="hlt">2</span>AX foci are generally regarded as markers of DSBs. A growing body of evidence demonstrates, however, that while induction of DSBs always brings about phosphorylation of histone H<span class="hlt">2</span>AX, the reverse is not true - the presence of γH<span class="hlt">2</span>AX foci should not be considered an unequivocal marker of DNA double-strand <span class="hlt">breaks</span>. We studied DNA damage <span class="hlt">induced</span> in A549 human lung adenocarcinoma cells by topoisomerase type I and II inhibitors (0.<span class="hlt">2</span> μM camptothecin, 10 μM etoposide or 0.<span class="hlt">2</span> μM mitoxantrone for 1 h), and using 3D high resolution quantitative confocal microscopy, assessed the number, size and the integrated intensity of immunofluorescence signals of individual γH<span class="hlt">2</span>AX foci <span class="hlt">induced</span> by these drugs. Also, investigated was spatial association between γH<span class="hlt">2</span>AX foci and foci of 53BP1, the protein involved in DSB repair, both in relation to DNA replication sites (factories) as revealed by labeling nascent DNA with EdU. Extensive 3D and correlation data analysis demonstrated that γH<span class="hlt">2</span>AX foci exhibit a wide range of sizes and levels of H<span class="hlt">2</span>AX phosphorylation, and correlate differently with 53BP1 and DNA replication. This is the first report showing lack of a link between low level phosphorylation γH<span class="hlt">2</span>AX sites and double-strand DNA <span class="hlt">breaks</span> in cells exposed to topoisomerase I or II inhibitors. The data are discussed in terms of mechanisms that may be involved in formation of γH<span class="hlt">2</span>AX sites of different sizes and intensities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27043378','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27043378"><span>Cytotoxic and genotoxic potential of Cr(VI), Cr(III)-nitrate and Cr(III)-EDTA complex in human hepatoma (Hep<span class="hlt">G</span><span class="hlt">2</span>) cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Novotnik, Breda; Ščančar, Janez; Milačič, Radmila; Filipič, Metka; Žegura, Bojana</p> <p>2016-07-01</p> <p>Chromium (Cr) and ethylenediaminetetraacetate (EDTA) are common environmental pollutants and can be present in high concentrations in surface waters at the same time. Therefore, chelation of Cr with EDTA can occur and thereby stable Cr(III)-EDTA complex is formed. Since there are no literature data on Cr(III)-EDTA toxicity, the aim of our work was to evaluate and compare Cr(III)-EDTA cytotoxic and genotoxic activity with those of Cr(VI) and Cr(III)-nitrate in human hepatoma (Hep<span class="hlt">G</span><span class="hlt">2</span>) cell line. First the effect of Cr(VI), Cr(III)-nitrate and Cr(III)-EDTA on cell viability was studied in the concentration range from 0.04 μ<span class="hlt">g</span> mL(-1) to 25 μ<span class="hlt">g</span> mL(-1) after 24 h exposure. Further the influence of non-cytotoxic concentrations of Cr(VI), Cr(III)-nitrate and Cr(III)-EDTA on DNA damage and genomic stability was determined with the comet assay and cytokinesis block micronucleus cytome assay, respectively. Cell viability was decreased only by Cr(VI) at concentrations above 1.0 μ<span class="hlt">g</span> mL(-1). Cr(VI) at ≥0.<span class="hlt">2</span> μ<span class="hlt">g</span> mL(-1) and Cr(III) at ≥1.0 μ<span class="hlt">g</span> mL(-1) <span class="hlt">induced</span> DNA damage, while after Cr(III)-EDTA exposure no formation DNA strand <span class="hlt">breaks</span> was determined. Statistically significant formation of micronuclei was <span class="hlt">induced</span> only by Cr(VI) at ≥0.<span class="hlt">2</span> μ<span class="hlt">g</span> mL(-1), while no influence on the frequency of nuclear buds nor nucleoplasmic bridges was observed at any exposure. This study provides the first evidence that Cr(III)-EDTA did not <span class="hlt">induce</span> DNA damage and had no influence on the genomic stability of Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Copyright © 2016 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11748978','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11748978"><span>Sister <span class="hlt">chromatid</span> exchange rate and alkaline comet assay scores in patients with ovarian cancer.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Baltaci, Volkan; Kayikçioğlu, Fulya; Alpas, Idil; Zeyneloğlu, Hulusi; Haberal, Ali</p> <p>2002-01-01</p> <p>Sister <span class="hlt">chromatid</span> exchange (SCE) frequencies were studied in patients with different types of ovarian malignancies and in healthy volunteers. The level of DNA damage in patients with ovarian malignancy and control subjects has also been studied by alkaline single cell gel electrophoresis (SCGE), also known as the comet assay. Peripheral blood was collected from 30 patients after histological confirmation of malignancy and 20 healthy female volunteers. The cells were evaluated according to their grade of damage. We found that the sister <span class="hlt">chromatid</span> exchange frequencies of cancer cases were significantly greater than that of controls (P < 0.001). The frequency of exchange in chromosomal groups A, B, and C, which include chromosomes 1-12, was higher than that of the other chromosomal groups in both groups. Comparison of the results of the alkaline comet assay in patient and control subjects showed a significant difference in the number of damaged cells. The frequency of limited migrated and extensive migrated cells in the women with ovarian malignancies was higher than that of control women (P < 0.001). SCE and SCGE can be used successfully to monitor DNA damage in women with ovarian cancer.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011PhRvD..84d4050M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011PhRvD..84d4050M"><span>D-foam-<span class="hlt">induced</span> flavor condensates and <span class="hlt">breaking</span> of supersymmetry in free Wess-Zumino fluids</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mavromatos, Nick E.; Sarkar, Sarben; Tarantino, Walter</p> <p>2011-08-01</p> <p>Recently [N. E. Mavromatos and S. Sarkar, New J. Phys. 10, 073009 (2008) NJOPFM1367-263010.1088/1367-2630/10/7/073009; N. E. Mavromatos, S. Sarkar, and W. Tarantino, Phys. Rev. DPRVDAQ1550-7998 80, 084046 (2009)10.1103/PhysRevD.80.084046], we argued that a particular model of string-inspired quantum space-time foam (D-foam) may <span class="hlt">induce</span> oscillations and mixing among flavored particles. As a result, rather than the mass-eigenstate vacuum, the correct ground state to describe the underlying dynamics is the flavor vacuum, proposed some time ago by Blasone and Vitiello as a description of quantum field theories with mixing. At the microscopic level, the <span class="hlt">breaking</span> of target-space supersymmetry is <span class="hlt">induced</span> in our space-time foam model by the relative transverse motion of brane defects. Motivated by these results, we show that the flavor vacuum, introduced through an inequivalent representation of the canonical (anti-) commutation relations, provides a vehicle for the <span class="hlt">breaking</span> of supersymmetry at a low-energy effective field-theory level; on considering the flavor-vacuum expectation value of the energy-momentum tensor and comparing with the form of a perfect relativistic fluid, it is found that the bosonic sector contributes as dark energy while the fermion contribution is like dust. This indicates a strong and novel <span class="hlt">breaking</span> of supersymmetry, of a nonperturbative nature, which may characterize the low-energy field theory of certain quantum-gravity models.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20100015501&hterms=epithelium&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Depithelium','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20100015501&hterms=epithelium&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Depithelium"><span>Distributions of Low- and High-LET Radiation-<span class="hlt">Induced</span> <span class="hlt">Breaks</span> in Chromosomes are Associated with Inter- and Intrachromosome Exchanges</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hada, Megumi; Zhang, Ye; Feiveson, Alan; Cucinotta, Francis A.; Wu, Honglu</p> <p>2010-01-01</p> <p>To study the breakpoint along the length of the chromosome <span class="hlt">induced</span> by low- and high-LET radiations, we exposed human epithelial cells in vitro to Cs-137 rays at both low and high dose rates, secondary neutrons at a low dose rate, and 600 MeV/u Fe ions at a high dose rate. The location of the <span class="hlt">breaks</span> was identified using the multicolor banding in situ hybridization (mBAND) that paints Chromosome 3 in 23 different colored bands. The breakpoint distributions were found to be similar between rays of low and high dose rates and between the two high-LET radiation types. Detailed analysis of the chromosome <span class="hlt">break</span> ends involved in inter- and intrachromosome exchanges revealed that only the <span class="hlt">break</span> ends participating in interchromosome exchanges contributed to the hot spots found for low-LET. For <span class="hlt">break</span> ends participating in intrachromosome exchanges, the distributions for all four radiation scenarios were similar with clusters of <span class="hlt">breaks</span> found in three regions. Analysis of the locations of the two <span class="hlt">break</span> ends in Chromosome 3 that joined to form an intrachromosome exchange demonstrated that two <span class="hlt">breaks</span> with a greater genomic separation may be more likely to rejoin than two closer <span class="hlt">breaks</span>, indicating that chromatin folding can play an important role in the rejoining of chromosome <span class="hlt">breaks</span>. Our study demonstrated that the gene-rich regions do not necessarily contain more <span class="hlt">breaks</span>. The breakpoint distribution depends more on the likelihood that a <span class="hlt">break</span> will join with another <span class="hlt">break</span> in the same chromosome or in a different chromosome.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29730927','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29730927"><span>Epoxy Stearic Acid, an Oxidative Product Derived from Oleic Acid, <span class="hlt">Induces</span> Cytotoxicity, Oxidative Stress, and Apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> Cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liu, Ying; Cheng, Yajun; Li, Jinwei; Wang, Yuanpeng; Liu, Yuanfa</p> <p>2018-05-23</p> <p>In the present study, effects of cis-9,10-epoxy stearic acid (ESA) generated by the thermal oxidation of oleic acid on Hep<span class="hlt">G</span><span class="hlt">2</span> cells, including cytotoxicity, apoptosis, and oxidative stress, were investigated. Our results revealed that ESA decreased the cell viability and <span class="hlt">induced</span> cell death. Cell cycle analysis with propidium iodide staining showed that ESA <span class="hlt">induced</span> cell cycle arrest at the <span class="hlt">G</span>0/<span class="hlt">G</span>1 phase in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Cell apoptosis analysis with annexin V and propidium iodide staining demonstrated that ESA <span class="hlt">induced</span> Hep<span class="hlt">G</span><span class="hlt">2</span> cell apoptotic events in a dose- and time-dependent manner; the apoptosis of cells after treated with 500 μM ESA for 12, 24, and 48 h was 32.16, 38.70, and 65.80%, respectively. Furthermore, ESA treatment to Hep<span class="hlt">G</span><span class="hlt">2</span> cells resulted in an increase in reactive oxygen species and malondialdehyde (from 0.84 ± 0.02 to 8.90 ± 0.50 nmol/mg of protein) levels and a reduction in antioxidant enzyme activity, including superoxide dismutase (from 1.34 ± 0.27 to 0.10 ± 0.007 units/mg of protein), catalase (from 100.04 ± 5.05 to 20.09 ± 3.00 units/mg of protein), and glutathione peroxidase (from 120.44 ± 7.62 to 35.84 ± 5.99 milliunits/mg of protein). These findings provide critical information on the effects of ESA on Hep<span class="hlt">G</span><span class="hlt">2</span> cells, particularly cytotoxicity and oxidative stress, which is important for the evaluation of the biosafety of the oxidative product of oleic acid.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24370495','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24370495"><span>Ursodeoxycholic acid inhibits overexpression of P-glycoprotein <span class="hlt">induced</span> by doxorubicin in Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Komori, Yuki; Arisawa, Sakiko; Takai, Miho; Yokoyama, Kunihiro; Honda, Minako; Hayashi, Kazuhiko; Ishigami, Masatoshi; Katano, Yoshiaki; Goto, Hidemi; Ueyama, Jun; Ishikawa, Tetsuya; Wakusawa, Shinya</p> <p>2014-02-05</p> <p>The hepatoprotective action of ursodeoxycholic acid (UDCA) was previously suggested to be partially dependent on its antioxidative effect. Doxorubicin (DOX) and reactive oxygen species have also been implicated in the overexpression of P-glycoprotein (P-gp), which is encoded by the MDR1 gene and causes antitumor multidrug resistance. In the present study, we assessed the effects of UDCA on the expression of MDR1 mRNA, P-gp, and intracellular reactive oxygen species levels in DOX-treated Hep<span class="hlt">G</span><span class="hlt">2</span> cells and compared them to those of other bile acids. DOX-<span class="hlt">induced</span> increases in reactive oxygen species levels and the expression of MDR1 mRNA were inhibited by N-acetylcysteine, an antioxidant, and the DOX-<span class="hlt">induced</span> increase in reactive oxygen species levels and DOX-<span class="hlt">induced</span> overexpression of MDR1 mRNA and P-gp were inhibited by UDCA. Cells treated with UDCA showed improved rhodamine 123 uptake, which was decreased in cells treated with DOX alone. Moreover, cells exposed to DOX for 24h combined with UDCA accumulated more DOX than that of cells treated with DOX alone. Thus, UDCA may have inhibited the overexpression of P-gp by suppressing DOX-<span class="hlt">induced</span> reactive oxygen species production. Chenodeoxycholic acid (CDCA) also exhibited these effects, whereas deoxycholic acid and litocholic acid were ineffective. In conclusion, UDCA and CDCA had an inhibitory effect on the induction of P-gp expression and reactive oxygen species by DOX in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. The administration of UDCA may be beneficial due to its ability to prevent the overexpression of reactive oxygen species and acquisition of multidrug resistance in hepatocellular carcinoma cells. Copyright © 2013 Elsevier B.V. All rights reserved.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li class="active"><span>14</span></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_14 --> <div id="page_15" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li class="active"><span>15</span></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="281"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3078076','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3078076"><span>Mitotic centromeric targeting of HP1 and its binding to Sgo1 are dispensable for sister-<span class="hlt">chromatid</span> cohesion in human cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Kang, Jungseog; Chaudhary, Jaideep; Dong, Hui; Kim, Soonjoung; Brautigam, Chad A.; Yu, Hongtao</p> <p>2011-01-01</p> <p>Human Shugoshin 1 (Sgo1) protects centromeric sister-<span class="hlt">chromatid</span> cohesion during prophase and prevents premature sister-<span class="hlt">chromatid</span> separation. Heterochromatin protein 1 (HP1) has been proposed to protect centromeric sister-<span class="hlt">chromatid</span> cohesion by directly targeting Sgo1 to centromeres in mitosis. Here we show that HP1α is targeted to mitotic centromeres by INCENP, a subunit of the chromosome passenger complex (CPC). Biochemical and structural studies show that both HP1–INCENP and HP1–Sgo1 interactions require the binding of the HP1 chromo shadow domain to PXVXL/I motifs in INCENP or Sgo1, suggesting that the INCENP-bound, centromeric HP1α is incapable of recruiting Sgo1. Consistently, a Sgo1 mutant deficient in HP1 binding is functional in centromeric cohesion protection and localizes normally to centromeres in mitosis. By contrast, INCENP or Sgo1 mutants deficient in HP1 binding fail to localize to centromeres in interphase. Therefore, our results suggest that HP1 binding by INCENP or Sgo1 is dispensable for centromeric cohesion protection during mitosis of human cells, but might regulate yet uncharacterized interphase functions of CPC or Sgo1 at the centromeres. PMID:21346195</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22649484','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22649484"><span>Verbs in the lexicon: Why is hitting easier than <span class="hlt">breaking</span>?</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>McKoon, Gail; Love, Jessica</p> <p>2011-11-01</p> <p>Adult speakers use verbs in syntactically appropriate ways. For example, they know implicitly that the boy hit at the fence is acceptable but the boy broke at the fence is not. We suggest that this knowledge is lexically encoded in semantic decompositions. The decomposition for <span class="hlt">break</span> verbs (e.<span class="hlt">g</span>. crack, smash) is hypothesized to be more complex than that for hit verbs (e.<span class="hlt">g</span>. kick, kiss). Specifically, the decomposition of a <span class="hlt">break</span> verb denotes that "an entity changes state as the result of some external force" whereas the decomposition for a hit verb denotes only that "an entity potentially comes in contact with another entity." In this article, verbs of the two types were compared in a lexical decision experiment - Experiment 1 - and they were compared in sentence comprehension experiments with transitive sentences (e.<span class="hlt">g</span>. the car hit the bicycle and the car broke the bicycle) - Experiments <span class="hlt">2</span> and 3. In Experiment 1, processing times were shorter for the hit than the <span class="hlt">break</span> verbs and in Experiments <span class="hlt">2</span> and 3, processing times were shorter for the hit sentences than the <span class="hlt">break</span> sentences, results that are in accord with the complexities of the postulated semantic decompositions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3961563','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3961563"><span>The MutSβ complex is a modulator of p53-driven tumorigenesis through its functions in both DNA double strand <span class="hlt">break</span> repair and mismatch repair</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>van Oers, Johanna M. M.; Edwards, Yasmin; Chahwan, Richard; Zhang, Weijia; Smith, Cameron; Pechuan, Joaquín; Schaetzlein, Sonja; Jin, Bo; Wang, Yuxun; Bergman, Aviv; Scharff, Matthew D.; Edelmann, Winfried</p> <p>2014-01-01</p> <p>Loss of the DNA mismatch repair protein MSH3 leads to the development of a variety of tumors in mice without significantly affecting survival rates, suggesting a modulating role for the MutSβ (MSH<span class="hlt">2</span>-MSH3) complex in late onset tumorigenesis. To better study the role of MSH3 in tumor progression, we crossed Msh3−/− mice onto a tumor predisposing p53-deficient background. Survival of Msh3/p53 mice was not reduced compared to single p53 mutant mice; however, the tumor spectrum changed significantly from lymphoma to sarcoma, indicating MSH3 as a potent modulator of p53-driven tumorigenesis. Interestingly, Msh3−/− mouse embryonic fibroblasts displayed increased <span class="hlt">chromatid</span> <span class="hlt">breaks</span> and persistence of γH<span class="hlt">2</span>AX foci following ionizing radiation, indicating a defect in DNA double strand <span class="hlt">break</span> repair. Msh3/p53 tumors showed increased loss of heterozygosity, elevated genome-wide copy number variation, and a moderate microsatellite instability phenotype compared to Msh<span class="hlt">2</span>/p53 tumors, revealing that MSH<span class="hlt">2</span>-MSH3 suppresses tumorigenesis by maintaining chromosomal stability. Our results show that the MSH<span class="hlt">2</span>-MSH3 complex is important for the suppression of late onset tumors due to its role in DNA double strand <span class="hlt">break</span> repair as well as in DNA mismatch repair. Furthermore, they demonstrate that MSH<span class="hlt">2</span>-MSH3 suppresses chromosomal instability and modulates the tumor spectrum in p53-deficient tumorigenesis, and possibly plays a role in other chromosomally unstable tumors as well. PMID:24013230</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013EPJB...86..243A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013EPJB...86..243A"><span>Statistical evidence of strain <span class="hlt">induced</span> <span class="hlt">breaking</span> of metallic point contacts</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Alwan, Monzer; Candoni, Nadine; Dumas, Philippe; Klein, Hubert R.</p> <p>2013-06-01</p> <p>A scanning tunneling microscopy in <span class="hlt">break</span> junction regime and a mechanically controllable <span class="hlt">break</span> junction are used to acquire thousands of conductance-elongation curves by stretching until <span class="hlt">breaking</span> and re-connecting Au junctions. From a robust statistical analysis performed on large sets of experiments, parameters such as lifetime, elongation and occurrence probabilities are extracted. The analysis of results obtained for different stretching speeds of the electrodes indicates that the <span class="hlt">breaking</span> mechanism of di- and mono-atomic junction is identical, and that the junctions undergo atomic rearrangement during their stretching and at the moment of <span class="hlt">breaking</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21381055','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21381055"><span>Reduction in fluoride-<span class="hlt">induced</span> genotoxicity in mouse bone marrow cells after substituting high fluoride-containing water with safe drinking water.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Podder, Santosh; Chattopadhyay, Ansuman; Bhattacharya, Shelley</p> <p>2011-10-01</p> <p>Treatment of mice with 15 mg l(-1) sodium fluoride (NaF) for 30 days increased the number of cell death, chromosomal aberrations (CAs) and 'cells with <span class="hlt">chromatid</span> <span class="hlt">breaks</span>' (aberrant cells) compared with control. The present study was intended to determine whether the fluoride (F)-<span class="hlt">induced</span> genotoxicity could be reduced by substituting high F-containing water after 30 days with safe drinking water, containing 0.1 mg F ions l(-1). A significant fall in percentage of CAs and aberrant cells after withdrawal of F-treatment following 30 days of safe water treatment in mice was observed which was highest after 90 days, although their levels still remained significantly high compared with the control group. This observation suggests that F-<span class="hlt">induced</span> genotoxicity could be reduced by substituting high F-containing water with safe drinking water. Further study is warranted with different doses and extended treatment of safe water to determine whether the <span class="hlt">induced</span> damages could be completely reduced or not. Copyright © 2011 John Wiley & Sons, Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19002846','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19002846"><span>The effects of boric acid on sister <span class="hlt">chromatid</span> exchanges and chromosome aberrations in cultured human lymphocytes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Arslan, Mehmet; Topaktas, Mehmet; Rencuzogullari, Eyyüp</p> <p>2008-02-01</p> <p>The aim of this study was to determine the possible genotoxic effects of boric acid (BA) (E284), which is used as an antimicrobial agent in food, by using sister <span class="hlt">chromatid</span> exchange (SCEs) and chromosome aberration (CAs) tests in human peripheral lymphocytes. The human lymphocytes were treated with 400, 600, 800, and 1000 mug/mL concentrations of BA dissolved in dimethyl sulfoxide (DMSO), for 24 h and 48 h treatment periods. BA did not increase the SCEs for all the concentrations and treatment periods when compared to control and solvent control (DMSO). BA <span class="hlt">induced</span> structural and total CAs at all the tested concentrations for 24 and 48 h treatment periods. The induction of the total CAs was dose dependent for the 24 h treatment period. However, BA did not cause numerical CAs. BA showed a cytotoxic effect by decreasing the replication index (RI) and mitotic index (MI). BA decreased the MI in a dose-dependent manner for the 24 h treatment period.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27412346','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27412346"><span>Differential cytotoxicity <span class="hlt">induced</span> by the Titanium(IV)Salan complex Tc52 in <span class="hlt">G</span><span class="hlt">2</span>-phase independent of DNA damage.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pesch, Theresa; Schuhwerk, Harald; Wyrsch, Philippe; Immel, Timo; Dirks, Wilhelm; Bürkle, Alexander; Huhn, Thomas; Beneke, Sascha</p> <p>2016-07-13</p> <p>Chemotherapy is one of the major treatment modalities for cancer. Metal-based compounds such as derivatives of cisplatin are in the front line of therapy against a subset of cancers, but their use is restricted by severe side-effects and the induction of resistance in treated tumors. Subsequent research focused on development of cytotoxic metal-complexes without cross-resistance to cisplatin and reduced side-effects. This led to the discovery of first-generation titanium(IV)salan complexes, which reached clinical trials but lacked efficacy. New-generation titanium (IV)salan-complexes show promising anti-tumor activity in mice, but their molecular mechanism of cytotoxicity is completely unknown. Four different human cell lines were analyzed in their responses to a toxic (Tc52) and a structurally highly related but non-toxic (Tc53) titanium(IV)salan complex. Viability assays were used to reveal a suitable treatment range, flow-cytometry analysis was performed to monitor the impact of dosage and treatment time on cell-cycle distribution and cell death. Potential DNA strand <span class="hlt">break</span> induction and crosslinking was investigated by immunostaining of damage markers as well as automated fluorometric analysis of DNA unwinding. Changes in nuclear morphology were analyzed by DAPI staining. Acidic beta-galactosidase activity together with morphological changes was monitored to detect cellular senescence. Western blotting was used to analyze induction of pro-apoptotic markers such as activated caspase7 and cleavage of PARP1, and general stress kinase p38. Here we show that the titanium(IV)salan Tc52 is effective in <span class="hlt">inducing</span> cell death in the lower micromolar range. Surprisingly, Tc52 does not target DNA contrary to expectations deduced from the reported activity of other titanium complexes. Instead, Tc52 application interferes with progression from <span class="hlt">G</span><span class="hlt">2</span>-phase into mitosis and <span class="hlt">induces</span> apoptotic cell death in tested tumor cells. Contrarily, human fibroblasts undergo senescence in a</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/6872102','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/6872102"><span>Alkaline DNA fragmentation, DNA disentanglement evaluated viscosimetrically and sister <span class="hlt">chromatid</span> exchanges, after treatment in vivo with nitrofurantoin.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Parodi, S; Pala, M; Russo, P; Balbi, C; Abelmoschi, M L; Taningher, M; Zunino, A; Ottaggio, L; de Ferrari, M; Carbone, A; Santi, L</p> <p>1983-07-01</p> <p>Nitrofurantoin was not positive as a carcinogen in long term assays. In vitro it was positive in some short term tests and negative in others. We have examined Nitrofurantoin for its capability of <span class="hlt">inducing</span> DNA damage in vivo. With the alkaline elution technique, Nitrofurantoin appeared clearly positive in all the tissues examined (liver, kidney, lung, spleen and bone marrow). In the liver we also observed some cross-linking effect. In bone marrow cells Nitrofurantoin was also clearly positive in terms of sister <span class="hlt">chromatid</span> exchanges (SCEs) induction. DNA damage in vivo was also examined with a viscosimetric method, more sensitive than alkaline elution. With this method the results were essentially negative, suggesting that the two methods detect different types of damage. In view of its positivity in many organs and in two short term tests in vivo, the carcinogenic potential of Nitrofurantoin should be reconsidered.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/10380173','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/10380173"><span>Detection of ozone-<span class="hlt">induced</span> DNA single strand <span class="hlt">breaks</span> in murine bronchoalveolar lavage cells acutely exposed in vivo.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Haney, J T; Connor, T H; Li, L</p> <p>1999-04-01</p> <p>Single-strand <span class="hlt">breaks</span> (SSBs) in DNA have been used a biomarker of oxidative damage. The comet assay, also known as single-cell gel electrophoresis, was used to investigate the ability of ozone (O(3)) to <span class="hlt">induce</span> DNA SSBs in murine bronchoalveolar lavage (BAL) cells. The comet assay is more sensitive than other techniques currently utilized for detecting SSBs and requires fewer cells. In the present study, 3 mice were exposed for 3 h to 0.25 ppm of O(3), and 3 to 0.5 ppm of O(3) for 3 h. Two air-exposed mice served as negative controls. All mice were euthanized 3 h after exposure, at which time BAL cells were recovered from the lungs and stained with ethidium bromide. BAL cells recovered from an air-exposed mouse were exposed to various concentrations of H(<span class="hlt">2</span>)O(<span class="hlt">2</span>) in vitro for 1 h at 4 degrees C. Excluding cells from the H(<span class="hlt">2</span>)O(<span class="hlt">2</span>) group (n = 25), 50 randomly selected BAL cells were graded by comet tail length into 1 of 4 categories: no damage (0 mm), low damage (1-10 mm), medium damage (11-30 mm), and high damage (31 + mm). The nonparametric Wilcoxon rank-sum test was used for statistical analysis, and p values lower than .05 were considered significant. The H(<span class="hlt">2</span>)O(<span class="hlt">2</span>) and the 0.25 and 0.5 ppm O3 groups showed statistically significant increases in DNA SSBs as compared to air-exposed controls. The results of this study indicate that (1) O(3) <span class="hlt">induces</span> DNA strand <span class="hlt">breaks</span> in murine BAL cells at 0.25 and 0.5 ppm, as evidenced by statistically significant increases in the length of comet tails for O(3)-exposed groups, and (<span class="hlt">2</span>) the comet assay can be used to assess O(3)-<span class="hlt">induced</span> SSBs for in vivo exposures. Therefore, it has the potential as a biomarker for in vivo oxidant exposures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/14630540','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/14630540"><span>Chromatin- and temperature-dependent modulation of radiation-<span class="hlt">induced</span> double-strand <span class="hlt">breaks</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Elmroth, K; Nygren, J; Stenerlöw, B; Hultborn, R</p> <p>2003-10-01</p> <p>To investigate the influence of chromatin organization and scavenging capacity in relation to irradiation temperature on the induction of double-strand <span class="hlt">breaks</span> (DSB) in structures derived from human diploid fibroblasts. Agarose plugs with different chromatin structures (intact cells+/-wortmannin, permeabilized cells with condensed chromatin, nucleoids and DNA) were prepared and irradiated with X-rays at <span class="hlt">2</span> or 37 degrees C and lysed using two different lysis protocols (new ice-cold lysis or standard lysis at 37 degrees C). Induction of DSB was determined by constant-field gel electrophoresis. The dose-modifying factor (DMF(temp)) for irradiation at 37 compared with <span class="hlt">2</span> degrees C was 0.92 in intact cells (i.e. more DSB <span class="hlt">induced</span> at <span class="hlt">2</span> degrees C), but gradually increased to 1.5 in permeabilized cells, <span class="hlt">2.2</span> in nucleoids and <span class="hlt">2</span>.6 in naked DNA, suggesting a role of chromatin organization for temperature modulation of DNA damage. In addition, DMF(temp) was influenced by the presence of 0.1 M DMSO or 30 mM glutathione, but not by post-irradiation temperature. The protective effect of low temperature was correlated to the indirect effects of ionizing radiation and was not dependent on post-irradiation temperature. Reasons for a dose modifying factor <1 in intact cells are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4768956','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4768956"><span>Myricetin <span class="hlt">induces</span> apoptosis via endoplasmic reticulum stress and DNA double-strand <span class="hlt">breaks</span> in human ovarian cancer cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>XU, YE; XIE, QI; WU, SHAOHUA; YI, DAN; YU, YANG; LIU, SHIBING; LI, SONGYAN; LI, ZHIXIN</p> <p>2016-01-01</p> <p>The mechanisms underlying myricetin-<span class="hlt">induced</span> cancer cell apoptosis remain to be elucidated. Certain previous studies have shown that myricetin <span class="hlt">induces</span> apoptosis through the mitochondrial pathway. Apoptosis, however, can also be <span class="hlt">induced</span> by other classical pathways, including endoplasmic reticulum (ER) stress and DNA double-strand <span class="hlt">breaks</span> (DSBs). The aim of the present study was to assess whether these two apoptotic pathways are involved in myricetin-<span class="hlt">induced</span> cell death in SKOV3 ovarian cancer cells. The results revealed that treatment with myricetin inhibited viability of SKOV3 cells in a dose-dependent manner. Myricetin <span class="hlt">induced</span> nuclear chromatin condensation and fragmentation, and also upregulated the protein levels of active caspase 3 in a time-dependent manner. In addition, myricetin upregulated ER stress-associated proteins, glucose-regulated protein-78 and C/EBP homologous protein in SKOV3 cells. Phosphorylation of H<span class="hlt">2</span>AX, a marker of DNA DSBs, was revealed to be upregulated in myricetin-treated cells. The data indicated that myricetin <span class="hlt">induces</span> DNA DSBs and ER stress, which leads to apoptosis in SKOV3 cells. PMID:26782830</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3979685','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3979685"><span>Differential Radiosensitivity Phenotypes of DNA-PKcs Mutations Affecting NHEJ and HRR Systems following Irradiation with Gamma-Rays or Very Low Fluences of Alpha Particles</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Little, John B.; Kato, Takamitsu A.; Shih, Hung-Ying; Xie, Xian-Jin; Wilson Jr., Paul F.; Brogan, John R.; Kurimasa, Akihiro; Chen, David J.; Bedford, Joel S.; Chen, Benjamin P. C.</p> <p>2014-01-01</p> <p>We have examined cell-cycle dependence of chromosomal aberration induction and cell killing after high or low dose-rate γ irradiation in cells bearing DNA-PKcs mutations in the S2056 cluster, the T2609 cluster, or the kinase domain. We also compared sister <span class="hlt">chromatid</span> exchanges (SCE) production by very low fluences of α-particles in DNA-PKcs mutant cells, and in homologous recombination repair (HRR) mutant cells including Rad51C, Rad51D, and Fancg/xrcc9. Generally, chromosomal aberrations and cell killing by γ-rays were similarly affected by mutations in DNA-PKcs, and these mutant cells were more sensitive in <span class="hlt">G</span>1 than in S/<span class="hlt">G</span><span class="hlt">2</span> phase. In <span class="hlt">G</span>1-irradiated DNA-PKcs mutant cells, both chromosome- and <span class="hlt">chromatid</span>-type <span class="hlt">breaks</span> and exchanges were in excess than wild-type cells. For cells irradiated in late S/<span class="hlt">G</span><span class="hlt">2</span> phase, mutant cells showed very high yields of <span class="hlt">chromatid</span> <span class="hlt">breaks</span> compared to wild-type cells. Few exchanges were seen in DNA-PKcs-null, Ku80-null, or DNA-PKcs kinase dead mutants, but exchanges in excess were detected in the S2506 or T2609 cluster mutants. SCE induction by very low doses of α-particles is resulted from bystander effects in cells not traversed by α-particles. SCE seen in wild-type cells was completely abolished in Rad51C- or Rad51D-deficient cells, but near normal in Fancg/xrcc9 cells. In marked contrast, very high levels of SCEs were observed in DNA-PKcs-null, DNA-PKcs kinase-dead and Ku80-null mutants. SCE induction was also abolished in T2609 cluster mutant cells, but was only slightly reduced in the S2056 cluster mutant cells. Since both non-homologous end-joining (NHEJ) and HRR systems utilize initial DNA lesions as a substrate, these results suggest the possibility of a competitive interference phenomenon operating between NHEJ and at least the Rad51C/D components of HRR; the level of interaction between damaged DNA and a particular DNA-PK component may determine the level of interaction of such DNA with a relevant HRR component. PMID:24714417</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4732212','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4732212"><span>Antigenic Determinants of the Bilobal Cockroach Allergen Bla <span class="hlt">g</span> <span class="hlt">2</span>*</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Woodfolk, Judith A.; Glesner, Jill; Wright, Paul W.; Kepley, Christopher L.; Li, Mi; Himly, Martin; Muehling, Lyndsey M.; Gustchina, Alla; Wlodawer, Alexander; Chapman, Martin D.; Pomés, Anna</p> <p>2016-01-01</p> <p>Bla <span class="hlt">g</span> <span class="hlt">2</span> is a major indoor cockroach allergen associated with the development of asthma. Antigenic determinants on Bla <span class="hlt">g</span> <span class="hlt">2</span> were analyzed by mutagenesis based on the structure of the allergen alone and in complex with monoclonal antibodies that interfere with IgE antibody binding. The structural analysis revealed mechanisms of allergen-antibody recognition through cation-π interactions. Single and multiple Bla <span class="hlt">g</span> <span class="hlt">2</span> mutants were expressed in Pichia pastoris and purified. The triple mutant K132A/K251A/F162Y showed an ∼100-fold reduced capacity to bind IgE, while preserving the native molecular fold, as proven by x-ray crystallography. This mutant was still able to <span class="hlt">induce</span> mast cell release. T-cell responses were assessed by analyzing Th1/Th<span class="hlt">2</span> cytokine production and the CD4+ T-cell phenotype in peripheral blood mononuclear cell cultures. Although T-cell activating capacity was similar for the KKF mutant and Bla <span class="hlt">g</span> <span class="hlt">2</span> based on CD25 expression, the KKF mutant was a weaker <span class="hlt">inducer</span> of the Th<span class="hlt">2</span> cytokine IL-13. Furthermore, this mutant <span class="hlt">induced</span> IL-10 from a non-T-cell source at higher levels that those <span class="hlt">induced</span> by Bla <span class="hlt">g</span> <span class="hlt">2</span>. Our findings demonstrate that a rational design of site-directed mutagenesis was effective in producing a mutant with only 3 amino acid substitutions that maintained the same fold as wild type Bla <span class="hlt">g</span> <span class="hlt">2</span>. These residues, which were involved in IgE antibody binding, endowed Bla <span class="hlt">g</span> <span class="hlt">2</span> with a T-cell modulatory capacity. The antigenic analysis of Bla <span class="hlt">g</span> <span class="hlt">2</span> will be useful for the subsequent development of recombinant allergen vaccines. PMID:26644466</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3788364','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3788364"><span>Translational Symmetry <span class="hlt">Breaking</span> and Gapping of Heavy-Quasiparticle Pocket in URu<span class="hlt">2</span>Si<span class="hlt">2</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Yoshida, Rikiya; Tsubota, Koji; Ishiga, Toshihiko; Sunagawa, Masanori; Sonoyama, Jyunki; Aoki, Dai; Flouquet, Jacques; Wakita, Takanori; Muraoka, Yuji; Yokoya, Takayoshi</p> <p>2013-01-01</p> <p>URu<span class="hlt">2</span>Si<span class="hlt">2</span> is a uranium compound that exhibits a so-called ‘hidden-order’ transition at ~17.5 K. However, the order parameter of this second-order transition as well as many of its microscopic properties remain unclarified despite considerable research. One of the key questions in this regard concerns the type of spontaneous symmetry <span class="hlt">breaking</span> occurring at the transition; although rotational symmetry <span class="hlt">breaking</span> has been detected, it is not clear whether another type of symmetry <span class="hlt">breaking</span> also occurs. Another key question concerns the property of Fermi-surface gapping in the momentum space. Here we address these key questions by a momentum-dependent observation of electronic states at the transition employing ultrahigh-resolution three-dimensional angle-resolved photoemission spectroscopy. Our results provide compelling evidence of the spontaneous <span class="hlt">breaking</span> of the lattice's translational symmetry and particle-hole asymmetric gapping of a heavy quasiparticle pocket at the transition. PMID:24084937</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013NIMPB.311...27L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013NIMPB.311...27L"><span>Comparison of direct DNA strand <span class="hlt">breaks</span> <span class="hlt">induced</span> by low energy electrons with different inelastic cross sections</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Jun-Li; Li, Chun-Yan; Qiu, Rui; Yan, Cong-Chong; Xie, Wen-Zhang; Zeng, Zhi; Tung, Chuan-Jong</p> <p>2013-09-01</p> <p>In order to study the influence of inelastic cross sections on the simulation of direct DNA strand <span class="hlt">breaks</span> <span class="hlt">induced</span> by low energy electrons, six different sets of inelastic cross section data were calculated and loaded into the Geant4-DNA code to calculate the DNA strand <span class="hlt">break</span> yields under the same conditions. The six sets of the inelastic cross sections were calculated by applying the dielectric function method of Emfietzoglou's optical-data treatments, with two different optical datasets and three different dispersion models, using the same Born corrections. Results show that the inelastic cross sections have a notable influence on the direct DNA strand <span class="hlt">break</span> yields. The yields simulated with the inelastic cross sections based on Hayashi's optical data are greater than those based on Heller's optical data. The discrepancies are about 30-45% for the single strand <span class="hlt">break</span> yields and 45-80% for the double strand <span class="hlt">break</span> yields. Among the yields simulated with cross sections of the three different dispersion models, generally the greatest are those of the extended-Drude dispersion model, the second are those of the extended-oscillator-Drude dispersion model, and the last are those of the Ashley's δ-oscillator dispersion model. For the single strand <span class="hlt">break</span> yields, the differences between the first two are very little and the differences between the last two are about 6-57%. For the double strand <span class="hlt">break</span> yields, the biggest difference between the first two can be about 90% and the differences between the last two are about 17-70%.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3440718','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3440718"><span>Trifluorothymidine exhibits potent antitumor activity via the induction of DNA double-strand <span class="hlt">breaks</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>SUZUKI, NORIHIKO; NAKAGAWA, FUMIO; NUKATSUKA, MAMORU; FUKUSHIMA, MASAKAZU</p> <p>2011-01-01</p> <p>TAS-102 is an oral anticancer drug composed of trifluorothymidine (TFT) and TPI (an inhibitor of thymidine phosphorylase that strongly inhibits the biodegradation of TFT). Similar to 5-fluorouracil (5FU) and 5-fluoro-<span class="hlt">2</span>′-deoxyuridine (FdUrd), TFT also inhibits thymidylate synthase (TS), a rate-limiting enzyme of DNA biosynthesis, and is incorporated into DNA. TFT exhibits an anticancer effect on colorectal cancer cells that have acquired 5FU and/or FdUrd resistance as a result of the overexpression of TS. Therefore, we examined the mode of action of TFT-<span class="hlt">induced</span> DNA damage after its incorporation into DNA. When HeLa cells were treated with TFT, the number of ring-open aldehyde forms at apurinic/apyrimidinic sites increased in a dose-dependent manner, although we previously reported that no detectable excisions of TFT paired to adenine were observed using uracil DNA glycosylases, thymine DNA glycosylase or methyl-Cp<span class="hlt">G</span> binding domain 4 and HeLa whole cell extracts. To investigate the functional mechanism of TFT-<span class="hlt">induced</span> DNA damage, we measured the phosphorylation of ATR, ATM, BRCA<span class="hlt">2</span>, chk1 and chk<span class="hlt">2</span> in nuclear extracts of HeLa cells after 0, 24, 48 or 72 h of exposure to an IC50 concentration of TFT, FdUrd or 5FU using Western blot analysis or an enzyme-linked immunosorbent assay (ELISA). Unlike FdUrd and 5FU, TFT resulted in an earlier phosphorylation of ATR and chk1 proteins after only 24 h of exposure, while phosphorylated ATM, BRCA<span class="hlt">2</span> and chk<span class="hlt">2</span> proteins were detected after more than 48 h of exposure to TFT. These results suggest that TFT causes single-strand <span class="hlt">breaks</span> followed by double-strand <span class="hlt">breaks</span> in the DNA of TFT-treated cells. TFT (as TAS-102) showed a more potent antitumor activity than oral 5FU on CO-3 colon cancer xenografts in mice, and such antitumor potency was supported by the increased number of double-strand <span class="hlt">breaks</span> occurring after single-strand <span class="hlt">breaks</span> in the DNA of the TFT-treated tumors. These results suggest that TFT causes single-strand <span class="hlt">breaks</span> after its</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28363167','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28363167"><span>Emodin targets mitochondrial cyclophilin D to <span class="hlt">induce</span> apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Ling; He, Dian; Li, Kun; Liu, Hongli; Wang, Baitao; Zheng, Lifang; Li, Jiazhong</p> <p>2017-06-01</p> <p>Emodin has demonstrated potent anticancer activity in human hepatocarcinoma cells and animal models, however, the cellular targets of emodin have not been fully defined. Here we report that emodin <span class="hlt">induces</span> the dysfunction of mitochondria and the apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells through an enrichment in mitochondria. Specifically, A mitochondrial matrix protein (cyclophilin D, CyPD) is involved in emodin-<span class="hlt">induced</span> apoptosis, and the inhibitor of CyPD (cyclosporin A) could almost completely suppressing the apoptosis; Moreover, as the expression of CyPD could be effectively inhibited by antioxidant N-acetyl-l-cysteine and epidermal growth factor (the activator of ERK), reactive oxygen species and ERK might be involved in the relevant role of CyPD. A further molecule-docking discloses the existence of three hydrogen-bonds in CyPD-emodin complex. Thus, target localization and CyPD in mitochondria provides an insight into the action of emodin in the treatment of liver cancer. Copyright © 2017 Elsevier Masson SAS. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3082747','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3082747"><span>Tetracycline-<span class="hlt">inducible</span> protein expression in pancreatic cancer cells: Effects of Cap<span class="hlt">G</span> overexpression</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Tonack, Sarah; Patel, Sabina; Jalali, Mehdi; Nedjadi, Taoufik; Jenkins, Rosalind E; Goldring, Christopher; Neoptolemos, John; Costello, Eithne</p> <p>2011-01-01</p> <p>AIM: To establish stable tetracycline-<span class="hlt">inducible</span> pancreatic cancer cell lines. METHODS: Suit-<span class="hlt">2</span>, MiaPaca-<span class="hlt">2</span>, and Panc-1 cells were transfected with a second generation reverse tetracycline-controlled transactivator protein (rtTA<span class="hlt">2</span>S-M<span class="hlt">2</span>), under the control of either a cytomegalovirus (CMV) or a chicken β-actin promoter, and the resulting clones were characterised. RESULTS: Use of the chicken (β-actin) promoter proved superior for both the production and maintenance of doxycycline-<span class="hlt">inducible</span> cell lines. The system proved versatile, enabling transient <span class="hlt">inducible</span> expression of a variety of genes, including GST-P, CYP<span class="hlt">2</span>E1, S100A6, and the actin capping protein, Cap<span class="hlt">G</span>. To determine the physiological utility of this system in pancreatic cancer cells, stable <span class="hlt">inducible</span> Cap<span class="hlt">G</span> expressors were established. Overexpressed Cap<span class="hlt">G</span> was localised to the cytoplasm and the nuclear membrane, but was not observed in the nucleus. High Cap<span class="hlt">G</span> levels were associated with enhanced motility, but not with changes to the cell cycle, or cellular proliferation. In Cap<span class="hlt">G</span>-overexpressing cells, the levels and phosphorylation status of other actin-moduating proteins (Cofilin and Ezrin/Radixin) were not altered. However, preliminary analyses suggest that the levels of other cellular proteins, such as ornithine aminotransferase and enolase, are altered upon Cap<span class="hlt">G</span> induction. CONCLUSION: We have generated pancreatic-cancer derived cell lines in which gene expression is fully controllable. PMID:21528072</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3631331','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3631331"><span>Quantitative analysis of the thermal requirements for stepwise physical dormancy-<span class="hlt">break</span> in seeds of the winter annual Geranium carolinianum (Geraniaceae)</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Gama-Arachchige, N. S.; Baskin, J. M.; Geneve, R. L.; Baskin, C. C.</p> <p>2013-01-01</p> <p>Background and Aims Physical dormancy (PY)-<span class="hlt">break</span> in some annual plant species is a two-step process controlled by two different temperature and/or moisture regimes. The thermal time model has been used to quantify PY-<span class="hlt">break</span> in several species of Fabaceae, but not to describe stepwise PY-<span class="hlt">break</span>. The primary aims of this study were to quantify the thermal requirement for sensitivity induction by developing a thermal time model and to propose a mechanism for stepwise PY-<span class="hlt">breaking</span> in the winter annual Geranium carolinianum. Methods Seeds of <span class="hlt">G</span>. carolinianum were stored under dry conditions at different constant and alternating temperatures to <span class="hlt">induce</span> sensitivity (step I). Sensitivity induction was analysed based on the thermal time approach using the Gompertz function. The effect of temperature on step II was studied by incubating sensitive seeds at low temperatures. Scanning electron microscopy, penetrometer techniques, and different humidity levels and temperatures were used to explain the mechanism of stepwise PY-<span class="hlt">break</span>. Key Results The base temperature (Tb) for sensitivity induction was 17·<span class="hlt">2</span> °C and constant for all seed fractions of the population. Thermal time for sensitivity induction during step I in the PY-<span class="hlt">breaking</span> process agreed with the three-parameter Gompertz model. Step II (PY-<span class="hlt">break</span>) did not agree with the thermal time concept. Q10 values for the rate of sensitivity induction and PY-<span class="hlt">break</span> were between <span class="hlt">2</span>·0 and 3·5 and between 0·02 and 0·1, respectively. The force required to separate the water gap palisade layer from the sub-palisade layer was significantly reduced after sensitivity induction. Conclusions Step I and step II in PY-<span class="hlt">breaking</span> of <span class="hlt">G</span>. carolinianum are controlled by chemical and physical processes, respectively. This study indicates the feasibility of applying the developed thermal time model to predict or manipulate sensitivity induction in seeds with two-step PY-<span class="hlt">breaking</span> processes. The model is the first and most detailed one yet developed for</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21999439','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21999439"><span>Evaluation of 8-hydroxy-<span class="hlt">2</span>'-deoxyguanosine (8-OHd<span class="hlt">G</span>) adduct levels and DNA strand <span class="hlt">breaks</span> in human peripheral blood lymphocytes exposed in vitro to polycyclic aromatic hydrocarbons with or without animal metabolic activation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Isabel, Rodríguez-Romero María; Sandra, Gómez-Arroyo; Rafael, Villalobos-Pietrini; Carmen, Martínez-Valenzuela; Josefina, Cortés-Eslava; del Carmen, Calderón-Ezquerro María; Rocío, García-Martínez; Francisco, Arenas-Huertero; Elena, Calderón-Segura María</p> <p>2012-04-01</p> <p>The polycyclic aromatic hydrocarbons (PAHs) dibenzo(a,h)anthracene, benzo(ghi)perylene, benzo(b)fluoranthene and benzo(a)pyrene have been identified in urban air from Mexico City and some of them are classified as human carcinogens. In the present study, human peripheral blood lymphocytes were exposed in vitro to different concentrations of PAHs with (+S9) or without (-S9) metabolic activation. The genotoxic and cytotoxic effects of each PAH were examined with an alkaline comet assay and trypan blue dye exclusion, and oxidative DNA damage was determined via the detection of 8-hydroxy-<span class="hlt">2</span>'-deoxyguanosine (8-Ohd<span class="hlt">G</span>) adduct levels by enzyme-linked immunosorbent assay (ELISA). The DNA damage was evaluated with two genotoxicity parameters: the frequency of comets and the comet tail length. Concentrations of 20, 40, 80, 160 and 320 µM DB(a,h)A-S9; 20, 40, 80, 160 and 240 µM B(ghi)P-S9; 20, 30, 40, 60 and 80 µM B(b)F-S9; and 80 µM B(a)P-S9 for 24 h <span class="hlt">induced</span> a small but significant increase in the means of comet frequency, in the tail length and in the 8-oHDg levels in relation to the control (0.5% DMSO-S9). However, all PAHs+S9 produced a more significant increase in DNA strand <span class="hlt">breaks</span> and the level of 8-OHd<span class="hlt">G</span> compared with the control (0.5% DMSO+S9), with a concentration-effect relationship. The viability of lymphocytes exposed to all PAHs-S9 and PAHs+S9 was not modified compared with the control. The results of this study demonstrate that the comet and ELISA are rapid, suitable and sensitive methods to detect in vitro PAH-<span class="hlt">induced</span> DNA damage in human peripheral lymphocytes.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li class="active"><span>15</span></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_15 --> <div id="page_16" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li class="active"><span>16</span></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="301"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25402647','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25402647"><span>Palmitic acid-<span class="hlt">induced</span> neuron cell cycle <span class="hlt">G</span><span class="hlt">2</span>/M arrest and endoplasmic reticular stress through protein palmitoylation in SH-SY5Y human neuroblastoma cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hsiao, Yung-Hsuan; Lin, Ching-I; Liao, Hsiang; Chen, Yue-Hua; Lin, Shyh-Hsiang</p> <p>2014-11-13</p> <p>Obesity-related neurodegenerative diseases are associated with elevated saturated fatty acids (SFAs) in the brain. An increase in SFAs, especially palmitic acid (PA), triggers neuron cell apoptosis, causing cognitive function to deteriorate. In the present study, we focused on the specific mechanism by which PA triggers SH-SY5Y neuron cell apoptosis. We found that PA <span class="hlt">induces</span> significant neuron cell cycle arrest in the <span class="hlt">G</span><span class="hlt">2</span>/M phase in SH-SY5Y cells. Our data further showed that <span class="hlt">G</span><span class="hlt">2</span>/M arrest is involved in elevation of endoplasmic reticular (ER) stress according to an increase in p-eukaryotic translation inhibition factor <span class="hlt">2</span>α, an ER stress marker. Chronic exposure to PA also accelerates beta-amyloid accumulation, a pathological characteristic of Alzheimer's disease. Interestingly, SFA-<span class="hlt">induced</span> ER stress, <span class="hlt">G</span><span class="hlt">2</span>/M arrest and cell apoptosis were reversed by treatment with <span class="hlt">2</span>-bromopalmitate, a protein palmitoylation inhibitor. These findings suggest that protein palmitoylation plays a crucial role in SFA-<span class="hlt">induced</span> neuron cell cycle <span class="hlt">G</span><span class="hlt">2</span>/M arrest, ER stress and apoptosis; this provides a novel strategy for preventing SFA-<span class="hlt">induced</span> neuron cell dysfunction.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011JPhCS.304a2040P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011JPhCS.304a2040P"><span>Inflammation response at the transcriptional level of Hep<span class="hlt">G</span><span class="hlt">2</span> cells <span class="hlt">induced</span> by multi-walled carbon nanotubes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Piret, Jean-Pascal; Vankoningsloo, Sébastien; Noël, Florence; Mejia Mendoza, Jorge; Lucas, Stéphane; Saout, Christelle; Toussaint, Olivier</p> <p>2011-07-01</p> <p>Poor information are currently available about the biological effects of multi-walled carbon nanotubes (MWCNT) on the liver. In this study, we evaluated the effects of MWCNT at the transcriptional level on the classical in vitro model of Hep<span class="hlt">G</span><span class="hlt">2</span> hepatocarcinoma cells. The expression levels of 96 transcript species implicated in the inflammatory and immune responses was studied after a 24h incubation of Hep<span class="hlt">G</span><span class="hlt">2</span> cells in presence of raw MWCNT dispersed in water by stirring. Among the 46 transcript species detected, only a few transcripts including mRNA coding for interleukine-7, chemokines receptor of the C-C families CCR7, as well as Endothelin-1, were statistically more abundant after treatment with MWCNT. Altogether, these data indicate that MWCNT can only <span class="hlt">induce</span> a weak inflammatory response in Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27577073','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27577073"><span>Hypoxia <span class="hlt">inducible</span> factor-1 mediates the expression of the immune checkpoint HLA-<span class="hlt">G</span> in glioma cells through hypoxia response element located in exon <span class="hlt">2</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yaghi, Layale; Poras, Isabelle; Simoes, Renata T; Donadi, Eduardo A; Tost, Jörg; Daunay, Antoine; de Almeida, Bibiana Sgorla; Carosella, Edgardo D; Moreau, Philippe</p> <p>2016-09-27</p> <p>HLA-<span class="hlt">G</span> is an immune checkpoint molecule with specific relevance in cancer immunotherapy. It was first identified in cytotrophoblasts, protecting the fetus from maternal rejection. HLA-<span class="hlt">G</span> tissue expression is very restricted but <span class="hlt">induced</span> in numerous malignant tumors such as glioblastoma, contributing to their immune escape. Hypoxia occurs during placenta and tumor development and was shown to activate HLA-<span class="hlt">G</span>. We aimed to elucidate the mechanisms of HLA-<span class="hlt">G</span> activation under conditions combining hypoxia-mimicking treatment and 5-aza-<span class="hlt">2</span>'deoxycytidine, a DNA demethylating agent used in anti-cancer therapy which also <span class="hlt">induces</span> HLA-<span class="hlt">G</span>. Both treatments enhanced the amount of HLA-<span class="hlt">G</span> mRNA and protein in HLA-<span class="hlt">G</span> negative U251MG glioma cells. Electrophoretic Mobility Shift Assays and luciferase reporter gene assays revealed that HLA-<span class="hlt">G</span> upregulation depends on Hypoxia <span class="hlt">Inducible</span> Factor-1 (HIF-1) and a hypoxia responsive element (HRE) located in exon <span class="hlt">2</span>. A polymorphic HRE at -966 bp in the 5'UT region may modulate the magnitude of the response mediated by the exon <span class="hlt">2</span> HRE. We suggest that therapeutic strategies should take into account that HLA-<span class="hlt">G</span> expression in response to hypoxic tumor environment is dependent on HLA-<span class="hlt">G</span> gene polymorphism and DNA methylation state at the HLA-<span class="hlt">G</span> locus.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5325396','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5325396"><span>Hypoxia <span class="hlt">inducible</span> factor-1 mediates the expression of the immune checkpoint HLA-<span class="hlt">G</span> in glioma cells through hypoxia response element located in exon <span class="hlt">2</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Yaghi, Layale; Poras, Isabelle; Simoes, Renata T.; Donadi, Eduardo A.; Tost, Jörg; Daunay, Antoine; de Almeida, Bibiana Sgorla; Carosella, Edgardo D.; Moreau, Philippe</p> <p>2016-01-01</p> <p>HLA-<span class="hlt">G</span> is an immune checkpoint molecule with specific relevance in cancer immunotherapy. It was first identified in cytotrophoblasts, protecting the fetus from maternal rejection. HLA-<span class="hlt">G</span> tissue expression is very restricted but <span class="hlt">induced</span> in numerous malignant tumors such as glioblastoma, contributing to their immune escape. Hypoxia occurs during placenta and tumor development and was shown to activate HLA-<span class="hlt">G</span>. We aimed to elucidate the mechanisms of HLA-<span class="hlt">G</span> activation under conditions combining hypoxia-mimicking treatment and 5-aza-<span class="hlt">2</span>′deoxycytidine, a DNA demethylating agent used in anti-cancer therapy which also <span class="hlt">induces</span> HLA-<span class="hlt">G</span>. Both treatments enhanced the amount of HLA-<span class="hlt">G</span> mRNA and protein in HLA-<span class="hlt">G</span> negative U251MG glioma cells. Electrophoretic Mobility Shift Assays and luciferase reporter gene assays revealed that HLA-<span class="hlt">G</span> upregulation depends on Hypoxia <span class="hlt">Inducible</span> Factor-1 (HIF-1) and a hypoxia responsive element (HRE) located in exon <span class="hlt">2</span>. A polymorphic HRE at −966 bp in the 5′UT region may modulate the magnitude of the response mediated by the exon <span class="hlt">2</span> HRE. We suggest that therapeutic strategies should take into account that HLA-<span class="hlt">G</span> expression in response to hypoxic tumor environment is dependent on HLA-<span class="hlt">G</span> gene polymorphism and DNA methylation state at the HLA-<span class="hlt">G</span> locus. PMID:27577073</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28381351','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28381351"><span>Dihydromyricetin <span class="hlt">induces</span> mitochondria-mediated apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells through down-regulation of the Akt/Bad pathway.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Zhuangwei; Zhang, Huiqin; Chen, Shiyong; Xu, Yan; Yao, Anjun; Liao, Qi; Han, Liyuan; Zou, Zuquan; Zhang, Xiaohong</p> <p>2017-02-01</p> <p>The plant flavonol dihydromyricetin (DHM) was reported to <span class="hlt">induce</span> apoptosis in human hepatocarcinoma Hep<span class="hlt">G</span><span class="hlt">2</span> cells. This study was undertaken to elucidate the underlying molecular mechanism of action of DHM. In the study, DHM down-regulated Akt expression and its phosphorylation at Ser473, up-regulated the levels of mitochondrial proapoptotic proteins Bax and Bad, and inhibited the phosphorylation of Bad at Ser136 and Ser112. It also inhibited the expression of the antiapoptotic protein Bcl-<span class="hlt">2</span> and enhanced the cleavage and activation of caspase-3 as well as the degradation of its downstream target poly(ADP-ribose) polymerase. Our results for the first time suggest that DHM-<span class="hlt">induced</span> apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells may come about by the inhibition of the Akt/Bad signaling pathway and stimulation of the mitochondrial apoptotic pathway. Dihydromyricetin may be a promising therapeutic medication for hepatocellular carcinoma. Copyright © 2017 Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5817260','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5817260"><span>Zanthoxylum ailanthoides Suppresses Oleic Acid-<span class="hlt">Induced</span> Lipid Accumulation through an Activation of LKB1/AMPK Pathway in Hep<span class="hlt">G</span><span class="hlt">2</span> Cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Kwon, Eun-Bin; Kang, Myung-Ji; Kim, Soo-Yeon; Lee, Yong-Moon; Lee, Mi-Kyeong; Yuk, Heung Joo; Ryu, Hyung Won; Lee, Su Ui</p> <p>2018-01-01</p> <p>Zanthoxylum ailanthoides (ZA) has been used as folk medicines in East Asian and recently reported to have several bioactivity; however, the studies of ZA on the regulation of triacylglycerol (TG) biosynthesis have not been elucidated yet. In this study, we examined whether the methanol extract of ZA (ZA-M) could reduce oleic acid- (OA-) <span class="hlt">induced</span> intracellular lipid accumulation and confirmed its mode of action in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. ZA-M was shown to promote the phosphorylation of AMPK and its upstream LKB1, followed by reduction of lipogenic gene expressions. As a result, treatment of ZA-M blocked de novo TG biosynthesis and subsequently mitigated intracellular neutral lipid accumulation in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. ZA-M also inhibited OA-<span class="hlt">induced</span> production of reactive oxygen species (ROS) and TNF-α, suggesting that ZA-M possess the anti-inflammatory feature in fatty acid over accumulated condition. Taken together, these results suggest that ZA-M attenuates OA-<span class="hlt">induced</span> lipid accumulation and inflammation through the activation of LKB1/AMPK signaling pathway in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. PMID:29507591</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29507591','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29507591"><span>Zanthoxylum ailanthoides Suppresses Oleic Acid-<span class="hlt">Induced</span> Lipid Accumulation through an Activation of LKB1/AMPK Pathway in Hep<span class="hlt">G</span><span class="hlt">2</span> Cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kwon, Eun-Bin; Kang, Myung-Ji; Kim, Soo-Yeon; Lee, Yong-Moon; Lee, Mi-Kyeong; Yuk, Heung Joo; Ryu, Hyung Won; Lee, Su Ui; Oh, Sei-Ryang; Moon, Dong-Oh; Lee, Hyun-Sun; Kim, Mun-Ock</p> <p>2018-01-01</p> <p>Zanthoxylum ailanthoides (ZA) has been used as folk medicines in East Asian and recently reported to have several bioactivity; however, the studies of ZA on the regulation of triacylglycerol (TG) biosynthesis have not been elucidated yet. In this study, we examined whether the methanol extract of ZA (ZA-M) could reduce oleic acid- (OA-) <span class="hlt">induced</span> intracellular lipid accumulation and confirmed its mode of action in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. ZA-M was shown to promote the phosphorylation of AMPK and its upstream LKB1, followed by reduction of lipogenic gene expressions. As a result, treatment of ZA-M blocked de novo TG biosynthesis and subsequently mitigated intracellular neutral lipid accumulation in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. ZA-M also inhibited OA-<span class="hlt">induced</span> production of reactive oxygen species (ROS) and TNF- α , suggesting that ZA-M possess the anti-inflammatory feature in fatty acid over accumulated condition. Taken together, these results suggest that ZA-M attenuates OA-<span class="hlt">induced</span> lipid accumulation and inflammation through the activation of LKB1/AMPK signaling pathway in Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5689654','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5689654"><span>The NEIL1 <span class="hlt">G</span>83D germline DNA glycosylase variant <span class="hlt">induces</span> genomic instability and cellular transformation</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Galick, Heather A.; Marsden, Carolyn G.; Kathe, Scott; Dragon, Julie A.; Volk, Lindsay; Nemec, Antonia A.; Wallace, Susan S.; Prakash, Aishwarya; Doublié, Sylvie; Sweasy, Joann B.</p> <p>2017-01-01</p> <p>Base excision repair (BER) is a key genome maintenance pathway. The NEIL1 DNA glycosylase recognizes oxidized bases, and likely removes damage in advance of the replication fork. The rs5745906 SNP of the NEIL1 gene is a rare human germline variant that encodes the NEIL1 <span class="hlt">G</span>83D protein, which is devoid of DNA glycosylase activity. Here we show that expression of <span class="hlt">G</span>83D NEIL1 in MCF10A immortalized but non-transformed mammary epithelial cells leads to replication fork stress. Upon treatment with hydrogen peroxide, we observe increased levels of stalled replication forks in cells expressing <span class="hlt">G</span>83D NEIL1 versus cells expressing the wild-type (WT) protein. Double-strand <span class="hlt">breaks</span> (DSBs) arise in <span class="hlt">G</span>83D-expressing cells during the S and <span class="hlt">G</span><span class="hlt">2</span>/M phases of the cell cycle. Interestingly, these <span class="hlt">breaks</span> result in genomic instability in the form of high levels of chromosomal aberrations and micronuclei. Cells expressing <span class="hlt">G</span>83D also grow in an anchorage independent manner, suggesting that the genomic instability results in a carcinogenic phenotype. Our results are consistent with the idea that an inability to remove oxidative damage in an efficient manner at the replication fork leads to genomic instability and mutagenesis. We suggest that individuals who harbor the <span class="hlt">G</span>83D NEIL1 variant face an increased risk for human cancer. PMID:29156764</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3555995','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3555995"><span>Genistein abrogates <span class="hlt">G</span><span class="hlt">2</span> arrest <span class="hlt">induced</span> by curcumin in p53 deficient T47D cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p></p> <p>2012-01-01</p> <p>Background The high cost and low level of cancer survival urge the finding of new drugs having better mechanisms. There is a high trend of patients to be “back to nature” and use natural products as an alternative way to cure cancer. The fact is that some of available anticancer drugs are originated from plants, such as taxane, vincristine, vinblastine, pacitaxel. Curcumin (diferuloylmethane), a dietary pigment present in Curcuma longa rizhome is reported to <span class="hlt">induce</span> cell cycle arrest in some cell lines. Other study reported that genistein isolated from Glycine max seed inhibited phosphorylation of cdk1, gene involved during <span class="hlt">G</span><span class="hlt">2</span>/M transition and thus could function as <span class="hlt">G</span><span class="hlt">2</span> checkpoint abrogator. The inhibition of cdk1 phosphorylation is one of alternative strategy which could selectively kill cancer cells and potentially be combined with DNA damaging agent such as curcumin. Methods T47D cell line was treated with different concentrations of curcumin and genistein, alone or in combination; added together or with interval time. Flow Cytometry and MTT assay were used to evaluate cell cycle distribution and viability, respectively. The presence of apoptotic cells was determined using acridine orange-ethidium bromide staining. Results In this study curcumin <span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span> arrest on p53 deficient T47D cells at the concentration of 10 μM. Increasing concentration up to 30 μM increased the number of cell death. Whilst genistein alone at low concentration (≤10 μM) <span class="hlt">induced</span> cell proliferation, addition of genistein (20 μM) 16 h after curcumin resulted in more cell death (89%), 34% higher than that administered at the same time (56%). The combination treatment resulted in apoptotic cell death. Combining curcumin with high dose of genistein (50 μM) <span class="hlt">induced</span> necrotic cells. Conclusions Genistein increased the death of curcumin treated T47D cells. Appropriate timing of administration and concentration of genistein determine the outcome of treatment and this method</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26644466','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26644466"><span>Antigenic Determinants of the Bilobal Cockroach Allergen Bla <span class="hlt">g</span> <span class="hlt">2</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Woodfolk, Judith A; Glesner, Jill; Wright, Paul W; Kepley, Christopher L; Li, Mi; Himly, Martin; Muehling, Lyndsey M; Gustchina, Alla; Wlodawer, Alexander; Chapman, Martin D; Pomés, Anna</p> <p>2016-01-29</p> <p>Bla <span class="hlt">g</span> <span class="hlt">2</span> is a major indoor cockroach allergen associated with the development of asthma. Antigenic determinants on Bla <span class="hlt">g</span> <span class="hlt">2</span> were analyzed by mutagenesis based on the structure of the allergen alone and in complex with monoclonal antibodies that interfere with IgE antibody binding. The structural analysis revealed mechanisms of allergen-antibody recognition through cation-π interactions. Single and multiple Bla <span class="hlt">g</span> <span class="hlt">2</span> mutants were expressed in Pichia pastoris and purified. The triple mutant K132A/K251A/F162Y showed an ∼100-fold reduced capacity to bind IgE, while preserving the native molecular fold, as proven by x-ray crystallography. This mutant was still able to <span class="hlt">induce</span> mast cell release. T-cell responses were assessed by analyzing Th1/Th<span class="hlt">2</span> cytokine production and the CD4(+) T-cell phenotype in peripheral blood mononuclear cell cultures. Although T-cell activating capacity was similar for the KKF mutant and Bla <span class="hlt">g</span> <span class="hlt">2</span> based on CD25 expression, the KKF mutant was a weaker <span class="hlt">inducer</span> of the Th<span class="hlt">2</span> cytokine IL-13. Furthermore, this mutant <span class="hlt">induced</span> IL-10 from a non-T-cell source at higher levels that those <span class="hlt">induced</span> by Bla <span class="hlt">g</span> <span class="hlt">2</span>. Our findings demonstrate that a rational design of site-directed mutagenesis was effective in producing a mutant with only 3 amino acid substitutions that maintained the same fold as wild type Bla <span class="hlt">g</span> <span class="hlt">2</span>. These residues, which were involved in IgE antibody binding, endowed Bla <span class="hlt">g</span> <span class="hlt">2</span> with a T-cell modulatory capacity. The antigenic analysis of Bla <span class="hlt">g</span> <span class="hlt">2</span> will be useful for the subsequent development of recombinant allergen vaccines. © 2016 by The American Society for Biochemistry and Molecular Biology, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5886692','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5886692"><span>DNA strand-exchange patterns associated with double-strand <span class="hlt">break-induced</span> and spontaneous mitotic crossovers in Saccharomyces cerevisiae</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p></p> <p>2018-01-01</p> <p>Mitotic recombination can result in loss of heterozygosity and chromosomal rearrangements that shape genome structure and initiate human disease. Engineered double-strand <span class="hlt">breaks</span> (DSBs) are a potent initiator of recombination, but whether spontaneous events initiate with the breakage of one or both DNA strands remains unclear. In the current study, a crossover (CO)-specific assay was used to compare heteroduplex DNA (hetDNA) profiles, which reflect strand exchange intermediates, associated with DSB-<span class="hlt">induced</span> versus spontaneous events in yeast. Most DSB-<span class="hlt">induced</span> CO products had the two-sided hetDNA predicted by the canonical DSB repair model, with a switch in hetDNA position from one product to the other at the position of the <span class="hlt">break</span>. Approximately 40% of COs, however, had hetDNA on only one side of the initiating <span class="hlt">break</span>. This anomaly can be explained by a modified model in which there is frequent processing of an early invasion (D-loop) intermediate prior to extension of the invading end. Finally, hetDNA tracts exhibited complexities consistent with frequent expansion of the DSB into a gap, migration of strand-exchange junctions, and template switching during gap-filling reactions. hetDNA patterns in spontaneous COs isolated in either a wild-type background or in a background with elevated levels of reactive oxygen species (tsa1Δ mutant) were similar to those associated with the DSB-<span class="hlt">induced</span> events, suggesting that DSBs are the major instigator of spontaneous mitotic recombination in yeast. PMID:29579095</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19478548','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19478548"><span>Cytotoxic effects of <span class="hlt">2</span>-methoxyestradiol in the hepatocellular carcinoma cell line Hep<span class="hlt">G</span><span class="hlt">2</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>El Naga, Reem N Abou; El-Demerdash, Ebtehal; Youssef, Samar S; Abdel-Naim, Ashraf B; El-Merzabani, Mahmoud</p> <p>2009-01-01</p> <p>The study was designed to examine the potential cytotoxicity of <span class="hlt">2</span>-methoxyestradiol (<span class="hlt">2</span>ME<span class="hlt">2</span>), a natural 17beta-estradiol metabolite, in hepatocellular carcinoma and the possible underlying mechanisms for this cytotoxicity. The cell line Hep<span class="hlt">G</span><span class="hlt">2</span> was treated with different concentrations of <span class="hlt">2</span>ME<span class="hlt">2</span> for 48 and 72 h. Using the sulforhodamine B assay, Hep<span class="hlt">G</span><span class="hlt">2</span> was sensitive to the cytotoxic effect of <span class="hlt">2</span>ME<span class="hlt">2</span>. <span class="hlt">2</span>ME<span class="hlt">2</span> <span class="hlt">induced</span> cell arrest at the <span class="hlt">G</span>(<span class="hlt">2</span>)/M phase and a significant high percentage of apoptotic cells compared to the control group. Also, <span class="hlt">2</span>ME<span class="hlt">2</span> <span class="hlt">induced</span> a significant increase in caspase 9 enzymatic activity after 48 and 72 h of treatment compared with control values. The DNA laddering was observed only in cells treated for 72 h. Furthermore, <span class="hlt">2</span>ME<span class="hlt">2</span> <span class="hlt">induced</span> a significant decrease in the expression levels of vascular endothelial growth factor (VEGF) gene compared to the control values. <span class="hlt">2</span>ME<span class="hlt">2</span> exerts cytotoxic activity in the Hep<span class="hlt">G</span><span class="hlt">2</span> cell line by preferential cell blocking at the <span class="hlt">G</span>(<span class="hlt">2</span>)/M phase as well as induction of apoptosis as evidenced by increased caspase 9 enzymatic activity and observed DNA laddering in <span class="hlt">2</span>ME<span class="hlt">2</span>-treated Hep<span class="hlt">G</span><span class="hlt">2</span> cells. In addition, a reduction in hypervascularity is an important postulated mechanism as indicated by the significant reduction in the expression of VGEF, one of the most important angiogenic factors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17353163','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17353163"><span>IL-27 <span class="hlt">induces</span> the production of Ig<span class="hlt">G</span>1 by human B cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Boumendjel, Amel; Tawk, Lina; Malefijt, René de Waal; Boulay, Vera; Yssel, Hans; Pène, Jérôme</p> <p>2006-12-01</p> <p>It has been reported that IL-27 specifically <span class="hlt">induces</span> the production of Ig<span class="hlt">G</span><span class="hlt">2</span>a by mouse B cells and inhibits IL-4-<span class="hlt">induced</span> Ig<span class="hlt">G</span>1 synthesis. Here, we show that human naïve cord blood expresses a functional IL-27 receptor, consisting of the TCCR and gp130 subunits, although at lower levels as compared to naïve and memory splenic B cells. IL-27 does not <span class="hlt">induce</span> proliferative responses and does not increase Ig<span class="hlt">G</span>1 production by CD19(+)CD27(+) memory B cells. However, it <span class="hlt">induces</span> a low, but significant production of Ig<span class="hlt">G</span>1 by naïve CD19(+)CD27(-)IgD(+)Ig<span class="hlt">G</span>(-) spleen and cord blood B cells, activated via CD40, whereas it has no effect on the production of the other Ig<span class="hlt">G</span> subclasses. In addition, IL-27 <span class="hlt">induces</span> the differentiation of a population of B cells that express high levels of CD38, in association with a down-regulation of surface IgD expression, and that are surface Ig<span class="hlt">G</span>(+/int), CD20(low), CD27(high), indicating that IL-27 promotes isotype switching and plasma cell differentiation of naive B cells. However, as compared to the effects of IL-21 and IL-10, both switch factors for human Ig<span class="hlt">G</span>1 and Ig<span class="hlt">G</span>3, those of IL-27 are modest and regulate exclusively the production of Ig<span class="hlt">G</span>1. Finally, although IL-27 has no effect on IL-4 and anti-CD40-<span class="hlt">induced</span> Cepsilon germline promoter activity, it up-regulates IL-4-<span class="hlt">induced</span> IgE production by naive B cells. These results point to a partial redundancy of switch factors regulating the production of Ig<span class="hlt">G</span>1 in humans, and furthermore indicate the existence of a common regulation of the human Ig<span class="hlt">G</span>1and murine Ig<span class="hlt">G</span><span class="hlt">2</span>a isotypes by IL-27.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24811113','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24811113"><span>Rule <span class="hlt">breaking</span> mediates the developmental association between GABRA<span class="hlt">2</span> and adolescent substance abuse.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Trucco, Elisa M; Villafuerte, Sandra; Heitzeg, Mary M; Burmeister, Margit; Zucker, Robert A</p> <p>2014-12-01</p> <p>This study's primary aim was to examine age-specific associations between GABRA<span class="hlt">2</span>, rule <span class="hlt">breaking</span>, problematic alcohol use, and substance abuse symptomatology. The secondary aim was to examine the extent to which rule <span class="hlt">breaking</span> mediates the GABRA<span class="hlt">2</span>-substance abuse relationship. A sample (n = 518) of primarily male (70.9%) and White (88.8%) adolescents from the Michigan Longitudinal Study was assessed from ages 11-18. Age-specific effects of GABRA<span class="hlt">2</span> on rule <span class="hlt">breaking</span>, problematic alcohol use, and substance abuse symptomatology were examined using nested path models. The role of rule <span class="hlt">breaking</span> as a mediator in the association between GABRA<span class="hlt">2</span> and substance abuse outcomes was tested using prospective cross-lagged path models. GABRA<span class="hlt">2</span> is significantly (p < 0.05) associated with rule <span class="hlt">breaking</span> in mid- to late-adolescence, but not substance abuse symptomatology across adolescence. GABRA<span class="hlt">2</span> effects on problematic alcohol use and substance abuse symptomatology operate largely (45.3% and 71.1%, respectively, p < 0.05) via rule <span class="hlt">breaking</span> in midadolescence. GABRA<span class="hlt">2</span> represents an early risk factor for an externalizing pathway to the development of problematic alcohol and drug use. © 2014 The Authors. Journal of Child Psychology and Psychiatry. © 2014 Association for Child and Adolescent Mental Health.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA013336','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA013336"><span>Laboratory Investigation of Wave <span class="hlt">Breaking</span>. Part <span class="hlt">2</span>. Deep Water Waves</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1975-06-01</p> <p>respectively, phase velocity is given implicitly by: C3 = [ + (f )<span class="hlt">2</span>] ( Levi - Civita , 1925) (<span class="hlt">2</span>a)C3 CS = F (1 + (c_-_)<span class="hlt">2</span> + (fH)4 (Beach Erosion Board, 1941...In view of the above, one is led to wonder why almost all wave- 4 oriented research within the past two decades has been directed towards wave growth...mechanisms, as opposed to wave <span class="hlt">breaking</span>. There seem to be ’’ at least two reasors. Wave <span class="hlt">breaking</span>--aidefined by turbulent energy loss- -is a non</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28263848','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28263848"><span>Phaleria macrocarpa (Boerl.) fruit <span class="hlt">induce</span> <span class="hlt">G</span>0/<span class="hlt">G</span>1 and <span class="hlt">G</span><span class="hlt">2</span>/M cell cycle arrest and apoptosis through mitochondria-mediated pathway in MDA-MB-231 human breast cancer cell.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kavitha, Nowroji; Ein Oon, Chern; Chen, Yeng; Kanwar, Jagat R; Sasidharan, Sreenivasan</p> <p>2017-04-06</p> <p>Phaleria macrocarpa (Scheff) Boerl, is a well-known folk medicinal plant in Indonesia. Traditionally, P. macrocarpa has been used to control cancer, impotency, hemorrhoids, diabetes mellitus, allergies, liver and hearth disease, kidney disorders, blood diseases, acne, stroke, migraine, and various skin diseases. The purpose of this study was to determine the in situ cytotoxicity effect P. macrocarpa fruit ethyl acetate fraction (PMEAF) and the underlying molecular mechanism of cell death. MDA-MB-231 cells were incubated with PMEAF for 24h. Cell cycle and viability were examined using flow cytometry analysis. Apoptosis was determined using the Annexin V assay and also by fluorescence microscopy. Apoptosis protein profiling was detected by RayBio® Human Apoptosis Array. The AO/PI staining and flow cytometric analysis of MDA-MB-231 cells treated with PMEAF were showed apoptotic cell death. The cell cycle analysis by flow cytometry analysis revealed that the accumulation of PMEAF treated MDA-MB-231 cells in <span class="hlt">G</span> 0 /<span class="hlt">G</span> 1 and <span class="hlt">G</span> <span class="hlt">2</span> /M-phase of the cell cycle. Moreover, the PMEAF exert cytotoxicity by increased the ROS production in MDA-MB-231 cells consistently stimulated the loss of mitochondrial membrane potential (∆ Ψm ) and <span class="hlt">induced</span> apoptosis cell death by activation of numerous signalling proteins. The results from apoptosis protein profiling array evidenced that PMEAF stimulated the expression of 9 pro-apoptotic proteins (Bax, Bid, caspase 3, caspase 8, cytochrome c, p21, p27, p53 and SMAC) and suppressed the 4 anti-apoptotic proteins (Bcl-<span class="hlt">2</span>, Bcl-w, XIAP and survivin) in MDA-MB-231 cells. The results indicated that PMEAF treatment <span class="hlt">induced</span> apoptosis in MDA-MB-231 cells through intrinsic mitochondrial related pathway with the participation of pro and anti-apoptotic proteins, caspases, <span class="hlt">G</span> 0 /<span class="hlt">G</span> 1 and <span class="hlt">G</span> <span class="hlt">2</span> /M-phases cell cycle arrest by p53-mediated mechanism. Copyright © 2017 Elsevier Ireland Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23274058','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23274058"><span>Jaceosidin, isolated from dietary mugwort (Artemisia princeps), <span class="hlt">induces</span> <span class="hlt">G</span><span class="hlt">2</span>/M cell cycle arrest by inactivating cdc25C-cdc<span class="hlt">2</span> via ATM-Chk1/<span class="hlt">2</span> activation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lee, Jong-Gyu; Kim, Ji-Hyun; Ahn, Ji-Hye; Lee, Kyung-Tae; Baek, Nam-In; Choi, Jung-Hye</p> <p>2013-05-01</p> <p>Jaceosidin, a flavonoid derived from Artemisia princeps (Japanese mugwort), has been shown to inhibit the growth of several human cancer cells, However, the exact mechanism for the cytotoxic effect of jaceosidin is not completely understood. In this study, we investigated the molecular mechanism involved in the antiproliferative effect of jaceosidin in human endometrial cancer cells. We demonstrated that jaceosidin is a more potent inhibitor of cell growth than cisplatin in human endometrial cancer cells. In contrast, jaceosidin-<span class="hlt">induced</span> cytotoxicity in normal endometrial cells was lower than that observed for cisplatin. Jaceosidin <span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span>/M phase cell cycle arrest and modulated the levels of cyclin B and p-Cdc<span class="hlt">2</span> in Hec1A cells. Knockdown of p21 using specific siRNAs partially abrogated jaceosidin-<span class="hlt">induced</span> cell growth inhibition. Additional mechanistic studies revealed that jaceosidin treatment resulted in an increase in phosphorylation of Cdc25C and ATM-Chk1/<span class="hlt">2</span>. Ku55933, an ATM inhibitor, reversed jaceosidin-<span class="hlt">induced</span> cell growth inhibition, in part. Moreover, jaceosidin treatment resulted in phosphorylation of ERK, and pretreatment with the ERK inhibitor, PD98059, attenuated cell growth inhibition by jaceosidin. These data suggest that jaceosidin, isolated from Japanese mugwort, modulates the ERK/ATM/Chk1/<span class="hlt">2</span> pathway, leading to inactivation of the Cdc<span class="hlt">2</span>-cyclin B1 complex, followed by <span class="hlt">G</span><span class="hlt">2</span>/M cell cycle arrest in endometrial cancer cells. Copyright © 2012 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=117544','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=117544"><span>UV-<span class="hlt">induced</span> replication arrest in the xeroderma pigmentosum variant leads to DNA double-strand <span class="hlt">breaks</span>, γ-H<span class="hlt">2</span>AX formation, and Mre11 relocalization</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Limoli, Charles L.; Giedzinski, Erich; Bonner, William M.; Cleaver, James E.</p> <p>2002-01-01</p> <p>UV-<span class="hlt">induced</span> replication arrest in the xeroderma pigmentosum variant (XPV) but not in normal cells leads to an accumulation of the Mre11/Rad50/Nbs1 complex and phosphorylated histone H<span class="hlt">2</span>AX (γ-H<span class="hlt">2</span>AX) in large nuclear foci at sites of stalled replication forks. These complexes have been shown to signal the presence of DNA damage, in particular, double-strand <span class="hlt">breaks</span> (DSBs). This finding suggests that UV damage leads to the formation of DSBs during the course of replication arrest. After UV irradiation, XPV cells showed a fluence-dependent increase in the yield of γ-H<span class="hlt">2</span>AX foci that paralleled the production of Mre11 foci. The percentage of foci-positive cells increased rapidly (10–15%) up to fluences of 10 J⋅m−<span class="hlt">2</span> before saturating at higher fluences. Frequencies of γ-H<span class="hlt">2</span>AX and Mre11 foci both reached maxima at 4 h after UV irradiation. This pattern contrasts sharply to the situation observed after x-irradiation, where peak levels of γ-H<span class="hlt">2</span>AX foci were found to precede the formation of Mre11 foci by several hours. The nuclear distributions of γ-H<span class="hlt">2</span>AX and Mre11 were found to colocalize spatially after UV- but not x-irradiation. UV-irradiated XPV cells showed a one-to-one correspondence between Mre11 and γ-H<span class="hlt">2</span>AX foci-positive cells. These results show that XPV cells develop DNA DSBs during the course of UV-<span class="hlt">induced</span> replication arrest. These UV-<span class="hlt">induced</span> foci occur in cells that are unable to carry out efficient bypass replication of UV damage and may contribute to further genetic variation. PMID:11756691</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27377964','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27377964"><span>Alpha-lipoic acid attenuates endoplasmic reticulum stress-<span class="hlt">induced</span> insulin resistance by improving mitochondrial function in Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lei, Lin; Zhu, Yiwei; Gao, Wenwen; Du, Xiliang; Zhang, Min; Peng, Zhicheng; Fu, Shoupeng; Li, Xiaobing; Zhe, Wang; Li, Xinwei; Liu, Guowen</p> <p>2016-10-01</p> <p>Alpha-lipoic acid (ALA) has been reported to have beneficial effects for improving insulin sensitivity. However, the underlying molecular mechanism of the beneficial effects remains poorly understood. Endoplasmic reticulum (ER) stress and mitochondrial dysfunction are considered causal factors that <span class="hlt">induce</span> insulin resistance. In this study, we investigated the effect of ALA on the modulation of insulin resistance in ER-stressed Hep<span class="hlt">G</span><span class="hlt">2</span> cells, and we explored the potential mechanism of this effect. Hep<span class="hlt">G</span><span class="hlt">2</span> cells were incubated with tunicamycin (Tun) for 6h to establish an ER stress cell model. Tun treatment <span class="hlt">induced</span> ER stress, mitochondrial dysfunction and insulin resistance. Interestingly, ALA had no significant effect on ER stress signals. Pretreatment of the ER stress cell model with ALA for 24h improved insulin sensitivity, restored the expression levels of mitochondrial oxidative phosphorylation (OXPHOS) complexes and increased intracellular ATP production. Moreover, ALA augmented the β-oxidation capacity of the mitochondria. Importantly, ALA treatment could decrease oligomycin-<span class="hlt">induced</span> mitochondrial dysfunction and then improved insulin resistance. Taken together, our data suggest that ALA prevents ER stress-<span class="hlt">induced</span> insulin resistance by enhancing mitochondrial function. Copyright © 2016 Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17451994','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17451994"><span>Vicia root-mirconucleus and sister <span class="hlt">chromatid</span> exchange assays on the genotoxicity of selenium compounds.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yi, Huilan; Si, Liangyan</p> <p>2007-06-15</p> <p>Selenium (Se) is an important metalloid with industrial, environmental, biological and toxicological significance. Excessive selenium in soil and water may contribute to environmental selenium pollution, and affect plant growth and human health. By using Vicia faba micronucleus (MN) and sister <span class="hlt">chromatid</span> exchange (SCE) tests, possible genotoxicity of sodium selenite and sodium biselenite was evaluated in this study. The results showed that sodium selenite, at concentrations from 0.01 to 10.0mg/L, <span class="hlt">induced</span> a 1.9-3.9-fold increase in MN frequency and a 1.5-1.6-fold increase in SCE frequency, with a statistically significantly difference from the control (P<0.05 and 0.01, respectively). Sodium selenite also caused mitotic delay and a 15-80% decrease in mitotic indices (MI), but at the lowest concentration (0.005mg/L), it slightly stimulated mitotic activity. Similarly, the frequencies of MN and SCE also increased significantly in sodium biselenite treated samples, with MI decline only at relatively higher effective concentrations. Results of the present study suggest that selenite is genotoxic to V. faba root cells and may be a genotoxic risk to human health.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li class="active"><span>16</span></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_16 --> <div id="page_17" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li class="active"><span>17</span></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="321"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1684051','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1684051"><span><span class="hlt">Chromatid</span> repulsion associated with Roberts/SC phocomelia syndrome is reduced in malignant cells and not expressed in interspecies somatic-cell hybrids.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Krassikoff, N E; Cowan, J M; Parry, D M; Francke, U</p> <p>1986-01-01</p> <p>Different cell types from a female patient with Roberts/SC phocomelia syndrome were evaluated quantitatively for the presence of repulsion of heterochromatin and satellite regions of mitotic chromosomes. Whereas EBV-transformed lymphoblasts from an established cell line revealed these phenomena at frequencies equal to those in PHA-stimulated lymphocytes and cultured skin fibroblasts, aneuploid cells from a metastatic melanoma displayed them at 50% lower frequency. Cocultivation of the patient's fibroblasts with either an immortal Chinese hamster cell line or with a human male fibroblast strain carrying a t(4;6)(p14;q21) translocation showed that the phenomenon was not corrected or <span class="hlt">induced</span> by a diffusible factor or by cell-to-cell contact. In each experiment, only the patient's metaphase spreads revealed <span class="hlt">chromatid</span> repulsion. In fusion hybrids between the patient's fibroblasts and an established Chinese hamster cell line, the human chromosomes behaved perfectly normally, suggesting that the gene product which is missing or mutant in Roberts/SC phocomelia syndrome is supplied by the Chinese hamster genome. Images Fig. 1 Fig. <span class="hlt">2</span> Fig. 3 PMID:3788975</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JNR....18..340Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JNR....18..340Y"><span>Nanosilica <span class="hlt">induced</span> dose-dependent cytotoxicity and cell type-dependent multinucleation in Hep<span class="hlt">G</span><span class="hlt">2</span> and L-02 cells</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yu, Yongbo; Duan, Junchao; Li, Yang; Yu, Yang; Hu, Hejing; Wu, Jing; Zhang, Yannan; Li, Yanbo; CaixiaGuo; Zhou, Xianqing; Sun, Zhiwei</p> <p>2016-11-01</p> <p>The prevalent exposure to nanosilica gained concerns about health effects of these particles on human beings. Although nanosilica-<span class="hlt">induced</span> multinucleation has been confirmed previously, the underlying mechanism was still not clear; this study was to investigate the origination of multinucleated cells caused by nanosilica (62 nm) in both Hep<span class="hlt">G</span><span class="hlt">2</span> and L-02 cells. Cell viability and cellular uptake was determined by MTT assay and transmission electron microscope (TEM), respectively. Giemsa staining was applied to detect multinucleation. To clarify the origination of multinucleated cells, fluorescent probes, PKH26 and PKH67, time-lapse observation were further conducted by confocal microscopy. Results indicated that nanosilica particles were internalized into cells and <span class="hlt">induced</span> cytotoxicity in a dose-dependent manner. Quantification analysis showed that nanosilica significantly increased the rates of binucleated and multinucleated cells, which suggested mitotic catastrophe induction. Moreover, dynamic visualization verified that multinucleation resulted from cell fusion in Hep<span class="hlt">G</span><span class="hlt">2</span> cells not in L-02 cells after nanosilica exposure, suggesting cell type-dependent multinucleation formation. Both multinucleation and cell fusion were involved in genetic instability, which emphasized the significance to explore the multinucleation <span class="hlt">induced</span> by nanosilica via environmental, occupational and consumer product exposure.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26192324','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26192324"><span>Disorder of <span class="hlt">G</span><span class="hlt">2</span>-M Checkpoint Control in Aniline-<span class="hlt">Induced</span> Cell Proliferation in Rat Spleen.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Jianling; Wang, Gangduo; Khan, M Firoze</p> <p>2015-01-01</p> <p>Aniline, a toxic aromatic amine, is known to cause hemopoietic toxicity both in humans and animals. Aniline exposure also leads to toxic response in spleen which is characterized by splenomegaly, hyperplasia, fibrosis and the eventual formation of tumors on chronic in vivo exposure. Previously, we have shown that aniline exposure leads to iron overload, oxidative DNA damage, and increased cell proliferation, which could eventually contribute to a tumorigenic response in the spleen. Despite our demonstration that cell proliferation was associated with deregulation of <span class="hlt">G</span>1 phase cyclins and increased expression of <span class="hlt">G</span>1 phase cyclin-dependent kinases (CDKs), molecular mechanisms, especially the regulation of <span class="hlt">G</span><span class="hlt">2</span> phase and contribution of epigenetic mechanisms in aniline-<span class="hlt">induced</span> splenic cellular proliferation remain largely unclear. This study therefore, mainly focused on the regulation of <span class="hlt">G</span><span class="hlt">2</span> phase in an animal model preceding a tumorigenic response. Male Sprague-Dawley rats were given aniline (0.5 mmol/kg/day) in drinking water or drinking water only (controls) for 30 days, and expression of <span class="hlt">G</span><span class="hlt">2</span> phase cyclins, CDK1, CDK inhibitors and miRNAs were measured in the spleen. Aniline treatment resulted in significant increases in cell cycle regulatory proteins, including cyclins A, B and CDK1, particularly phosphor-CDK1, and decreases in CDK inhibitors p21 and p27, which could promote the splenocytes to go through <span class="hlt">G</span><span class="hlt">2</span>/M transition. Our data also showed upregulation of tumor markers Trx-1 and Ref-1 in rats treated with aniline. More importantly, we observed lower expression of miRNAs including Let-7a, miR-15b, miR24, miR-100 and miR-125, and greater expression of CDK inhibitor regulatory miRNAs such as miR-181a, miR-221 and miR-222 in the spleens of aniline-treated animals. Our findings suggest that significant increases in the expression of cyclins, CDK1 and aberrant regulation of miRNAs could lead to an accelerated <span class="hlt">G</span><span class="hlt">2</span>/M transition of the splenocytes, and potentially to a</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16131840','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16131840"><span>Telomere sister <span class="hlt">chromatid</span> exchange in telomerase deficient murine cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Yisong; Giannone, Richard J; Liu, Yie</p> <p>2005-10-01</p> <p>We have recently demonstrated that several types of genomic rearrangements (i.e., telomere sister <span class="hlt">chromatid</span> exchange (T-SCE), genomic-SCE, or end-to-end fusions) were more often detected in long-term cultured murine telomerase deficient embryonic stem (ES) cells than in freshly prepared murine splenocytes, even through they possessed similar frequencies of critically short telomeres. The high rate of genomic rearrangements in telomerase deficient ES cells, when compared to murine splenocytes, may reflect the cultured cells' gained ability to protect chromosome ends with eroded telomeres allowing them to escape "end crisis". However, the possibility that ES cells were more permissive to genomic rearrangements than other cell types or that differences in the microenvironment or genetic background of the animals might consequentially determine the rate of T-SCEs or other genomic rearrangements at critically short telomeres could not be ruled out.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/1898293','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/1898293"><span>Physiologic bases of <span class="hlt">G-induced</span> loss of consciousness (<span class="hlt">G</span>-LOC).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Werchan, P M</p> <p>1991-07-01</p> <p>Exposure of pilots to high sustained +Gz (head to feet) or rapid onset of +Gz can produce a variety of pathophysiologic effects ranging from the loss of peripheral vision to total blackout and, finally, <span class="hlt">G-induced</span> loss of consciousness (<span class="hlt">G</span>-LOC). A <span class="hlt">G</span>-LOC research program divided into four phases has been organized at USAFSAM/Crew Technology Division. In contrast to previous studies in acceleration, this program will focus exclusively on the ultimate problem in <span class="hlt">G</span>-LOC; namely, inadequate cerebral perfusion leading to impaired brain energy metabolism, structure and function. The primary objective of this research program is to identify and arrange chronologically the numerous physiological and biochemical alterations in the brain that comprise the mechanism of <span class="hlt">G</span>-LOC.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017TCry...11.2711H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017TCry...11.2711H"><span>Wave-<span class="hlt">induced</span> stress and <span class="hlt">breaking</span> of sea ice in a coupled hydrodynamic discrete-element wave-ice model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Herman, Agnieszka</p> <p>2017-11-01</p> <p>In this paper, a coupled sea ice-wave model is developed and used to analyze wave-<span class="hlt">induced</span> stress and <span class="hlt">breaking</span> in sea ice for a range of wave and ice conditions. The sea ice module is a discrete-element bonded-particle model, in which ice is represented as cuboid <q>grains</q> floating on the water surface that can be connected to their neighbors by elastic joints. The joints may <span class="hlt">break</span> if instantaneous stresses acting on them exceed their strength. The wave module is based on an open-source version of the Non-Hydrostatic WAVE model (NHWAVE). The two modules are coupled with proper boundary conditions for pressure and velocity, exchanged at every wave model time step. In the present version, the model operates in two dimensions (one vertical and one horizontal) and is suitable for simulating compact ice in which heave and pitch motion dominates over surge. In a series of simulations with varying sea ice properties and incoming wavelength it is shown that wave-<span class="hlt">induced</span> stress reaches maximum values at a certain distance from the ice edge. The value of maximum stress depends on both ice properties and characteristics of incoming waves, but, crucially for ice <span class="hlt">breaking</span>, the location at which the maximum occurs does not change with the incoming wavelength. Consequently, both regular and random (Jonswap spectrum) waves <span class="hlt">break</span> the ice into floes with almost identical sizes. The width of the zone of broken ice depends on ice strength and wave attenuation rates in the ice.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3958969','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3958969"><span>Gold nanoparticle–M<span class="hlt">2</span>e conjugate coformulated with Cp<span class="hlt">G</span> <span class="hlt">induces</span> protective immunity against influenza A virus</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Tao, Wenqian; Ziemer, Katherine S; Gill, Harvinder S</p> <p>2014-01-01</p> <p>Aim: This study aimed to develop a novel influenza A vaccine by conjugating the highly conserved extracellular region of the matrix <span class="hlt">2</span> protein (M<span class="hlt">2</span>e) of influenza A virus to gold nanoparticles (AuNPs) and to test the vaccine in a mouse influenza challenge model. Materials & methods: Citrate-reduced AuNPs (diameter: 12 nm) were synthesized, and characterized by transmission electron microscopy and dynamic light scattering. M<span class="hlt">2</span>e was conjugated to AuNPs through thiol–gold interactions to form M<span class="hlt">2</span>e–AuNP conjugates. Particle stability was confirmed by UV–visible spectra, and M<span class="hlt">2</span>e conjugation was further characterized by x-ray photoelectron spectroscopy. Mice were immunized with M<span class="hlt">2</span>e–AuNPs with or without Cp<span class="hlt">G</span> (cytosine-guanine rich oligonucleotide) as an adjuvant with appropriate control groups. Sera was collected and M<span class="hlt">2</span>e-specific immunoglobulin (Ig<span class="hlt">G</span>) was measured, and immunized mice were challenged with PR8-H1N1 influenza virus. Results: M<span class="hlt">2</span>e-capped AuNPs could be lyophilized and stably resuspended in water. Intranasal vaccination of mice with M<span class="hlt">2</span>e–AuNP conjugates <span class="hlt">induced</span> M<span class="hlt">2</span>e-specific Ig<span class="hlt">G</span> serum antibodies, which significantly increased upon addition of soluble Cp<span class="hlt">G</span> as adjuvant. Upon challenge with lethal PR8, mice vaccinated with M<span class="hlt">2</span>e-AuNP conjugates were only partially protected, while mice that received soluble Cp<span class="hlt">G</span> as adjuvant in addition to M<span class="hlt">2</span>e–AuNP were fully protected. Conclusion: Overall, this study demonstrates the potential of using the M<span class="hlt">2</span>e–AuNP conjugates with Cp<span class="hlt">G</span> as an adjuvant as a platform for developing an influenza A vaccine. PMID:23829488</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25529822','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25529822"><span>3-(3-Hydroxy-4-methoxyphenyl)-4-(3,4,5-trimethoxyphenyl)-1,<span class="hlt">2</span>,5-selenadiazole (<span class="hlt">G</span>-1103), a novel combretastatin A-4 analog, <span class="hlt">induces</span> <span class="hlt">G</span><span class="hlt">2</span>/M arrest and apoptosis by disrupting tubulin polymerization in human cervical HeLa cells and fibrosarcoma HT-1080 cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zuo, Daiying; Guo, Dandan; Jiang, Xuewei; Guan, Qi; Qi, Huan; Xu, Jingwen; Li, Zengqiang; Yang, Fushan; Zhang, Weige; Wu, Yingliang</p> <p>2015-02-05</p> <p>Microtubule is a popular target for anticancer drugs. In this study, we describe the effect 3-(3-hydroxy-4-methoxyphenyl)-4-(3,4,5-trimethoxyphenyl)-1,<span class="hlt">2</span>,5-selenadiazole (<span class="hlt">G</span>-1103), a newly synthesized analog of combretastatin A-4 (CA-4), showing a strong time- and dose-dependent anti-proliferative effect on human cervical cancer HeLa cells and human fibrosarcoma HT-1080 cells. We demonstrated that the growth inhibitory effects of <span class="hlt">G</span>-1103 in HeLa and HT-1080 cells were associated with microtubule depolymerization and proved that <span class="hlt">G</span>-1103 acted as microtubule destabilizing agent. Furthermore, cell cycle analysis revealed that <span class="hlt">G</span>-1103 treatment resulted in cell cycle arrest at the <span class="hlt">G</span><span class="hlt">2</span>/M phase in a time-dependent manner with subsequent apoptosis induction. Western blot analysis revealed that down-regulation of cdc25c and up-regulation of cyclin B1 was related with <span class="hlt">G</span><span class="hlt">2</span>/M arrest in HeLa and HT-1080 cells treatment with <span class="hlt">G</span>-1103. In addition, <span class="hlt">G</span>-1103 <span class="hlt">induced</span> HeLa cell apoptosis by up-regulating cleaved caspase-3, Fas, cleaved caspase-8 expression, which indicated that <span class="hlt">G</span>-1103 <span class="hlt">induced</span> HeLa cell apoptosis was mainly associated with death receptor pathway. However, <span class="hlt">G</span>-1103 <span class="hlt">induced</span> HT-1080 cell apoptosis by up-regulating cleaved caspase-3, Fas, cleaved caspase-8, Bax and cleaved caspase-9 expression and down-regulating anti-apoptotic protein Bcl-<span class="hlt">2</span> expression, which indicated that <span class="hlt">G</span>-1103 <span class="hlt">induced</span> HT-1080 cell apoptosis was associated with both mitochondrial and death receptor pathway. Taken together, all the data demonstrated that <span class="hlt">G</span>-1103 exhibited its antitumor activity through disrupting the microtubule assembly, causing cell cycle arrest and consequently <span class="hlt">inducing</span> apoptosis in HeLa and HT-1080 cells. Therefore, the novel compound <span class="hlt">G</span>-1103 is a promising microtubule inhibitor that has great potentials for therapeutic treatment of various malignancies. Copyright © 2014 Elsevier Ireland Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28538144','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28538144"><span>Probing the Potential Role of Non-B DNA Structures at Yeast Meiosis-Specific DNA Double-Strand <span class="hlt">Breaks</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kshirsagar, Rucha; Khan, Krishnendu; Joshi, Mamata V; Hosur, Ramakrishna V; Muniyappa, K</p> <p>2017-05-23</p> <p>A plethora of evidence suggests that different types of DNA quadruplexes are widely present in the genome of all organisms. The existence of a growing number of proteins that selectively bind and/or process these structures underscores their biological relevance. Moreover, <span class="hlt">G</span>-quadruplex DNA has been implicated in the alignment of four sister <span class="hlt">chromatids</span> by forming parallel guanine quadruplexes during meiosis; however, the underlying mechanism is not well defined. Here we show that a <span class="hlt">G</span>/C-rich motif associated with a meiosis-specific DNA double-strand <span class="hlt">break</span> (DSB) in Saccharomyces cerevisiae folds into <span class="hlt">G</span>-quadruplex, and the C-rich sequence complementary to the <span class="hlt">G</span>-rich sequence forms an i-motif. The presence of <span class="hlt">G</span>-quadruplex or i-motif structures upstream of the green fluorescent protein-coding sequence markedly reduces the levels of gfp mRNA expression in S. cerevisiae cells, with a concomitant decrease in green fluorescent protein abundance, and blocks primer extension by DNA polymerase, thereby demonstrating the functional significance of these structures. Surprisingly, although S. cerevisiae Hop1, a component of synaptonemal complex axial/lateral elements, exhibits strong affinity to <span class="hlt">G</span>-quadruplex DNA, it displays a much weaker affinity for the i-motif structure. However, the Hop1 C-terminal but not the N-terminal domain possesses strong i-motif binding activity, implying that the C-terminal domain has a distinct substrate specificity. Additionally, we found that Hop1 promotes intermolecular pairing between <span class="hlt">G</span>/C-rich DNA segments associated with a meiosis-specific DSB site. Our results support the idea that the <span class="hlt">G</span>/C-rich motifs associated with meiosis-specific DSBs fold into intramolecular <span class="hlt">G</span>-quadruplex and i-motif structures, both in vitro and in vivo, thus revealing an important link between non-B form DNA structures and Hop1 in meiotic chromosome synapsis and recombination. Copyright © 2017 Biophysical Society. Published by Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3964074','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3964074"><span>Activation <span class="hlt">induced</span> deaminase C-terminal domain links DNA <span class="hlt">breaks</span> to end protection and repair during class switch recombination</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Zahn, Astrid; Eranki, Anil K.; Patenaude, Anne-Marie; Methot, Stephen P.; Fifield, Heather; Cortizas, Elena M.; Foster, Paul; Imai, Kohsuke; Durandy, Anne; Larijani, Mani; Verdun, Ramiro E.; Di Noia, Javier M.</p> <p>2014-01-01</p> <p>Activation-<span class="hlt">induced</span> deaminase (AID) triggers antibody class switch recombination (CSR) in B cells by initiating DNA double strand <span class="hlt">breaks</span> that are repaired by nonhomologous end-joining pathways. A role for AID at the repair step is unclear. We show that specific inactivation of the C-terminal AID domain encoded by exon 5 (E5) allows very efficient deamination of the AID target regions but greatly impacts the efficiency and quality of subsequent DNA repair. Specifically eliminating E5 not only precludes CSR but also, causes an atypical, enzymatic activity-dependent dominant-negative effect on CSR. Moreover, the E5 domain is required for the formation of AID-dependent Igh-cMyc chromosomal translocations. DNA <span class="hlt">breaks</span> at the Igh switch regions <span class="hlt">induced</span> by AID lacking E5 display defective end joining, failing to recruit DNA damage response factors and undergoing extensive end resection. These defects lead to nonproductive resolutions, such as rearrangements and homologous recombination that can antagonize CSR. Our results can explain the autosomal dominant inheritance of AID variants with truncated E5 in patients with hyper-IgM syndrome <span class="hlt">2</span> and establish that AID, through the E5 domain, provides a link between DNA damage and repair during CSR. PMID:24591601</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28356992','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28356992"><span>Arctigenin, a natural lignan compound, <span class="hlt">induces</span> <span class="hlt">G</span>0/<span class="hlt">G</span>1 cell cycle arrest and apoptosis in human glioma cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Maimaitili, Aisha; Shu, Zunhua; Cheng, Xiaojiang; Kaheerman, Kadeer; Sikandeer, Alifu; Li, Weimin</p> <p>2017-02-01</p> <p>The aim of the current study was to investigate the anticancer potential of arctigenin, a natural lignan compound, in malignant gliomas. The U87MG and T98<span class="hlt">G</span> human glioma cell lines were treated with various concentrations of arctigenin for 48 h and the effects of arctigenin on the aggressive phenotypes of glioma cells were assessed. The results demonstrated that arctigenin dose-dependently inhibited the growth of U87MG and T98<span class="hlt">G</span> cells, as determined using 3-(4,5-dimethylthiazol-<span class="hlt">2</span>-yl)-<span class="hlt">2</span>,5-diphenyltetrazolium bromide and bromodeoxyuridine incorporation assays. Arctigenin exposure also <span class="hlt">induced</span> a 60-75% reduction in colony formation compared with vehicle-treated control cells. However, arctigenin was not observed to affect the invasiveness of glioma cells. Arctigenin significantly increased the proportion of cells in the <span class="hlt">G</span> 0 /<span class="hlt">G</span> 1 phase and reduced the number of cells in the S phase, as compared with the control group (P<0.05). Western blot analysis demonstrated that arctigenin increased the expression levels of p21, retinoblastoma and p53 proteins, and significantly decreased the expression levels of cyclin D1 and cyclin-dependent kinase 4 proteins. Additionally, arctigenin was able to <span class="hlt">induce</span> apoptosis in glioma cells, coupled with increased expression levels of cleaved caspase-3 and the pro-apoptotic BCL<span class="hlt">2</span>-associated X protein. Furthermore, arctigenin-<span class="hlt">induced</span> apoptosis was significantly suppressed by the pretreatment of cells with Z-DEVD-FMK, a caspase-3 inhibitor. In conclusion, the results suggest that arctigenin is able to inhibit cell proliferation and may <span class="hlt">induce</span> apoptosis and cell cycle arrest at the <span class="hlt">G</span> 0 /<span class="hlt">G</span> 1 phase in glioma cells. These results warrant further investigation of the anticancer effects of arctigenin in animal models of gliomas.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5351207','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5351207"><span>Arctigenin, a natural lignan compound, <span class="hlt">induces</span> <span class="hlt">G</span>0/<span class="hlt">G</span>1 cell cycle arrest and apoptosis in human glioma cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Maimaitili, Aisha; Shu, Zunhua; Cheng, Xiaojiang; Kaheerman, Kadeer; Sikandeer, Alifu; Li, Weimin</p> <p>2017-01-01</p> <p>The aim of the current study was to investigate the anticancer potential of arctigenin, a natural lignan compound, in malignant gliomas. The U87MG and T98<span class="hlt">G</span> human glioma cell lines were treated with various concentrations of arctigenin for 48 h and the effects of arctigenin on the aggressive phenotypes of glioma cells were assessed. The results demonstrated that arctigenin dose-dependently inhibited the growth of U87MG and T98<span class="hlt">G</span> cells, as determined using 3-(4,5-dimethylthiazol-<span class="hlt">2</span>-yl)-<span class="hlt">2</span>,5-diphenyltetrazolium bromide and bromodeoxyuridine incorporation assays. Arctigenin exposure also <span class="hlt">induced</span> a 60–75% reduction in colony formation compared with vehicle-treated control cells. However, arctigenin was not observed to affect the invasiveness of glioma cells. Arctigenin significantly increased the proportion of cells in the <span class="hlt">G</span>0/<span class="hlt">G</span>1 phase and reduced the number of cells in the S phase, as compared with the control group (P<0.05). Western blot analysis demonstrated that arctigenin increased the expression levels of p21, retinoblastoma and p53 proteins, and significantly decreased the expression levels of cyclin D1 and cyclin-dependent kinase 4 proteins. Additionally, arctigenin was able to <span class="hlt">induce</span> apoptosis in glioma cells, coupled with increased expression levels of cleaved caspase-3 and the pro-apoptotic BCL<span class="hlt">2</span>-associated X protein. Furthermore, arctigenin-<span class="hlt">induced</span> apoptosis was significantly suppressed by the pretreatment of cells with Z-DEVD-FMK, a caspase-3 inhibitor. In conclusion, the results suggest that arctigenin is able to inhibit cell proliferation and may <span class="hlt">induce</span> apoptosis and cell cycle arrest at the <span class="hlt">G</span>0/<span class="hlt">G</span>1 phase in glioma cells. These results warrant further investigation of the anticancer effects of arctigenin in animal models of gliomas. PMID:28356992</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3413253','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3413253"><span>Saponins isolated from Asparagus <span class="hlt">induce</span> apoptosis in human hepatoma cell line Hep<span class="hlt">G</span><span class="hlt">2</span> through a mitochondrial-mediated pathway</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Ji, Y.; Ji, C.; Yue, L.; Xu, H.</p> <p>2012-01-01</p> <p>Objective Many scientific studies have shown that Asparagus officinalis has an antitumour effect and enhances human immunity, but the active components and the antitumour mechanisms are unclear. We investigated the effects of saponins isolated from Asparagus on proliferation and apoptosis in the human hepatoma cell line Hep<span class="hlt">G</span><span class="hlt">2</span>. Methods Hep<span class="hlt">G</span><span class="hlt">2</span> cells were treated with varying concentrations of Asparagus saponins at various times. Using mtt and flow cytometry assays, we evaluated the effects of Asparagus saponins on the growth and apoptosis of Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Transmission electron microscopy was used to observe the morphology of cell apoptosis. Confocal laser scanning microscopy was used to analyze intracellular calcium ion concentration, mitochondrial permeability transition pore (mptp), and mitochondrial membrane potential (mmp). Spectrophotometry was applied to quantify the activity of caspase-9 and caspase-3. Flow cytometry was used to investigate the levels of reactive oxygen species (ros) and pH, and the expressions of Bcl<span class="hlt">2</span>, Bax, CytC, and caspase-3, in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Results Asparagus saponins inhibited the growth of Hep<span class="hlt">G</span><span class="hlt">2</span> cells in a dose-dependent manner. The median inhibitory concentration (IC50) was 101.15 mg/L at 72 hours. The apoptosis morphology at 72 hours of treatment was obvious, showing cell protuberance, concentrated cytoplasm, and apoptotic bodies. The apoptotic rates at 72 hours were 30.9%, 51.7%, and 62.1% (for saponin concentrations of 50 mg/L, 100 mg/L, 200 mg/L). Treatment with Asparagus saponins for 24 hours increased the intracellular level of ros and Ca<span class="hlt">2</span>+, lowered the pH, activated intracellular mptp, and decreased mmp in a dose-dependent manner. Treatment also increased the activity of caspase-9 and caspase-3, downregulated the expression of Bcl<span class="hlt">2</span>, upregulated the expression of Bax, and <span class="hlt">induced</span> release of CytC and activation of caspase-3. Conclusions Asparagus saponins <span class="hlt">induce</span> apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells through a mitochondrial-mediated and caspase</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28713162','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28713162"><span>ABT-263 <span class="hlt">induces</span> <span class="hlt">G</span>1/<span class="hlt">G</span>0-phase arrest, apoptosis and autophagy in human esophageal cancer cells in vitro.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lin, Qing-Huan; Que, Fu-Chang; Gu, Chun-Ping; Zhong, De-Sheng; Zhou, Dan; Kong, Yi; Yu, Le; Liu, Shu-Wen</p> <p>2017-12-01</p> <p>Both the anti- and pro-apoptotic members of the Bcl-<span class="hlt">2</span> family are regulated by a conserved Bcl-<span class="hlt">2</span> homology (BH3) domain. ABT-263 (Navitoclax), a novel BH3 mimetic and orally bioavailable Bcl-<span class="hlt">2</span> family inhibitor with high affinity for Bcl-xL, Bcl-<span class="hlt">2</span> and Bcl-w has entered clinical trials for cancer treatment. But the anticancer mechanisms of ABT-263 have not been fully elucidated. In this study we investigated the effects of ABT-263 on human esophageal cancer cells in vitro and to explore its anticancer mechanisms. Treatment with ABT-263 dose-dependently suppressed the viability of 3 human esophageal cancer cells with IC 50 values of 10.7±1.4, 7.1±1.5 and 8.<span class="hlt">2</span>±1.6 μmol/L, in EC109, HKESC-<span class="hlt">2</span> and CaES-17 cells, respectively. ABT-263 (5-20 μmol/L) dose-dependently <span class="hlt">induced</span> <span class="hlt">G</span> 1 /<span class="hlt">G</span> 0 -phase arrest in the 3 cancer cell lines and <span class="hlt">induced</span> apoptosis evidenced by increased the Annexin V-positive cell population and elevated levels of cleaved caspase 3, cleaved caspase 9 and PARP. We further demonstrated that ABT-263 treatment markedly increased the expression of p21 Waf1/Cip1 and decreased the expression of cyclin D1 and phospho-Rb (retinoblastoma tumor suppressor protein) (Ser780) proteins that contributed to the <span class="hlt">G</span> 1 /<span class="hlt">G</span> 0 -phase arrest. Knockdown of p21 Waf1/Cip1 attenuated ABT-263-<span class="hlt">induced</span> <span class="hlt">G</span> 1 /<span class="hlt">G</span> 0 -phase arrest. Moreover, ABT-263 treatment enhanced pro-survival autophagy, shown as the increased LC3-II levels and decreased p62 levels, which counteracted its anticancer activity. Our results suggest that ABT-263 exerts cytostatic and cytotoxic effects on human esophageal cancer cells in vitro and enhances pro-survival autophagy, which counteracts its anticancer activity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27627923','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27627923"><span>Sulforaphane <span class="hlt">Induces</span> Cell Death Through <span class="hlt">G</span><span class="hlt">2</span>/M Phase Arrest and Triggers Apoptosis in HCT 116 Human Colon Cancer Cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liu, Kuo-Ching; Shih, Ting-Ying; Kuo, Chao-Lin; Ma, Yi-Shih; Yang, Jiun-Long; Wu, Ping-Ping; Huang, Yi-Ping; Lai, Kuang-Chi; Chung, Jing-Gung</p> <p>2016-01-01</p> <p>Sulforaphane (SFN), an isothiocyanate, exists exclusively in cruciferous vegetables, and has been shown to possess potent antitumor and chemopreventive activity. However, there is no available information that shows SFN affecting human colon cancer HCT 116 cells. In the present study, we found that SFN <span class="hlt">induced</span> cell morphological changes, which were photographed by contrast-phase microscopy, and decreased viability. SFN also <span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span>/M phase arrest and cell apoptosis in HCT 116 cells, which were measured with flow cytometric assays. Western blotting indicated that SFN increased Cyclin A, cdk <span class="hlt">2</span>, Cyclin B and WEE1, but decreased Cdc 25C, cdk1 protein expressions that led to <span class="hlt">G</span><span class="hlt">2</span>/M phase arrest. Apoptotic cell death was also confirmed by Annexin V/PI and DAPI staining and DNA gel electrophoresis in HCT 116 cells after exposure to SFN. The flow cytometric assay also showed that SFN <span class="hlt">induced</span> the generation of reactive oxygen species (ROS) and Ca[Formula: see text] and decreased mitochondria membrane potential and increased caspase-8, -9 and -3 activities in HCT 116 cell. Western blotting also showed that SFN <span class="hlt">induced</span> the release of cytochrome c, and AIF, which was confirmed by confocal microscopy examination. SFN <span class="hlt">induced</span> ER stress-associated protein expression. Based on those observations, we suggest that SFN may be used as a novel anticancer agent for the treatment of human colon cancer in the future.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23949848','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23949848"><span>Effect of 3<span class="hlt">G</span> cell phone exposure with computer controlled <span class="hlt">2</span>-D stepper motor on non-thermal activation of the hsp27/p38MAPK stress pathway in rat brain.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kesari, Kavindra Kumar; Meena, Ramovatar; Nirala, Jayprakash; Kumar, Jitender; Verma, H N</p> <p>2014-03-01</p> <p>Cell phone radiation exposure and its biological interaction is the present concern of debate. Present study aimed to investigate the effect of 3<span class="hlt">G</span> cell phone exposure with computer controlled <span class="hlt">2</span>-D stepper motor on 45-day-old male Wistar rat brain. Animals were exposed for <span class="hlt">2</span> h a day for 60 days by using mobile phone with angular movement up to zero to 30°. The variation of the motor is restricted to 90° with respect to the horizontal plane, moving at a pre-determined rate of <span class="hlt">2</span>° per minute. Immediately after 60 days of exposure, animals were scarified and numbers of parameters (DNA double-strand <span class="hlt">break</span>, micronuclei, caspase 3, apoptosis, DNA fragmentation, expression of stress-responsive genes) were performed. Result shows that microwave radiation emitted from 3<span class="hlt">G</span> mobile phone significantly <span class="hlt">induced</span> DNA strand <span class="hlt">breaks</span> in brain. Meanwhile a significant increase in micronuclei, caspase 3 and apoptosis were also observed in exposed group (P < 0.05). Western blotting result shows that 3<span class="hlt">G</span> mobile phone exposure causes a transient increase in phosphorylation of hsp27, hsp70, and p38 mitogen-activated protein kinase (p38MAPK), which leads to mitochondrial dysfunction-mediated cytochrome c release and subsequent activation of caspases, involved in the process of radiation-<span class="hlt">induced</span> apoptotic cell death. Study shows that the oxidative stress is the main factor which activates a variety of cellular signal transduction pathways, among them the hsp27/p38MAPK is the pathway of principle stress response. Results conclude that 3<span class="hlt">G</span> mobile phone radiations affect the brain function and cause several neurological disorders.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22617334','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22617334"><span>Phospho-Bcl-x(L)(Ser62) plays a key role at DNA damage-<span class="hlt">induced</span> <span class="hlt">G</span>(<span class="hlt">2</span>) checkpoint.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Jianfang; Beauchemin, Myriam; Bertrand, Richard</p> <p>2012-06-01</p> <p>Accumulating evidence suggests that Bcl-xL, an anti-apoptotic member of the Bcl-<span class="hlt">2</span> family, also functions in cell cycle progression and cell cycle checkpoints. Analysis of a series of phosphorylation site mutants reveals that cells expressing Bcl-xL(Ser62Ala) mutant are less stable at the <span class="hlt">G</span> <span class="hlt">2</span> checkpoint and enter mitosis more rapidly than cells expressing wild-type Bcl-xL or Bcl-xL phosphorylation site mutants, including Thr41Ala, Ser43Ala, Thr47Ala, Ser56Ala and Thr115Ala. Analysis of the dynamic phosphorylation and location of phospho-Bcl-xL(Ser62) in unperturbed, synchronized cells and during DNA damage-<span class="hlt">induced</span> <span class="hlt">G</span> <span class="hlt">2</span> arrest discloses that a pool of phospho-Bcl-xL(Ser62) accumulates into nucleolar structures in etoposide-exposed cells during <span class="hlt">G</span> <span class="hlt">2</span> arrest. In a series of in vitro kinase assays, pharmacological inhibitors and specific siRNAs experiments, we found that Polo kinase 1 and MAPK9/JNK<span class="hlt">2</span> are major protein kinases involved in Bcl-xL(Ser62) phosphorylation and accumulation into nucleolar structures during the <span class="hlt">G</span> <span class="hlt">2</span> checkpoint. In nucleoli, phospho-Bcl-xL(Ser62) binds to and co-localizes with Cdk1(cdc<span class="hlt">2</span>), the key cyclin-dependent kinase required for entry into mitosis. These data indicate that during <span class="hlt">G</span> <span class="hlt">2</span> checkpoint, phospho-Bcl-xL(Ser62) stabilizes <span class="hlt">G</span> <span class="hlt">2</span> arrest by timely trapping of Cdk1(cdc<span class="hlt">2</span>) in nucleolar structures to slow mitotic entry. It also highlights that DNA damage affects the dynamic composition of the nucleolus, which now emerges as a piece of the DNA damage response.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.6076B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.6076B"><span>Different modes of continental <span class="hlt">break</span>-up triggered by a sole mantle plume: a <span class="hlt">2</span>D and 3D numerical study</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Beniest, Anouk; Koptev, Alexander; Leroy, Sylvie; Burov, Evgueni</p> <p>2017-04-01</p> <p>We used <span class="hlt">2</span>D and 3D numerical models to investigate the impact of a single mantle plume on continental rifting and breakup processes. We varied the thermo-rheological structure of the continental lithosphere, its geometry and the initial plume position. Based on the results of our <span class="hlt">2</span>D experiments, three continental <span class="hlt">break</span>-up modes can be distinguished: A) 'central' continental <span class="hlt">break</span>-up, the <span class="hlt">break</span>-up center is located directly above the original mantle anomaly position, B) 'shifted' <span class="hlt">break</span>-up, the <span class="hlt">break</span>-up center is 50 to 200 km displaced from the initial plume location and C) 'distant' <span class="hlt">break</span>-up, due to convection and/or slab-subduction/delamination, the <span class="hlt">break</span>-up center is considerably shifted (300 to 800 km) from the primary plume position. Our 3D model, with a laterally homogeneous initial setup also results in continental <span class="hlt">break</span>-up with the axis of continental <span class="hlt">break</span>-up hundreds of kilometers shifted from the original plume location. The model results show that the classical, 'central' view of mantle plume <span class="hlt">induced</span> continental <span class="hlt">break</span>-up is not the only mode of <span class="hlt">break</span>-up. When considering a diversity of <span class="hlt">break</span>-up styles, it is possible to explain a variety of observed geophysical and geological features. For example, the mantle material glued to the base of the lithosphere at shallower depths corresponds geometrically and location-wise to high-velocity/high-density bodies observed on seismic data below the thinned continental lithosphere and the transition zone of the South Atlantic domain. During migration, products of partial melting of the mantle material can move vertically to (shallow) lower crustal levels. They might resemble high density bodies observed at lower crustal levels inside continental crust with similar geometries observed with gravity modelling. Also, topographic variation form in the very early stages of rifting on the first impingement of upwelled plume material. These variations remain visible, as the final position of the spreading center is shifted</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/10812246','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/10812246"><span>Immunostimulatory Cp<span class="hlt">G</span>-oligonucleotides <span class="hlt">induce</span> functional high affinity IL-<span class="hlt">2</span> receptors on B-CLL cells: costimulation with IL-<span class="hlt">2</span> results in a highly immunogenic phenotype.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Decker, T; Schneller, F; Kronschnabl, M; Dechow, T; Lipford, G B; Wagner, H; Peschel, C</p> <p>2000-05-01</p> <p>Cp<span class="hlt">G</span>-oligodeoxynucleotides (Cp<span class="hlt">G</span>-ODN) have been shown to <span class="hlt">induce</span> proliferation, cytokine production, and surface molecule regulation in normal and malignant human B cells. In the present study, we investigated the potential of Cp<span class="hlt">G</span>-ODN to <span class="hlt">induce</span> functional high-affinity receptors in leukemic and normal B cells and the effects of costimulation with IL-<span class="hlt">2</span> on proliferation, cytokine secretion, and surface molecule regulation. Highly purified B cells from B-CLL patients and normal controls were stimulated with Cp<span class="hlt">G</span>-ODN with or without IL-<span class="hlt">2</span>. Expression of CD25 was determined using FACS, and the presence of high-affinity IL-<span class="hlt">2</span> receptors was determined by scatchard analysis. Costimulatory effects of IL-<span class="hlt">2</span> and Cp<span class="hlt">G</span>-ODN were investigated using proliferation assays, ELISA (IL-6, TNF-alpha), and FACS analysis (CD80, CD86 expression). Reactivity of autologous and allogeneic T cells toward activated B-CLL cells was determined in mixed lymphocyte reactions and Interferon-gamma Elispot assays. The Cp<span class="hlt">G</span>-ODN DSP30 caused a significantly stronger induction of the IL-<span class="hlt">2</span> receptor alpha chain in malignant as compared with normal B cells (p = 0.03). This resulted in the expression of functional high-affinity IL-<span class="hlt">2</span> receptors in B-CLL cells, but fewer numbers of receptors with less affinity were expressed in normal B cells. Although addition of IL-<span class="hlt">2</span> to Cp<span class="hlt">G</span>-ODN-stimulated cells augmented proliferation in both normal B cells and B-CLL cells, no costimulatory effect on cytokine production or surface molecule expression could be observed in normal B cells. In contrast, TNF-alpha and IL-6 production was increased in B-CLL cells, and the expression of CD80 and CD86 was further enhanced when IL-<span class="hlt">2</span> was used as a costimulus. Autologous and allogeneic immune recognition of B-CLL cells stimulated with Cp<span class="hlt">G</span>-ODN and IL-<span class="hlt">2</span> was increased compared with B-CLL cells stimulated with Cp<span class="hlt">G</span>-ODN alone. Stimulation of B-CLL cells with Cp<span class="hlt">G</span>-ODN and IL-<span class="hlt">2</span> might be an attractive strategy for potential immunotherapies for B</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=39928&Lab=ORD&keyword=infusion&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50','EPA-EIMS'); return false;" href="https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=39928&Lab=ORD&keyword=infusion&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50"><span>INDUCTION, ACCUMULATION, AND PERSISTENCE OF SISTER <span class="hlt">CHROMATID</span> EXCHANGES IN WOMEN WITH BREAST CANCER RECEIVING CYCLOPHOSPHAMIDE, ANDRIAMYCIN, AND 5-FLUOROACIL CHEMOTHERAPY</span></a></p> <p><a target="_blank" href="http://oaspub.epa.gov/eims/query.page">EPA Science Inventory</a></p> <p></p> <p></p> <p>The induction, stimulation, and persistence of sister <span class="hlt">chromatid</span> exchanges (SCE's) and high SCE frequency cells (HFC's) was measured in peripheral lymphocytes of women with breast cancer before chemotherapy and on multiple occasions during and after therapy. Chemotherapy consisted...</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li class="active"><span>17</span></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_17 --> <div id="page_18" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li class="active"><span>18</span></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="341"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015OcMod..87...30S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015OcMod..87...30S"><span>Scaling depth-<span class="hlt">induced</span> wave-<span class="hlt">breaking</span> in two-dimensional spectral wave models</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Salmon, J. E.; Holthuijsen, L. H.; Zijlema, M.; van Vledder, G. Ph.; Pietrzak, J. D.</p> <p>2015-03-01</p> <p>Wave <span class="hlt">breaking</span> in shallow water is still poorly understood and needs to be better parameterized in <span class="hlt">2</span>D spectral wave models. Significant wave heights over horizontal bathymetries are typically under-predicted in locally generated wave conditions and over-predicted in non-locally generated conditions. A joint scaling dependent on both local bottom slope and normalized wave number is presented and is shown to resolve these issues. Compared to the 12 wave <span class="hlt">breaking</span> parameterizations considered in this study, this joint scaling demonstrates significant improvements, up to ∼50% error reduction, over 1D horizontal bathymetries for both locally and non-locally generated waves. In order to account for the inherent differences between uni-directional (1D) and directionally spread (<span class="hlt">2</span>D) wave conditions, an extension of the wave <span class="hlt">breaking</span> dissipation models is presented. By including the effects of wave directionality, rms-errors for the significant wave height are reduced for the best performing parameterizations in conditions with strong directional spreading. With this extension, our joint scaling improves modeling skill for significant wave heights over a verification data set of 11 different 1D laboratory bathymetries, 3 shallow lakes and 4 coastal sites. The corresponding averaged normalized rms-error for significant wave height in the <span class="hlt">2</span>D cases varied between 8% and 27%. In comparison, using the default setting with a constant scaling, as used in most presently operating <span class="hlt">2</span>D spectral wave models, gave equivalent errors between 15% and 38%.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003PhRvC..68e4304M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003PhRvC..68e4304M"><span>Systematics of first <span class="hlt">2</span>+ state <span class="hlt">g</span> factors around mass 80</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mertzimekis, T. J.; Stuchbery, A. E.; Benczer-Koller, N.; Taylor, M. J.</p> <p>2003-11-01</p> <p>The systematics of the first <span class="hlt">2</span>+ state <span class="hlt">g</span> factors in the mass 80 region are investigated in terms of an IBM-II analysis, a pairing-corrected geometrical model, and a shell-model approach. Subshell closure effects at N=38 and overall trends were examined using IBM-II. A large-space shell-model calculation was successful in describing the behavior for N=48 and N=50 nuclei, where single-particle features are prominent. A schematic truncated-space calculation was applied to the lighter isotopes. The variations of the effective boson <span class="hlt">g</span> factors are discussed in connection with the role of F -spin <span class="hlt">breaking</span>, and comparisons are made between the mass 80 and mass 180 regions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1572797','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1572797"><span>Endothelium-dependent relaxation <span class="hlt">induced</span> by cathepsin <span class="hlt">G</span> in porcine pulmonary arteries</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Glusa, Erika; Adam, Christine</p> <p>2001-01-01</p> <p>Serine proteinases elicit profound cellular effects in various tissues mediated by activation of proteinase-activated receptors (PAR). In the present study, we investigated the vascular effects of cathepsin <span class="hlt">G</span>, a serine proteinase that is present in the azurophil granules of leukocytes and is known to activate several cells that express PARs. In prostaglandin F<span class="hlt">2</span>α (3 μM)-precontracted rings from porcine pulmonary arteries with intact endothelium, cathepsin <span class="hlt">G</span> caused concentration-dependent relaxant responses (pEC50=9.64±0.12). The endothelium-dependent relaxant effect of cathepsin <span class="hlt">G</span> could also be demonstrated in porcine coronary arteries (pEC50=9.23±0.07). In pulmonary arteries the cathepsin <span class="hlt">G-induced</span> relaxation was inhibited after blockade of nitric oxide synthesis by L-NAME (200 μM) and was absent in endothelium-denuded vessels. Bradykinin- and cathepsin <span class="hlt">G-induced</span> relaxant effects were associated with a 5.7 fold and <span class="hlt">2</span>.4 fold increase in the concentration of cyclic GMP, respectively. Compared with thrombin and trypsin, which also produced an endothelium-dependent relaxation in pulmonary arteries, cathepsin <span class="hlt">G</span> was <span class="hlt">2</span>.5 and four times more potent, respectively. Cathepsin <span class="hlt">G</span> caused only small homologous desensitization. In cathepsin <span class="hlt">G</span>-challenged vessels, thrombin was still able to elicit a relaxant effect. The effects of cathepsin <span class="hlt">G</span> were blocked by soybean trypsin inhibitor (IC50=0.043 μg ml−1), suggesting that proteolytic activity is essential for induction of relaxation. Recombinant acetyl-eglin C proved to be a potent inhibitor (IC50=0.14 μg ml−1) of the cathepsin <span class="hlt">G</span> effect, whereas neither indomethacin (3 μM) nor the thrombin inhibitor hirudin (5 ATU ml−1) elicited any inhibitory activity. Due to their polyanionic structure defibrotide (IC50=0.11 μg ml−1), heparin (IC50=0.48 μg ml−1) and suramin (IC50=1.85 μg ml−1) diminished significantly the relaxation in response to the basic protein cathepsin <span class="hlt">G</span>. In conclusion, like</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1976452','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1976452"><span><span class="hlt">G</span>-quadruplex <span class="hlt">induced</span> stabilization by <span class="hlt">2</span>′-deoxy-<span class="hlt">2</span>′-fluoro-d-arabinonucleic acids (<span class="hlt">2</span>′F-ANA)</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Peng, Chang Geng; Damha, Masad J.</p> <p>2007-01-01</p> <p>The impact of <span class="hlt">2</span>′-deoxy-<span class="hlt">2</span>′-fluoroarabinonucleotide residues (<span class="hlt">2</span>′F-araN) on different <span class="hlt">G</span>-quadruplexes derived from a thrombin-binding DNA aptamer d(<span class="hlt">G</span><span class="hlt">2</span>T<span class="hlt">2</span><span class="hlt">G</span><span class="hlt">2</span>TGTG<span class="hlt">2</span>T<span class="hlt">2</span><span class="hlt">G</span><span class="hlt">2</span>), an anti-HIV phosphorothioate aptamer PS-d(T<span class="hlt">2</span><span class="hlt">G</span>4T<span class="hlt">2</span>) and a DNA telomeric sequence d(<span class="hlt">G</span>4T4<span class="hlt">G</span>4) via UV thermal melting (Tm) and circular dichroism (CD) experiments has been investigated. Generally, replacement of deoxyguanosines that adopt the anti conformation (anti-guanines) with <span class="hlt">2</span>′F-ara<span class="hlt">G</span> can stabilize <span class="hlt">G</span>-quartets and maintain the quadruplex conformation, while replacement of syn-guanines with <span class="hlt">2</span>′F-ara<span class="hlt">G</span> is not favored and results in a dramatic switch to an alternative quadruplex conformation. It was found that incorporation of <span class="hlt">2</span>′F-ara<span class="hlt">G</span> or T residues into a thrombin-binding DNA <span class="hlt">G</span>-quadruplex stabilizes the complex (ΔTm up to ∼+3°C/<span class="hlt">2</span>′F-araN modification); <span class="hlt">2</span>′F-araN units also increased the half-life in 10% fetal bovine serum (FBS) up to 48-fold. Two modified thrombin-binding aptamers (PG13 and PG14) show an approximately 4-fold increase in binding affinity to thrombin, as assessed via a nitrocellulose filter binding assay, both with increased thermal stability (∼1°C/<span class="hlt">2</span>′F-ANA modification increase in Tm) and nuclease resistance (4–7-fold) as well. Therefore, the <span class="hlt">2</span>′-deoxy-<span class="hlt">2</span>′-fluoro-d-arabinonucleic acid (<span class="hlt">2</span>′F-ANA) modification is well suited to tune (and improve) the physicochemical and biological properties of naturally occurring DNA <span class="hlt">G</span>-quartets. PMID:17636049</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040173255&hterms=genetic+fish&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dgenetic%2Bfish','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040173255&hterms=genetic+fish&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dgenetic%2Bfish"><span>Rejoining and misrejoining of radiation-<span class="hlt">induced</span> chromatin <span class="hlt">breaks</span>. II. Biophysical Model</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Wu, H.; Durante, M.; George, K.; Goodwin, E. H.; Yang, T. C.</p> <p>1996-01-01</p> <p>A biophysical model for the kinetics of the formation of radiation-<span class="hlt">induced</span> chromosome aberrations is developed to account for the recent experimental results obtained with a combination of the premature chromosome condensation (PCC) and fluorescence in situ hybridization (FISH) techniques. In this model, we consider the broken ends of DNA double-strand <span class="hlt">breaks</span> (DSBs) to be reactant and make use of the interaction distance hypothesis. The repair/misrepair process between broken ends is suggested to consist of two steps; the first step represents the two <span class="hlt">break</span> ends approaching each other, and the second step represents the enzymatic processes leading to DNA end-to-end rejoining. Only the second step is reflected in the kinetics observed in experiments using PCC. The model appears to be able to fit existing data for human cells. It is shown that the kinetics of the formation of chromosome aberrations can be explained by a single rate that characterizes both rejoining and misrejoining of DSBs, suggesting that repair and misrepair share the same mechanism. Fast repair (completed in minutes) in a subset of DSBs is suggested as an explanation of the complete exchanges observed with PCC in human lymphocytes immediately after irradiation. The fast repair component seems to be absent in human fibroblasts.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3935595','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3935595"><span>Lithium Causes <span class="hlt">G</span><span class="hlt">2</span> Arrest of Renal Principal Cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>de Groot, Theun; Alsady, Mohammad; Jaklofsky, Marcel; Otte-Höller, Irene; Baumgarten, Ruben; Giles, Rachel H.</p> <p>2014-01-01</p> <p>Vasopressin-regulated expression and insertion of aquaporin-<span class="hlt">2</span> channels in the luminal membrane of renal principal cells is essential for urine concentration. Lithium affects urine concentrating ability, and approximately 20% of patients treated with lithium develop nephrogenic diabetes insipidus (NDI), a disorder characterized by polyuria and polydipsia. Lithium-<span class="hlt">induced</span> NDI is caused by aquaporin-<span class="hlt">2</span> downregulation and a reduced ratio of principal/intercalated cells, yet lithium <span class="hlt">induces</span> principal cell proliferation. Here, we studied how lithium-<span class="hlt">induced</span> principal cell proliferation can lead to a reduced ratio of principal/intercalated cells using two-dimensional and three-dimensional polarized cultures of mouse renal collecting duct cells and mice treated with clinically relevant lithium concentrations. DNA image cytometry and immunoblotting revealed that lithium initiated proliferation of mouse renal collecting duct cells but also increased the <span class="hlt">G</span><span class="hlt">2</span>/S ratio, indicating <span class="hlt">G</span><span class="hlt">2</span>/M phase arrest. In mice, treatment with lithium for 4, 7, 10, or 13 days led to features of NDI and an increase in the number of principal cells expressing PCNA in the papilla. Remarkably, 30%–40% of the PCNA-positive principal cells also expressed pHistone-H3, a late <span class="hlt">G</span><span class="hlt">2</span>/M phase marker detected in approximately 20% of cells during undisturbed proliferation. Our data reveal that lithium treatment initiates proliferation of renal principal cells but that a significant percentage of these cells are arrested in the late <span class="hlt">G</span><span class="hlt">2</span> phase, which explains the reduced principal/intercalated cell ratio and may identify the molecular pathway underlying the development of lithium-<span class="hlt">induced</span> renal fibrosis. PMID:24408872</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28257055','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28257055"><span>A Novel Polysaccharide Conjugate from Bullacta exarata <span class="hlt">Induces</span> <span class="hlt">G</span>1-Phase Arrest and Apoptosis in Human Hepatocellular Carcinoma Hep<span class="hlt">G</span><span class="hlt">2</span> Cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liao, Ningbo; Sun, Liang; Chen, Jiang; Zhong, Jianjun; Zhang, Yanjun; Zhang, Ronghua</p> <p>2017-03-01</p> <p>Bullacta exarata has been consumed in Asia, not only as a part of the normal diet, but also as a traditional Chinese medicine with liver- and kidney-benefitting functions. Several scientific investigations involving extraction of biomolecules from this mollusk and pharmacological studies on their biological activities have been carried out. However, little is known regarding the antitumor properties of polysaccharides from B. exarata , hence the polysaccharides from B. exarata have been investigated here. One polysaccharide conjugate BEPS-IA was isolated and purified from B. exarata . It mainly consisted of mannose and glucose in a molar ratio of 1:<span class="hlt">2</span>, with an average molecular weight of 127 kDa. Thirteen general amino acids were identified to be components of the protein-bound polysaccharide. Methylation and NMR studies revealed that BEPS-IA is a heteropolysaccharide consisting of 1,4-linked-α-d-Glc, 1,6-linked-α-d-Man, 1,3,6-linked-α-d-Man, and 1-linked-α-d-Man residue, in a molar ratio of 6:1:1:1. In order to test the antitumor activity of BEPS-IA, we investigated its effect against the growth of human hepatocellular carcinoma cells Hep<span class="hlt">G</span><span class="hlt">2</span> in vitro. The result showed that BEPS-IA dose-dependently exhibited an effective Hep<span class="hlt">G</span><span class="hlt">2</span> cells growth inhibition with an IC 50 of 112.4 μ<span class="hlt">g</span>/mL. Flow cytometry analysis showed that BEPS-IA increased the populations of both apoptotic sub-<span class="hlt">G</span>1 and <span class="hlt">G</span>1 phase. The result obtained from TUNEL assay corroborated apoptosis which was shown in flow cytometry. Western blot analysis suggested that BEPS-IA <span class="hlt">induced</span> apoptosis and growth inhibition were associated with up-regulation of p53, p21 and Bax, down-regulation of Bcl-<span class="hlt">2</span>. These findings suggest that BEPS-IA may serve as a potential novel dietary agent for hepatocellular carcinoma.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/978252-telomere-sister-chromatid-exchange-telomerase-deficient-murine-cells','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/978252-telomere-sister-chromatid-exchange-telomerase-deficient-murine-cells"><span>Telomere sister <span class="hlt">chromatid</span> exchange in telomerase deficient murine cells</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Wang, Yisong; Giannone, Richard J; Liu, Yie</p> <p>2005-01-01</p> <p>We have recently demonstrated that several types of genomic rearrangements (i.e., telomere sister <span class="hlt">chromatid</span> exchange (T-SCE), genomic-SCE, or end-to-end fusions) were more often detected in long-term cultured murine telomerase deficient embryonic stem (ES) cells than in freshly prepared murine splenocytes, even through they possessed similar frequencies of critically short telomeres. The high rate of genomic rearrangements in telomerase deficient ES cells, when compared to murine splenocytes, may reflect the cultured cells' gained ability to protect chromosome ends with eroded telomeres allowing them to escape 'end crisis'. However, the possibility that ES cells were more permissive to genomic rearrangements than othermore » cell types or that differences in the microenvironment or genetic background of the animals might consequentially determine the rate of T-SCEs or other genomic rearrangements at critically short telomeres could not be ruled out.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2893262','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2893262"><span>Influence of homologous recombinational repair on cell survival and chromosomal aberration induction during the cell cycle in γ-irradiated CHO cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Wilson, Paul F.; Hinz, John M.; Urbin, Salustra S.; Nham, Peter B.; Thompson, Larry H.</p> <p>2010-01-01</p> <p>The repair of DNA double-strand <span class="hlt">breaks</span> (DSB) by homologous recombinational repair (HRR) underlies the high radioresistance and low mutability observed in S-phase mammalian cells. To evaluate the contributions of HRR and nonhomologous end-joining (NHEJ) to overall DSB repair capacity throughout the cell cycle after γ-irradiation, we compared HRR-deficient RAD51D-knockout 51D1 to CgRAD51D-complemented 51D1 (51D1.3) CHO cells for survival and chromosomal aberrations (CAs). Asynchronous cultures were irradiated with 150 or 300 cGy and separated by cell size using centrifugal elutriation. Cell survival of each synchronous fraction (~20 fractions total from early <span class="hlt">G</span>1 to late <span class="hlt">G</span><span class="hlt">2</span>/M) was measured by colony formation. 51D1.3 cells were most resistant in S, while 51D1 cells were most resistant in early <span class="hlt">G</span>1 (with survival and chromosome-type CA levels similar to 51D1.3) and became progressively more sensitive throughout S and <span class="hlt">G</span><span class="hlt">2</span>. Both cell lines experienced significantly reduced survival from late S into <span class="hlt">G</span><span class="hlt">2</span>. Metaphases were collected from every third elutriation fraction at the first post-irradiation mitosis and scored for CAs. 51D1 cells irradiated in S and <span class="hlt">G</span><span class="hlt">2</span> had ~<span class="hlt">2</span>-fold higher <span class="hlt">chromatid</span>-type CAs and a remarkable ~25-fold higher level of complex <span class="hlt">chromatid</span>-type exchanges compared to 51D1.3 cells. Complex exchanges in 51D1.3 cells were only observed in <span class="hlt">G</span><span class="hlt">2</span>. These results show an essential role for HRR in preventing gross chromosomal rearrangements in proliferating cells and, with our previous report of reduced survival of <span class="hlt">G</span><span class="hlt">2</span>-phase NHEJ-deficient prkdc CHO cells [Hinz et al. DNA Repair 4, 782–792, 2005], imply reduced activity/efficiency of both HRR and NHEJ as cells transition from S to <span class="hlt">G</span><span class="hlt">2</span>. PMID:20434408</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11098853','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11098853"><span>Radiation-<span class="hlt">induced</span> double-strand <span class="hlt">breaks</span> in mammalian DNA: influence of temperature and DMSO.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Elmroth, K; Nygren, J; Erkell, L J; Hultborn, R</p> <p>2000-11-01</p> <p>To investigate the effects of subphysiological irradiation temperature (<span class="hlt">2</span> 28 degrees C) and the influence of the radical scavenger DMSO on the induction of double-strand <span class="hlt">breaks</span> (DSB) in chromosomal DNA from a human breast cancer cell line (MCF-7) as well as in intact cells. The rejoining of DSB in cells irradiated at <span class="hlt">2</span> degrees C or 37 degrees C was also investigated. Agarose plugs with [14C]thymidine labelled MCF-7 cells were lysed in EDTA-NLS-proteinase-K buffer. The plugs containing chromosomal DNA were irradiated with X-rays under different temperatures and scavenging conditions. Intact MCF-7 cells were irradiated in Petri dishes and plugs were made. The cells were then lysed in EDTA-NLS-proteinase-K buffer. The induction of DSB was studied by constant field gel electrophoresis and expressed as DSB/100/Mbp, calculated from the fraction of activity released into the gel. The induction of DSB in chromosomal DNA was reduced by a decrease in temperature. This protective effect of low temperature was inhibited when the DNA was irradiated in the presence of DMSO. No difference was found when intact cells were irradiated at different temperatures. However, the rapid phase of rejoining was slower in cells irradiated at 37 degrees C than at <span class="hlt">2</span> degrees C. The induction of DSB in naked DNA was reduced by hypothermic irradiation. The temperature had no influence on the induction of DSB in the presence of a high concentration of DMSO, indicating that the temperature effect is mediated via the indirect effects of ionizing radiation. Results are difficult to interpret in intact cells. Rejoining during irradiation at the higher temperature may counteract an increased induction. The difference in rejoining may be interpreted in terms of qualitative differences between <span class="hlt">breaks</span> <span class="hlt">induced</span> at the two temperatures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21554918','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21554918"><span>Eupatilin, a dietary flavonoid, <span class="hlt">induces</span> <span class="hlt">G</span><span class="hlt">2</span>/M cell cycle arrest in human endometrial cancer cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cho, Jung-Hoon; Lee, Jong-Gyu; Yang, Yeong-In; Kim, Ji-Hyun; Ahn, Ji-Hye; Baek, Nam-In; Lee, Kyung-Tae; Choi, Jung-Hye</p> <p>2011-08-01</p> <p>This study is the first to investigate the antiproliferative effect of eupatilin in human endometrial cancer cells. Eupatilin, a naturally occurring flavonoid isolated from Artemisia princeps, has anti-inflammatory, anti-oxidative, and anti-tumor activities. In the present study, we investigated the potential effect of eupatilin on cell growth and its molecular mechanism of action in human endometrial cancer cells. Eupatilin was more potent than cisplatin in inhibiting cell viability in the human endometrial cancer cell lines Hec1A and KLE. Eupatilin showed relatively low cytotoxicity in normal human endometrial cells HES and HESC cells when compared to cisplatin. Eupatilin <span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span>/M phase cell cycle arrest in a time- and dose-dependent manner, as indicated by flow cytometry analysis. In addition, treatment of Hec1A cells with eupatilin resulted in a significant increase in the expression of p21(WAF1/CIP1) and in the phosphorylation of Cdc25C and Cdc<span class="hlt">2</span>. Knockdown of p21 using specific siRNAs significantly compromised eupatilin-<span class="hlt">induced</span> cell growth inhibition. Interestingly, levels of mutant p53 in Hec1A cells decreased markedly upon treatment with eupatilin, and p53 siRNA significantly increased p21 expression. Moreover, eupatilin modulated the phosphorylation of protein kinases ERK1/<span class="hlt">2</span>, Akt, ATM, and Chk<span class="hlt">2</span>. These results suggest that eupatilin inhibits the growth of human endometrial cancer cells via <span class="hlt">G</span><span class="hlt">2</span>/M phase cell cycle arrest through the up-regulation of p21 by the inhibition of mutant p53 and the activation of the ATM/Chk<span class="hlt">2</span>/Cdc25C/Cdc<span class="hlt">2</span> checkpoint pathway. Copyright © 2011 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16296884','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16296884"><span>DNA strand <span class="hlt">breaks</span> signal the induction of DNA double-strand <span class="hlt">break</span> repair in Saccharomyces cerevisiae.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Singh, Rakesh Kumar; Krishna, Malini</p> <p>2005-12-01</p> <p>Genotoxic stress <span class="hlt">induces</span> a checkpoint signaling cascade to generate a stress response. Saccharomyces cerevisiae shows an altered radiation response under different type of stress. Although the induction of repair has been implicated in enhanced survival after exposure to the challenging stress, the nature of the signal remains poorly understood. This study demonstrates that low doses of gamma radiation and bleomycin <span class="hlt">induce</span> RAD52-dependent recombination repair pathway in the wild-type strain D-261. Prior exposure of cells to DNA-damaging agents (gamma radiation or bleomycin) equips them better for the subsequent damage caused by challenging doses. However, exposure to UV light, which does not cause strand <span class="hlt">breaks</span>, was ineffective. This was confirmed by PFGE studies. This indicates that the strand <span class="hlt">breaks</span> probably serve as the signal for induction of the recombination repair pathway while pyrimidine dimers do not. The nature of the <span class="hlt">induced</span> repair was investigated by mutation scoring in special strain D-7, which showed that the <span class="hlt">induced</span> repair is essentially error free.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5371300','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5371300"><span>Heat-<span class="hlt">induced</span> symmetry <span class="hlt">breaking</span> in ant (Hymenoptera: Formicidae) escape behavior</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Chung, Yuan-Kai</p> <p>2017-01-01</p> <p>The collective egress of social insects is important in dangerous situations such as natural disasters or enemy attacks. Some studies have described the phenomenon of symmetry <span class="hlt">breaking</span> in ants, with two exits <span class="hlt">induced</span> by a repellent. However, whether symmetry <span class="hlt">breaking</span> occurs under high temperature conditions, which are a common abiotic stress, remains unknown. In our study, we deposited a group of Polyrhachis dives ants on a heated platform and counted the number of escaping ants with two identical exits. We discovered that ants asymmetrically escaped through two exits when the temperature of the heated platform was >32.75°C. The degree of asymmetry increased linearly with the temperature of the platform. Furthermore, the higher the temperature of heated platform was, the more ants escaped from the heated platform. However, the number of escaping ants decreased for 3 min when the temperature was higher than the critical thermal limit (39.46°C), which is the threshold for ants to endure high temperature without a loss of performance. Moreover, the ants tended to form small groups to escape from the thermal stress. A preparatory formation of ant grouping was observed before they reached the exit, indicating that the ants actively clustered rather than accidentally gathered at the exits to escape. We suggest that a combination of individual and grouping ants may help to optimize the likelihood of survival during evacuation. PMID:28355235</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24013230','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24013230"><span>The MutSβ complex is a modulator of p53-driven tumorigenesis through its functions in both DNA double-strand <span class="hlt">break</span> repair and mismatch repair.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>van Oers, J M M; Edwards, Y; Chahwan, R; Zhang, W; Smith, C; Pechuan, X; Schaetzlein, S; Jin, B; Wang, Y; Bergman, A; Scharff, M D; Edelmann, W</p> <p>2014-07-24</p> <p>Loss of the DNA mismatch repair (MMR) protein MSH3 leads to the development of a variety of tumors in mice without significantly affecting survival rates, suggesting a modulating role for the MutSβ (MSH<span class="hlt">2</span>-MSH3) complex in late-onset tumorigenesis. To better study the role of MSH3 in tumor progression, we crossed Msh3(-/-) mice onto a tumor predisposing p53-deficient background. Survival of Msh3/p53 mice was not reduced compared with p53 single mutant mice; however, the tumor spectrum changed significantly from lymphoma to sarcoma, indicating MSH3 as a potent modulator of p53-driven tumorigenesis. Interestingly, Msh3(-/-) mouse embryonic fibroblasts displayed increased <span class="hlt">chromatid</span> <span class="hlt">breaks</span> and persistence of γH<span class="hlt">2</span>AX foci following ionizing radiation, indicating a defect in DNA double-strand <span class="hlt">break</span> repair (DSBR). Msh3/p53 tumors showed increased loss of heterozygosity, elevated genome-wide copy-number variation and a moderate microsatellite instability phenotype compared with Msh<span class="hlt">2</span>/p53 tumors, revealing that MSH<span class="hlt">2</span>-MSH3 suppresses tumorigenesis by maintaining chromosomal stability. Our results show that the MSH<span class="hlt">2</span>-MSH3 complex is important for the suppression of late-onset tumors due to its roles in DNA DSBR as well as in DNA MMR. Further, they demonstrate that MSH<span class="hlt">2</span>-MSH3 suppresses chromosomal instability and modulates the tumor spectrum in p53-deficient tumorigenesis and possibly has a role in other chromosomally unstable tumors as well.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4507860','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4507860"><span>Disorder of <span class="hlt">G</span><span class="hlt">2</span>-M Checkpoint Control in Aniline-<span class="hlt">Induced</span> Cell Proliferation in Rat Spleen</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Wang, Jianling; Wang, Gangduo; Khan, M. Firoze</p> <p>2015-01-01</p> <p>Aniline, a toxic aromatic amine, is known to cause hemopoietic toxicity both in humans and animals. Aniline exposure also leads to toxic response in spleen which is characterized by splenomegaly, hyperplasia, fibrosis and the eventual formation of tumors on chronic in vivo exposure. Previously, we have shown that aniline exposure leads to iron overload, oxidative DNA damage, and increased cell proliferation, which could eventually contribute to a tumorigenic response in the spleen. Despite our demonstration that cell proliferation was associated with deregulation of <span class="hlt">G</span>1 phase cyclins and increased expression of <span class="hlt">G</span>1 phase cyclin-dependent kinases (CDKs), molecular mechanisms, especially the regulation of <span class="hlt">G</span><span class="hlt">2</span> phase and contribution of epigenetic mechanisms in aniline-<span class="hlt">induced</span> splenic cellular proliferation remain largely unclear. This study therefore, mainly focused on the regulation of <span class="hlt">G</span><span class="hlt">2</span> phase in an animal model preceding a tumorigenic response. Male Sprague-Dawley rats were given aniline (0.5 mmol/kg/day) in drinking water or drinking water only (controls) for 30 days, and expression of <span class="hlt">G</span><span class="hlt">2</span> phase cyclins, CDK1, CDK inhibitors and miRNAs were measured in the spleen. Aniline treatment resulted in significant increases in cell cycle regulatory proteins, including cyclins A, B and CDK1, particularly phosphor-CDK1, and decreases in CDK inhibitors p21 and p27, which could promote the splenocytes to go through <span class="hlt">G</span><span class="hlt">2</span>/M transition. Our data also showed upregulation of tumor markers Trx-1 and Ref-1 in rats treated with aniline. More importantly, we observed lower expression of miRNAs including Let-7a, miR-15b, miR24, miR-100 and miR-125, and greater expression of CDK inhibitor regulatory miRNAs such as miR-181a, miR-221 and miR-222 in the spleens of aniline-treated animals. Our findings suggest that significant increases in the expression of cyclins, CDK1 and aberrant regulation of miRNAs could lead to an accelerated <span class="hlt">G</span><span class="hlt">2</span>/M transition of the splenocytes, and potentially to a</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28118114','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28118114"><span>Coordination of the Ser2056 and Thr2609 Clusters of DNA-PKcs in Regulating Gamma Rays and Extremely Low Fluencies of Alpha-Particle Irradiation to <span class="hlt">G</span>0/<span class="hlt">G</span>1 Phase Cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nagasawa, Hatsumi; Lin, Yu-Fen; Kato, Takamitsu A; Brogan, John R; Shih, Hung-Ying; Kurimasa, Akihiro; Bedford, Joel S; Chen, Benjamin P C; Little, John B</p> <p>2017-02-01</p> <p>The catalytic subunit of DNA dependent protein kinase (DNA-PKcs) and its kinase activity are critical for mediation of non-homologous end-joining (NHEJ) of DNA double-strand <span class="hlt">breaks</span> (DSB) in mammalian cells after gamma-ray irradiation. Additionally, DNA-PKcs phosphorylations at the T2609 cluster and the S2056 cluster also affect DSB repair and cellular sensitivity to gamma radiation. Previously we reported that phosphorylations within these two regions affect not only NHEJ but also homologous recombination repair (HRR) dependent DSB repair. In this study, we further examine phenotypic effects on cells bearing various combinations of mutations within either or both regions. Effects studied included cell killing as well as chromosomal aberration induction after 0.5-8 Gy gamma-ray irradiation delivered to synchronized cells during the <span class="hlt">G</span> 0 /<span class="hlt">G</span> 1 phase of the cell cycle. Blocking phosphorylation within the T2609 cluster was most critical regarding sensitization and depended on the number of available phosphorylation sites. It was also especially interesting that only one substitution of alanine in each of the two clusters separately abolished the restoration of wild-type sensitivity by DNA-PKcs. Similar patterns were seen for induction of chromosomal aberrations, reflecting their connection to cell killing. To study possible change in coordination between HRR and NHEJ directed repair in these DNA-PKcs mutant cell lines, we compared the induction of sister <span class="hlt">chromatid</span> exchanges (SCEs) by very low fluencies of alpha particles with mutant cells defective in the HRR pathway that is required for induction of SCEs. Levels of true SCEs <span class="hlt">induced</span> by very low fluence of alpha-particle irradiation normally seen in wild-type cells were only slightly decreased in the S2056 cluster mutants, but were completely abolished in the T2609 cluster mutants and were indistinguishable from levels seen in HRR deficient cells. Again, a single substitution in the S2056 together with a single</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040173256&hterms=incubation+time&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dincubation%2Btime','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040173256&hterms=incubation+time&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dincubation%2Btime"><span>Rejoining and misrejoining of radiation-<span class="hlt">induced</span> chromatin <span class="hlt">breaks</span>. I. experiments with human lymphocytes</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Durante, M.; George, K.; Wu, H.; Yang, T. C.</p> <p>1996-01-01</p> <p>Fluorescence in situ hybridization with a composite probe for human chromosome 4 and a probe that stained all centromeres was used to study gamma-ray <span class="hlt">induced</span> breakage, rejoining and misrejoining in prematurely condensed chromosomes in human lymphocytes. Dose-response curves for the induction of all types of aberrations in prematurely condensed human chromosomes 4 were determined immediately after irradiation and after 8 h postirradiation incubation. In addition, aberrations were measured after various incubation times from 0 to 18 h after a dose of 7 Gy. Unrejoined chromosome <span class="hlt">breaks</span> were the most frequent type of aberration observed immediately after irradiation. Approximately 15% of total aberrations observed were chromosome exchanges. After 8 h postirradiation incubation, the frequency of <span class="hlt">breaks</span> in prematurely condensed chromosomes declined to about 20% of the initial value, and chromosomal exchanges became the most frequent aberration. Results of metaphase analysis were similar to those for prematurely condensed chromosomes after 8 h incubation with the exception that a significantly lower frequency of fragments was observed. Symmetrical and asymmetrical interchanges were found at similar frequencies at all doses. No complex exchanges were observed in lymphocyte chromosomes immediately after exposure. They accounted for about 1% of total exchanges in metaphase chromosomes at doses <3 Gy and about 14% at 7 Gy. Incomplete exchanges amounted to approximately 15% of total exchanges at all doses. The kinetics of <span class="hlt">break</span> rejoining was exponential, and the frequency of exchanges increased with kinetics similar to that observed for the rejoining of the <span class="hlt">breaks</span>. This increase in the total exchanges as a function of the time between irradiation and fusion was due to a rapid increase in reciprocal interchanges, and a slower increase in complex exchanges; the frequency of incomplete exchanges increased initially, then decreased with prolonged incubation to the level observed</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27885940','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27885940"><span>Lignans from Opuntia ficus-indica seeds protect rat primary hepatocytes and Hep<span class="hlt">G</span><span class="hlt">2</span> cells against ethanol-<span class="hlt">induced</span> oxidative stress.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kim, Jung Wha; Yang, Heejung; Kim, Hyeon Woo; Kim, Hong Pyo; Sung, Sang Hyun</p> <p>2017-01-01</p> <p>Bioactivity-guided isolation of Opuntia ficus-indica (Cactaceae) seeds against ethanol-treated primary rat hepatocytes yielded six lignan compounds. Among the isolates, furofuran lignans 4-6, significantly protected rat hepatocytes against ethanol-<span class="hlt">induced</span> oxidative stress by reducing intracellular reactive oxygen species levels, preserving antioxidative defense enzyme activities, and maintaining the glutathione content. Moreover, 4 dose-dependently <span class="hlt">induced</span> the heme oxygenase-1 expression in Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3448514','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3448514"><span>Unstabilized DNA <span class="hlt">breaks</span> in HTLV-1 Tax expressing cells correlate with functional targeting of Ku80, not PKcs, XRCC4, or H<span class="hlt">2</span>AX</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p></p> <p>2012-01-01</p> <p>Background Expression of the human T-cell leukemia virus type 1 (HTLV-1) Tax oncoprotein rapidily <span class="hlt">induces</span> a significant increase of micronuclei (MN) and unstabilized DNA <span class="hlt">breaks</span> in cells. Unstabilized DNA <span class="hlt">breaks</span> can have free 3′-OH ends accessible to in situ addition of digoxygenin (DIG)-labeled dUTP using terminal deoxynucleotidyl transferase. In the present work, we used a GFP-Tax (green fluorescent protein) plasmid, which produces a functionally active GFP-tagged Tax protein, to detect the cellular target(s) for Tax which might mechanistically explain the clastogenic phenomenon. We examined the induction of MN and unstabilized DNA <span class="hlt">breaks</span> in wild type cells and cells individually knocked out for Ku80, PKcs, XRCC4, and H<span class="hlt">2</span>AX proteins. We also assessed in the same cells, the signal strengths produced by DIG-dUTP incorporation at the unstable DNA <span class="hlt">breaks</span> in the presence and absence of Tax. Results Cells mutated for PKcs, XRCC4 and H<span class="hlt">2</span>AX showed increased frequency of MN and unstabilized DNA <span class="hlt">breaks</span> in response to the expression of Tax, while cells genetically mutated for Ku80 were refractory to Tax’s induction of these cytogenetic effects. Moreover, by measuring the size of DIG-dUTP incorporation signal, which indicates the extent of unstable DNA ends, we found that Tax <span class="hlt">induces</span> larger signals than those in control cells. However, in xrs-6 cells deficient for Ku80, this Tax effect was not seen. Conclusions The data here demonstrate that clastogenic DNA damage in Tax expressing cells is explained by Tax targeting of Ku80, but not PKcs, XRCC4 or H<span class="hlt">2</span>AX, which are all proteins directly or indirectly related to the non-homologous end-joining (NHEJ) repair system. Of note, the Ku80 protein plays an important role at the initial stage of the NHEJ repair system, protecting and stabilizing DNA-<span class="hlt">breaks</span>. Accordingly, HTLV-1 Tax is shown to interfere with a normal cellular protective mechanism for stabilizing DNA <span class="hlt">breaks</span>. These DNA <span class="hlt">breaks</span>, unprotected by Ku80, are unstable and are</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4077132','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4077132"><span>Novel microtubule-targeted agent 6-chloro-4-(methoxyphenyl) coumarin <span class="hlt">induces</span> <span class="hlt">G</span><span class="hlt">2</span>-M arrest and apoptosis in HeLa cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Ma, Yi-ming; Zhou, Yu-bo; Xie, Chuan-ming; Chen, Dong-mei; Li, Jia</p> <p>2012-01-01</p> <p>Aim: To identify a novel coumarin analogue with the highest anticancer activity and to further investigate its anticancer mechanisms. Methods: The viability of cancer cells was investigated using the MTT assay. The cell cycle progression was evaluated using both flow cytometric and Western blotting analysis. Microtubule depolymerization was observed with immunocytochemistry in vivo and a tubulin depolymerization assay in vitro. Apoptosis was demonstrated using Annexin V/Propidium Iodide (PI) double-staining and sub-<span class="hlt">G</span>1 analysis. Results: Among 36 analogues of coumarin, 6-chloro-4-(methoxyphenyl) coumarin showed the best anticancer activity (IC50 value about 200 nmol/L) in HCT116 cells. The compound had a broad spectrum of anticancer activity against 9 cancer cell lines derived from colon cancer, breast cancer, liver cancer, cervical cancer, leukemia, epidermoid cancer with IC50 value of 75 nmol/L–1.57 μmol/L but with low cytotocitity against WI-38 human lung fibroblasts (IC50 value of 12.128 μmol/L). The compound (0.04–10 μmol/L) <span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span>-M phase arrest in HeLa cells in a dose-dependent manner, which was reversible after the compound was removed. The compound (10–300 μmol/L) <span class="hlt">induced</span> the depolymerization of purified porcine tubulin in vitro. Finally, the compound (0.04–<span class="hlt">2</span>.5 μmol/L) <span class="hlt">induced</span> apoptosis of HeLa cells in dose- and time-dependent manners. Conclusion: 6-Chloro-4-(methoxyphenyl) coumarin is a novel microtubule-targeting agent that <span class="hlt">induces</span> <span class="hlt">G</span><span class="hlt">2</span>–M arrest and apoptosis in HeLa cells. PMID:22266726</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li class="active"><span>18</span></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_18 --> <div id="page_19" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li class="active"><span>19</span></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="361"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29673453','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29673453"><span>[Knockdown of STAT3 inhibits proliferation and migration of Hep<span class="hlt">G</span><span class="hlt">2</span> hepatoma cells <span class="hlt">induced</span> by IFN1].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Li, Xiaofang; Wang, Yuqi; Yan, Ben; Fang, Peipei; Ma, Chao; Xu, Ning; Fu, Xiaoyan; Liang, Shujuan</p> <p>2018-02-01</p> <p>Objective To prepare lentiviruses expressing shRNA sequences targeting human signal transducer and activator of transcription 3 (STAT3) and detect the effect of STAT3 knockdown on type I interferon (IFN1)-<span class="hlt">induced</span> proliferation and migration in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Methods Four STAT3-targeting shRNA sequences (shRNA1-shRNA4) and one control sequence (Ctrl shRNA) were selected and cloned respectively into pLKO.1-sp6-pgk-GFP to construct shRNA-expressing vectors. Along with backbone psPAX<span class="hlt">2</span> and pMD<span class="hlt">2</span>.<span class="hlt">G</span> vectors, they were separately transfected into HEK293T cells to prepare lentiviruses. Hep<span class="hlt">G</span><span class="hlt">2</span> cells were infected with the lentiviruses. Cytoplastic STAT3 level was detected by Western blotting to screen effective shRNA sequence(s) targeting STAT3. Proliferation and migration of Hep<span class="hlt">G</span><span class="hlt">2</span> cells were analyzed by CCK-8 assay and Transwell TM migration and scratching assay, respectively. To detect the effect of IFN1 on cell proliferation and migration of Hep<span class="hlt">G</span><span class="hlt">2</span> cells, the cells were treated with 2000 U/mL IFNα<span class="hlt">2</span>b for indicated time and the activation of IFN-triggered STAT1 signal transduction was assayed by Western blotting. Results Two most effective STAT3-targeting shRNA sequences shRNA1 and shRNA<span class="hlt">2</span> were selected, and the expression of both STAT3 shRNA significantly decreased proliferation and migration of Hep<span class="hlt">G</span><span class="hlt">2</span> cells. When treated with IFNα<span class="hlt">2</span>b, 2000 U/mL of IFN1 showed more competent in attenuating growth and migration of Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Our data further proved that knockdown of STAT3 increased the phosphorylation of STAT1, and IFNα<span class="hlt">2</span>b further enhanced the activation of STAT1 signaling in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Conclusion Knockdown of STAT3 inhibits cell migration and growth, and rescues IFN response through up-regulating STAT1 signal transduction in Hep<span class="hlt">G</span><span class="hlt">2</span> hepatoma cells.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040173105&hterms=non+ionizing+radiation&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dnon%2Bionizing%2Bradiation','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040173105&hterms=non+ionizing+radiation&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dnon%2Bionizing%2Bradiation"><span>Non-random distribution of DNA double-strand <span class="hlt">breaks</span> <span class="hlt">induced</span> by particle irradiation</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lobrich, M.; Cooper, P. K.; Rydberg, B.; Chatterjee, A. (Principal Investigator)</p> <p>1996-01-01</p> <p>Induction of DNA double-strand <span class="hlt">breaks</span> (dsbs) in mammalian cells is dependent on the spatial distribution of energy deposition from the ionizing radiation. For high LET particle radiations the primary ionization sites occur in a correlated manner along the track of the particles, while for X-rays these sites are much more randomly distributed throughout the volume of the cell. It can therefore be expected that the distribution of dsbs linearly along the DNA molecule also varies with the type of radiation and the ionization density. Using pulsed-field gel and conventional gel techniques, we measured the size distribution of DNA molecules from irradiated human fibroblasts in the total range of 0.1 kbp-10 Mbp for X-rays and high LET particles (N ions, 97 keV/microns and Fe ions, 150 keV/microns). On a mega base pair scale we applied conventional pulsed-field gel electrophoresis techniques such as measurement of the fraction of DNA released from the well (FAR) and measurement of breakage within a specific NotI restriction fragment (hybridization assay). The induction rate for widely spaced <span class="hlt">breaks</span> was found to decrease with LET. However, when the entire distribution of radiation-<span class="hlt">induced</span> fragments was analysed, we detected an excess of fragments with sizes below about 200 kbp for the particles compared with X-irradiation. X-rays are thus more effective than high LET radiations in producing large DNA fragments but less effective in the production of smaller fragments. We determined the total induction rate of dsbs for the three radiations based on a quantitative analysis of all the measured radiation-<span class="hlt">induced</span> fragments and found that the high LET particles were more efficient than X-rays at <span class="hlt">inducing</span> dsbs, indicating an increasing total efficiency with LET. Conventional assays that are based only on the measurement of large fragments are therefore misleading when determining total dsb induction rates of high LET particles. The possible biological significance of this non</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22215943-mitochondrial-aquaporin-knockdown-human-hepatoma-hepg2-cells-causes-ros-induced-mitochondrial-depolarization-loss-viability','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22215943-mitochondrial-aquaporin-knockdown-human-hepatoma-hepg2-cells-causes-ros-induced-mitochondrial-depolarization-loss-viability"><span>Mitochondrial aquaporin-8 knockdown in human hepatoma Hep<span class="hlt">G</span><span class="hlt">2</span> cells causes ROS-<span class="hlt">induced</span> mitochondrial depolarization and loss of viability</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Marchissio, Maria Julia; Francés, Daniel Eleazar Antonio; Carnovale, Cristina Ester</p> <p></p> <p>Human aquaporin-8 (AQP8) channels facilitate the diffusional transport of H{sub <span class="hlt">2</span>}O{sub <span class="hlt">2</span>} across membranes. Since AQP8 is expressed in hepatic inner mitochondrial membranes, we studied whether mitochondrial AQP8 (mtAQP8) knockdown in human hepatoma Hep<span class="hlt">G</span><span class="hlt">2</span> cells impairs mitochondrial H{sub <span class="hlt">2</span>}O{sub <span class="hlt">2</span>} release, which may lead to organelle dysfunction and cell death. We confirmed AQP8 expression in Hep<span class="hlt">G</span><span class="hlt">2</span> inner mitochondrial membranes and found that 72 h after cell transfection with siRNAs targeting two different regions of the human AQP8 molecule, mtAQP8 protein specifically decreased by around 60% (p < 0.05). Studies in isolated mtAQP8-knockdown mitochondria showed that H{sub <span class="hlt">2</span>}O{sub <span class="hlt">2</span>} release, assessedmore » by Amplex Red, was reduced by about 45% (p < 0.05), an effect not observed in digitonin-permeabilized mitochondria. mtAQP8-knockdown cells showed an increase in mitochondrial ROS, assessed by dichlorodihydrofluorescein diacetate (+ 120%, p < 0.05) and loss of mitochondrial membrane potential (− 80%, p < 0.05), assessed by tetramethylrhodamine-coupled quantitative fluorescence microscopy. The mitochondria-targeted antioxidant MitoTempol prevented ROS accumulation and dissipation of mitochondrial membrane potential. Cyclosporin A, a mitochondrial permeability transition pore blocker, also abolished the mtAQP8 knockdown-<span class="hlt">induced</span> mitochondrial depolarization. Besides, the loss of viability in mtAQP8 knockdown cells verified by MTT assay, LDH leakage, and trypan blue exclusion test could be prevented by cyclosporin A. Our data on human hepatoma Hep<span class="hlt">G</span><span class="hlt">2</span> cells suggest that mtAQP8 facilitates mitochondrial H{sub <span class="hlt">2</span>}O{sub <span class="hlt">2</span>} release and that its defective expression causes ROS-<span class="hlt">induced</span> mitochondrial depolarization via the mitochondrial permeability transition mechanism, and cell death. -- Highlights: ► Aquaporin-8 is expressed in mitochondria of human hepatoma Hep<span class="hlt">G</span><span class="hlt">2</span> cells. ► Aquaporin-8 knockdown impairs mitochondrial H{sub <span class="hlt">2</span>}O{sub <span class="hlt">2</span>} release and increases ROS. </p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21140965-alpha-estradiol-arrests-cell-cycle-progression-sub-induces-apoptotic-cell-death-human-acute-leukemia-jurkat-cells','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21140965-alpha-estradiol-arrests-cell-cycle-progression-sub-induces-apoptotic-cell-death-human-acute-leukemia-jurkat-cells"><span>17{alpha}-Estradiol arrests cell cycle progression at <span class="hlt">G</span>{sub <span class="hlt">2</span>}/M and <span class="hlt">induces</span> apoptotic cell death in human acute leukemia Jurkat T cells</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Jun, Do Youn; Park, Hae Sun; Kim, Jun Seok</p> <p>2008-09-15</p> <p>A pharmacological dose (<span class="hlt">2</span>.5-10 {mu}M) of 17{alpha}-estradiol (17{alpha}-E{sub <span class="hlt">2</span>}) exerted a cytotoxic effect on human leukemias Jurkat T and U937 cells, which was not suppressed by the estrogen receptor (ER) antagonist ICI 182,780. Along with cytotoxicity in Jurkat T cells, several apoptotic events including mitochondrial cytochrome c release, activation of caspase-9, -3, and -8, PARP degradation, and DNA fragmentation were <span class="hlt">induced</span>. The cytotoxicity of 17{alpha}-E{sub <span class="hlt">2</span>} was not blocked by the anti-Fas neutralizing antibody ZB-4. While undergoing apoptosis, there was a remarkable accumulation of <span class="hlt">G</span>{sub <span class="hlt">2</span>}/M cells with the upregulatoin of cdc<span class="hlt">2</span> kinase activity, which was reflected in the Thr56more » phosphorylation of Bcl-<span class="hlt">2</span>. Dephosphorylation at Tyr15 and phosphorylation at Thr161 of cdc<span class="hlt">2</span>, and significant increase in the cyclin B1 level were underlying factors for the cdc<span class="hlt">2</span> kinase activation. Whereas the 17{alpha}-E{sub <span class="hlt">2</span>}-<span class="hlt">induced</span> apoptosis was completely abrogated by overexpression of Bcl-<span class="hlt">2</span> or by pretreatment with the pan-caspase inhibitor z-VAD-fmk, the accumulation of <span class="hlt">G</span>{sub <span class="hlt">2</span>}/M cells significantly increased. The caspase-8 inhibitor z-IETD-fmk failed to influence 17{alpha}-E{sub <span class="hlt">2</span>}-mediated caspase-9 activation, but it markedly reduced caspase-3 activation and PARP degradation with the suppression of apoptosis, indicating the contribution of caspase-8; not as an upstream event of the mitochondrial cytochrome c release, but to caspase-3 activation. In the presence of hydroxyurea, which blocked the cell cycle progression at the <span class="hlt">G</span>{sub 1}/S boundary, 17{alpha}-E{sub <span class="hlt">2</span>} failed to <span class="hlt">induce</span> the <span class="hlt">G</span>{sub <span class="hlt">2</span>}/M arrest as well as apoptosis. These results demonstrate that the cytotoxicity of 17{alpha}-E{sub <span class="hlt">2</span>} toward Jurkat T cells is attributable to apoptosis mainly <span class="hlt">induced</span> in <span class="hlt">G</span>{sub <span class="hlt">2</span>}/M-arrested cells, in an ER-independent manner, via a mitochondria-dependent caspase pathway regulated by Bcl-<span class="hlt">2</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26371841','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26371841"><span>Chronic cat allergen exposure <span class="hlt">induces</span> a TH<span class="hlt">2</span> cell-dependent Ig<span class="hlt">G</span>4 response related to low sensitization.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Renand, Amedee; Archila, Luis D; McGinty, John; Wambre, Erik; Robinson, David; Hales, Belinda J; Thomas, Wayne R; Kwok, William W</p> <p>2015-12-01</p> <p>In human subjects, allergen tolerance has been observed after high-dose allergen exposure or after completed allergen immunotherapy, which is related to the accumulation of anti-inflammatory Ig<span class="hlt">G</span>4. However, the specific T-cell response that leads to Ig<span class="hlt">G</span>4 induction during chronic allergen exposure remains poorly understood. We sought to evaluate the relationship between cat allergen-specific T-cell frequency, cat allergen-specific IgE and Ig<span class="hlt">G</span>4 titers, and clinical status in adults with cat allergy with and without cat ownership and the cellular mechanism by which Ig<span class="hlt">G</span>4 is produced. Fel d 1-, Fel d 4-, Fel d 7-, and Fel d 8-specific T-cell responses were characterized by CD154 expression after antigen stimulation. In allergic subjects without cat ownership, the frequency of cat allergen (Fel d 1 and Fel d 4)-specific TH<span class="hlt">2</span> (sTH<span class="hlt">2</span>) cells correlates with higher IgE levels and is linked to asthma. Paradoxically, we observed that subjects with cat allergy and chronic cat exposure maintain a high frequency of sTH<span class="hlt">2</span> cells, which correlates with higher Ig<span class="hlt">G</span>4 levels and low sensitization. B cells from allergic, but not nonallergic subjects, are able to produce Ig<span class="hlt">G</span>4 after cognate interactions with sTH<span class="hlt">2</span> clones and Fel d 1 peptide or the Fel d 1 recombinant protein. These experiments suggest that (1) allergen-experienced B cells with the capacity to produce Ig<span class="hlt">G</span>4 are present in allergic subjects and (<span class="hlt">2</span>) cat allergen exposure <span class="hlt">induces</span> an Ig<span class="hlt">G</span>4 response in a TH<span class="hlt">2</span> cell-dependent manner. Thus Ig<span class="hlt">G</span>4 accumulation could be mediated by chronic activation of the TH<span class="hlt">2</span> response, which in turn drives desensitization. Copyright © 2015 American Academy of Allergy, Asthma & Immunology. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29570431','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29570431"><span><span class="hlt">G</span> protein-coupled receptor kinase-<span class="hlt">2</span>-deficient mice are protected from dextran sodium sulfate-<span class="hlt">induced</span> acute colitis.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Steury, Michael D; Kang, Ho Jun; Lee, Taehyung; Lucas, Peter C; McCabe, Laura R; Parameswaran, Narayanan</p> <p>2018-06-01</p> <p><span class="hlt">G</span> protein-coupled receptor kinase <span class="hlt">2</span> (GRK<span class="hlt">2</span>) is a serine/threonine kinase and plays a key role in different disease processes. Previously, we showed that GRK<span class="hlt">2</span> knockdown enhances wound healing in colonic epithelial cells. Therefore, we hypothesized that ablation of GRK<span class="hlt">2</span> would protect mice from dextran sodium sulfate (DSS)-<span class="hlt">induced</span> acute colitis. To test this, we administered DSS to wild-type (GRK<span class="hlt">2</span> +/+ ) and GRK<span class="hlt">2</span> heterozygous (GRK +/- ) mice in their drinking water for 7 days. As predicted, GRK<span class="hlt">2</span> +/- mice were protected from colitis as demonstrated by decreased weight loss (20% loss in GRK<span class="hlt">2</span> +/+ vs. 11% loss in GRK<span class="hlt">2</span> +/- ). lower disease activity index (GRK<span class="hlt">2</span> +/+ 9.1 vs GRK<span class="hlt">2</span> +/- 4.1), and increased colon lengths (GRK<span class="hlt">2</span> +/+ 4.7 cm vs GRK<span class="hlt">2</span> +/- 5.3 cm). To examine the mechanisms by which GRK<span class="hlt">2</span> +/- mice are protected from colitis, we investigated expression of inflammatory genes in the colon as well as immune cell profiles in colonic lamina propria, mesenteric lymph node, and in bone marrow. Our results did not reveal differences in immune cell profiles between the two genotypes. However, expression of inflammatory genes was significantly decreased in DSS-treated GRK<span class="hlt">2</span> +/- mice compared with GRK<span class="hlt">2</span> +/+ . To understand the mechanisms, we generated myeloid-specific GRK<span class="hlt">2</span> knockout mice and subjected them to DSS-<span class="hlt">induced</span> colitis. Similar to whole body GRK<span class="hlt">2</span> heterozygous knockout mice, myeloid-specific knockout of GRK<span class="hlt">2</span> was sufficient for the protection from DSS-<span class="hlt">induced</span> colitis. Together our results indicate that deficiency of GRK<span class="hlt">2</span> protects mice from DSS-<span class="hlt">induced</span> colitis and further suggests that the mechanism of this effect is likely via GRK<span class="hlt">2</span> regulation of inflammatory genes in the myeloid cells.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22465688-phosphoramide-mustard-exposure-induces-dna-adduct-formation-dna-damage-repair-response-rat-ovarian-granulosa-cells','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22465688-phosphoramide-mustard-exposure-induces-dna-adduct-formation-dna-damage-repair-response-rat-ovarian-granulosa-cells"><span>Phosphoramide mustard exposure <span class="hlt">induces</span> DNA adduct formation and the DNA damage repair response in rat ovarian granulosa cells</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Ganesan, Shanthi, E-mail: shanthig@iastate.edu; Keating, Aileen F., E-mail: akeating@iastate.edu</p> <p></p> <p>Phosphoramide mustard (PM), the ovotoxic metabolite of the anti-cancer agent cyclophosphamide (CPA), destroys rapidly dividing cells by forming NOR-<span class="hlt">G</span>-OH, NOR-<span class="hlt">G</span> and <span class="hlt">G-NOR-G</span> adducts with DNA, potentially leading to DNA damage. A previous study demonstrated that PM <span class="hlt">induces</span> ovarian DNA damage in rat ovaries. To investigate whether PM <span class="hlt">induces</span> DNA adduct formation, DNA damage and induction of the DNA repair response, rat spontaneously immortalized granulosa cells (SIGCs) were treated with vehicle control (1% DMSO) or PM (3 or 6 μM) for 24 or 48 h. Cell viability was reduced (P < 0.05) after 48 h of exposure to 3 or 6more » μM PM. The NOR-<span class="hlt">G</span>-OH DNA adduct was detected after 24 h of 6 μM PM exposure, while the more cytotoxic <span class="hlt">G-NOR-G</span> DNA adduct was formed after 48 h by exposure to both PM concentrations. Phosphorylated H<span class="hlt">2</span>AX (γH<span class="hlt">2</span>AX), a marker of DNA double stranded <span class="hlt">break</span> occurrence, was also increased by PM exposure, coincident with DNA adduct formation. Additionally, induction of genes (Atm, Parp1, Prkdc, Xrcc6, and Brca1) and proteins (ATM, γH<span class="hlt">2</span>AX, PARP-1, PRKDC, XRCC6, and BRCA1) involved in DNA repair were observed in both a time- and dose-dependent manner. These data support that PM <span class="hlt">induces</span> DNA adduct formation in ovarian granulosa cells, <span class="hlt">induces</span> DNA damage and elicits the ovarian DNA repair response. - Highlights: • PM forms ovarian DNA adducts. • DNA damage marker γH<span class="hlt">2</span>AX increased by PM exposure. • PM <span class="hlt">induces</span> ovarian DNA double strand <span class="hlt">break</span> repair.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29078142','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29078142"><span>Electronegative L5-LDL <span class="hlt">induces</span> the production of <span class="hlt">G</span>-CSF and GM-CSF in human macrophages through LOX-1 involving NF-κB and ERK<span class="hlt">2</span> activation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yang, Tzu-Ching; Chang, Po-Yuan; Kuo, Tzu-Ling; Lu, Shao-Chun</p> <p>2017-12-01</p> <p>Circulating levels of granulocyte colony-stimulating factor (<span class="hlt">G</span>-CSF) and granulocyte macrophage colony-stimulating factor (GM-CSF) are associated with the severity of acute myocardial infarction (AMI). However, what causes increases in <span class="hlt">G</span>-CSF and GM-CSF is unclear. In this study, we investigated whether L5-low-density lipoprotein (LDL), a mildly oxidized LDL from AMI, can <span class="hlt">induce</span> <span class="hlt">G</span>-CSF and GM-CSF production in human macrophages. L1-LDL and L5-LDL were isolated through anion-exchange chromatography from AMI plasma. Human macrophages derived from THP-1 and peripheral blood mononuclear cells were treated with L1-LDL, L5-LDL, or copper-oxidized LDL (Cu-oxLDL) and <span class="hlt">G</span>-CSF and GM-CSF protein levels in the medium were determined. In addition, the effects of L5-LDL on <span class="hlt">G</span>-CSF and GM-CSF production were tested in lectin-type oxidized LDL receptor-1 (LOX-1), CD36, extracellular signal-regulated kinase (ERK) 1, and ERK<span class="hlt">2</span> knockdown THP-1 macrophages. L5-LDL but not L1-LDL or Cu-oxLDL significantly <span class="hlt">induced</span> production of <span class="hlt">G</span>-CSF and GM-CSF in macrophages. In vitro oxidation of L1-LDL and L5-LDL altered their ability to <span class="hlt">induce</span> <span class="hlt">G</span>-CSF and GM-CSF, suggesting that the degree of oxidation is critical for the effects. Knockdown and antibody neutralization experiments suggested that the effects were caused by LOX-1. In addition, nuclear factor (NF)-κB and ERK1/<span class="hlt">2</span> inhibition resulted in marked reductions of L5-LDL-<span class="hlt">induced</span> <span class="hlt">G</span>-CSF and GM-CSF production. Moreover, knockdown of ERK<span class="hlt">2</span>, but not ERK1, hindered L5-LDL-<span class="hlt">induced</span> <span class="hlt">G</span>-CSF and GM-CSF production. The results indicate that L5-LDL, a naturally occurring mild oxidized LDL, <span class="hlt">induced</span> <span class="hlt">G</span>-CSF and GM-CSF production in human macrophages through LOX-1, ERK<span class="hlt">2</span>, and NF-κB dependent pathways. Copyright © 2017 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24451844','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24451844"><span>Active immunity <span class="hlt">induced</span> by passive Ig<span class="hlt">G</span> post-exposure protection against ricin.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hu, Charles Chen; Yin, Junfei; Chau, Damon; Cherwonogrodzky, John W; Hu, Wei-Gang</p> <p>2014-01-21</p> <p>Therapeutic antibodies can confer an instant protection against biothreat agents when administered. In this study, intact Ig<span class="hlt">G</span> and F(ab')<span class="hlt">2</span> from goat anti-ricin hyperimmune sera were compared for the protection against lethal ricin mediated intoxication. Similar ricin-binding affinities and neutralizing activities in vitro were observed between Ig<span class="hlt">G</span> and F(ab')<span class="hlt">2</span> when compared at the same molar concentration. In a murine ricin intoxication model, both Ig<span class="hlt">G</span> and F(ab')<span class="hlt">2</span> could rescue 100% of the mice by one dose (3 nmol) administration of antibodies 1 hour after 5 × LD50 ricin challenge. Nine days later, when the rescued mice received a second ricin challenge (5 × LD50), only the Ig<span class="hlt">G</span>-treated mice survived; the F(ab')<span class="hlt">2</span>-treated mice did not. The experimental design excluded the possibility of residual goat Ig<span class="hlt">G</span> responsible for the protection against the second ricin challenge. Results confirmed that the active immunity against ricin in mice was <span class="hlt">induced</span> quickly following the passive delivery of a single dose of goat Ig<span class="hlt">G</span> post-exposure. Furthermore, it was demonstrated that the <span class="hlt">induced</span> active immunity against ricin in mice lasted at least 5 months. Therefore, passive Ig<span class="hlt">G</span> therapy not only provides immediate protection to the victim after ricin exposure, but also elicits an active immunity against ricin that subsequently results in long term protection.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3920268','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3920268"><span>Active Immunity <span class="hlt">Induced</span> by Passive Ig<span class="hlt">G</span> Post-Exposure Protection against Ricin</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Hu, Charles Chen; Yin, Junfei; Chau, Damon; Cherwonogrodzky, John W.; Hu, Wei-Gang</p> <p>2014-01-01</p> <p>Therapeutic antibodies can confer an instant protection against biothreat agents when administered. In this study, intact Ig<span class="hlt">G</span> and F(ab’)<span class="hlt">2</span> from goat anti-ricin hyperimmune sera were compared for the protection against lethal ricin mediated intoxication. Similar ricin-binding affinities and neutralizing activities in vitro were observed between Ig<span class="hlt">G</span> and F(ab’)<span class="hlt">2</span> when compared at the same molar concentration. In a murine ricin intoxication model, both Ig<span class="hlt">G</span> and F(ab’)<span class="hlt">2</span> could rescue 100% of the mice by one dose (3 nmol) administration of antibodies 1 hour after 5 × LD50 ricin challenge. Nine days later, when the rescued mice received a second ricin challenge (5 × LD50), only the Ig<span class="hlt">G</span>-treated mice survived; the F(ab’)<span class="hlt">2</span>-treated mice did not. The experimental design excluded the possibility of residual goat Ig<span class="hlt">G</span> responsible for the protection against the second ricin challenge. Results confirmed that the active immunity against ricin in mice was <span class="hlt">induced</span> quickly following the passive delivery of a single dose of goat Ig<span class="hlt">G</span> post-exposure. Furthermore, it was demonstrated that the <span class="hlt">induced</span> active immunity against ricin in mice lasted at least 5 months. Therefore, passive Ig<span class="hlt">G</span> therapy not only provides immediate protection to the victim after ricin exposure, but also elicits an active immunity against ricin that subsequently results in long term protection. PMID:24451844</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27396604','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27396604"><span>Inhibition of Aurora A Kinase by Alisertib <span class="hlt">Induces</span> Autophagy and Cell Cycle Arrest and Increases Chemosensitivity in Human Hepatocellular Carcinoma Hep<span class="hlt">G</span><span class="hlt">2</span> Cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhu, Qiaohua; Yu, Xinfa; Zhou, Zhi-Wei; Zhou, Chengyu; Chen, Xiao-Wu; Zhou, Shu-Feng</p> <p>2017-01-01</p> <p>Aurora A kinase represent a feasible target in cancer therapy. To evaluate the proteomic response of human liver carcinoma cells to alisertib (ALS) and identify the molecular targets of ALS, we examined the effects of ALS on the proliferation, cell cycle, autophagy, apoptosis, and chemosensitivity in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. The stable-isotope labeling by amino acids in cell culture (SILAC) based quantitative proteomic study was performed to evaluate the proteomic response to ALS. Cell cycle distribution and apoptosis were assessed using flow cytometry and autophagy was determined using flow cytometry and confocal microscopy. Our SILAC proteomic study showed that ALS regulated the expression of 914 proteins, with 407 molecules being up-regulated and 507 molecules being down-regulated in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Ingenuity pathway analysis (IPA) and KEGG pathway analysis identified 146 and 32 signaling pathways were regulated by ALS, respectively, which were associated with cell survival, programmed cell death, and nutrition-energy metabolism. Subsequently, the verification experiments showed that ALS remarkably arrested Hep<span class="hlt">G</span><span class="hlt">2</span> cells in <span class="hlt">G</span><span class="hlt">2</span>/M phase and led to an accumulation of aneuploidy via regulating the expression of key cell cycle regulators. ALS <span class="hlt">induced</span> a marked autophagy in a concentration- and time-dependent manner via the phosphatidylinositol 3-kinase (PI3K)/protein kinase B (Akt)/mammalian target of rapamycin (mTOR) signaling pathway. Autophagy inhibition promoted the pro-apoptotic effect of ALS, indicating a cyto-protective role of ALS-<span class="hlt">induced</span> autophagy. ALS increased the chemosensitivity of Hep<span class="hlt">G</span><span class="hlt">2</span> cells to cisplatin and doxorubicin. Taken together, ALS <span class="hlt">induces</span> autophagy and cell cycle arrest in Hep<span class="hlt">G</span><span class="hlt">2</span> cells via PI3K/Akt/mTOR-mediated pathway. Autophagy inhibition may promote the anticancer effect of ALS and sensitize the chemotherapy in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Copyright© Bentham Science Publishers; For any queries, please email at epub@benthamscience.org.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29383524','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29383524"><span>DNA double-strand <span class="hlt">breaks</span> in blood lymphocytes <span class="hlt">induced</span> by two-day 99mTc-MIBI myocardial perfusion scintigraphy.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Rief, Matthias; Hartmann, Lisa; Geisel, Dominik; Richter, Felicitas; Brenner, Winfried; Dewey, Marc</p> <p>2018-07-01</p> <p>To investigate DNA double-strand <span class="hlt">breaks</span> (DSBs) in blood lymphocytes <span class="hlt">induced</span> by two-day 99m Tc-MIBI myocardial perfusion scintigraphy (MPS) using y-H<span class="hlt">2</span>AX immunofluorescence microscopy and to correlate the results with 99m Tc activity in blood samples. Eleven patients who underwent two-day MPS were included. DSB blood sampling was performed before and 5min, 1h and 24h after the first and second radiotracer injections. 99m Tc activity was measured in each blood sample. For immunofluorescence microscopy, distinct foci representing DSBs were quantified in lymphocytes after staining for the phosphorylated histone variant y-H<span class="hlt">2</span>AX. The 99m Tc-MIBI activity measured on days one and two was similar (254±25 and 258±27 MBq; p=0.594). Compared with baseline DSB foci (0.09±0.05/cell), a significant increase was found at 5min (0.19±0.04/cell) and 1h (0.18±0.04/cell) after the first injection and at 5min and 1h after the second injection (0.21±0.03 and 0.19±0.04/cell, respectively; p=0.003 for both). At 24h after the first and second injections, the number of DSB foci had returned to baseline (0.06±0.02 and 0.12±0.05/cell, respectively). 99m Tc activity levels in peripheral blood samples correlated well with DSB counts (r=0.451). DSB counts reflect 99m Tc-MIBI activity after injection for two-day MPS, and might allow individual monitoring of biological effects of cardiac nuclear imaging. • Myocardial perfusion scintigraphy using 99m Tc <span class="hlt">induces</span> time-dependent double-strand <span class="hlt">breaks</span> (DSBs) • γ-H<span class="hlt">2</span>AX immunofluorescence microscopy shows DSB as an early response to radiotracer injection • Activity measurements of 99m Tc correlate well with detected DSB • DSB foci <span class="hlt">induced</span> by 99m Tc return to baseline 24h after radiotracer injection.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24632106','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24632106"><span>Tributyltin <span class="hlt">induces</span> a <span class="hlt">G</span><span class="hlt">2</span>/M cell cycle arrest in human amniotic cells via PP<span class="hlt">2</span>A inhibition-mediated inactivation of the ERK1/<span class="hlt">2</span> cascades.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Yali; Guo, Zonglou; Xu, Lihong</p> <p>2014-03-01</p> <p>The molecular mechanisms underlying the cell cycle alterations <span class="hlt">induced</span> by tributyltin (TBT), a highly toxic environmental contaminant, remain elusive. In this study, cell cycle progression and some key regulators in <span class="hlt">G</span><span class="hlt">2</span>/M phase were investigated in human amniotic cells treated with TBT. Furthermore, protein phosphatase (PP) <span class="hlt">2</span>A and the ERK cascades were examined. The results showed that TBT caused a <span class="hlt">G</span><span class="hlt">2</span>/M cell cycle arrest that was accompanied by a decrease in the total cdc25C protein level and an increase in the p-cdc<span class="hlt">2</span> level in the nucleus. TBT caused a decrease in PP<span class="hlt">2</span>A activity and inhibited the ERK cascade by inactivating Raf-1, resulting in the dephosphorylation of MEK1/<span class="hlt">2</span>, ERK1/<span class="hlt">2</span>, and c-Myc. Taken together, TBT leads to a <span class="hlt">G</span><span class="hlt">2</span>/M cell cycle arrest in FL cells, an increase in p-cdc<span class="hlt">2</span> and a decrease in the levels of total cdc25C protein, which may be caused by the PP<span class="hlt">2</span>A inhibition-mediated inactivation of the ERK1/<span class="hlt">2</span> cascades. Copyright © 2014 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27261572','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27261572"><span>Procyanidins, from Castanea mollissima Bl. shell, <span class="hlt">induces</span> autophagy following apoptosis associated with PI3K/AKT/mTOR inhibition in Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Haihui; Luo, Xiaoping; Ke, Jiajia; Duan, Yuqing; He, Yuanqing; Zhang, Di; Cai, Meihong; Sun, Guibo; Sun, Xiaobo</p> <p>2016-07-01</p> <p>Procyanidins from Castanea mollissima Bl. shell (CSPCs) <span class="hlt">induced</span> autophagy and apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells and its mechanism remains to be examined. In this paper, autophagy was measured by the lipid modification of light chain-3 (LC3) and the formation of autophagosomes. Hoechst 33258 staining and flow cytometer analysis were used to measure apoptosis. The western blot analysis was used to examine the effects of CSPCs on the expression of LC3, PI3K, phosphorylation of AKT, mTOR, Bcl-<span class="hlt">2</span>, Bad, Bax, BID and cleaved caspase 3 in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. The results showed that 3-methyladenine (3-MA) and apoptosis inhibitor (Z-VAD) could inhibited the death of Hep<span class="hlt">G</span><span class="hlt">2</span> <span class="hlt">induced</span> by CSPCs for 48h (150μ<span class="hlt">g</span>/mL). CSPCs <span class="hlt">induced</span> the accumulation of autophagosomes and microtubule-associated proteins light chain 3-II (LC3-II, a marker of autophagy). P-AKT, PI3K and mTOR were significantly decreased on CSPCs exposure. However, these phenomena were not observed in the group pretreated with the autophagy inhibitor 3-MA and Z-VAD. CSPCs also <span class="hlt">induced</span> the expression of Bad, Bax and Beclin-1 proteins and decreased the expression of Bcl-<span class="hlt">2</span>, which was inhibited by 3-MA and Z-VAD. Moreover the apoptotic cell death could be inhibited by 3-MA. In addition, inhibition of LC3-II by siRNA-dependent knockdown attenuated the cleavage of caspase 3. These results suggested CSPCs could trigger autophagy via inhibition of the PI3K/AKT/mTOR signaling pathway, enhanced apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells which may be associated with the mitochondria-dependent signaling way. Copyright © 2016 Elsevier Masson SAS. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/7304942-binding-radiation-induced-phenylalanine-radicals-dna-influence-biological-activity-dna-its-sensitivity-induction-breaks-gamma-rays','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/7304942-binding-radiation-induced-phenylalanine-radicals-dna-influence-biological-activity-dna-its-sensitivity-induction-breaks-gamma-rays"><span>Binding of radiation-<span class="hlt">induced</span> phenylalanine radicals to DNA: influence on the biological activity of the DNA and on its sensitivity to the induction of <span class="hlt">breaks</span> by gamma rays</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Vanderschans, G.P.; Vanrijn, C.J.S.; Bleichrodt, J.F.</p> <p>1975-11-01</p> <p>When an aqueous solution of double-stranded deoxyribonucleic acid (DNA) of bacteriophage PM<span class="hlt">2</span> containing phenylalanine and saturated with N<span class="hlt">2</span>O is irradiated with gamma rays, radiation <span class="hlt">induced</span> phenylalanine radicals are bound covalently. Under the conditions used about 25 phenylalanine molecules may be bound per lethal hit. Also for single-stranded PM<span class="hlt">2</span> DNA most of the phenylalanine radicals bound are nonlethal. Evidence is presented that in double-stranded DNA an appreciable fraction of the single-strand <span class="hlt">breaks</span> is <span class="hlt">induced</span> by phenylalanine radicals. Radiation products of phenylalanine and the phenylalanine bound to the DNA decrease the sensitivity of the DNA to the induction of single-strand <span class="hlt">breaks</span>. Theremore » are indications that the high efficiency of protection by radiation products of phenylalanine is due to their positive charge, which will result in a relatively high concentration of these compounds in the vicinity of the negatively charged DNA molecules. (Author) (GRA)« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24222901','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24222901"><span>Depigmented allergoids reveal new epitopes with capacity to <span class="hlt">induce</span> Ig<span class="hlt">G</span> blocking antibodies.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>López-Matas, M Angeles; Gallego, Mayte; Iraola, Víctor; Robinson, Douglas; Carnés, Jerónimo</p> <p>2013-01-01</p> <p>The synthesis of allergen-specific blocking IgGs that interact with IgE after allergen immunotherapy (SIT) has been related to clinical efficacy. The objectives were to investigate the epitope specificity of Ig<span class="hlt">G</span>-antibodies <span class="hlt">induced</span> by depigmented-polymerized (Dpg-Pol) allergoids and unmodified allergen extracts, and examine IgE-blocking activity of <span class="hlt">induced</span> Ig<span class="hlt">G</span>-antibodies. Rabbits were immunized with native and Dpg-Pol extracts of birch pollen, and serum samples were obtained. Recognition of linear Ig<span class="hlt">G</span>-epitopes of Bet v 1 and Bet v <span class="hlt">2</span> and the capacity of these Ig<span class="hlt">G</span>-antibodies to block binding of human-IgE was determined. Serum from rabbits immunized with native extracts recognised 11 linear epitopes from Bet v 1, while that from Dpg-Pol-immunized animals recognised 8. For Bet v <span class="hlt">2</span>, 8 epitopes were recognized by Ig<span class="hlt">G</span> from native immunized animals, and 9 from Dpg-Pol immunized one. Dpg-Pol and native immunized serum did not always recognise the same epitopes, but specific-Ig<span class="hlt">G</span> from both could block human-IgE binding sites for native extract. Depigmented-polymerized birch extract stimulates the synthesis of specific Ig<span class="hlt">G</span>-antibodies which recognize common but also novel epitopes compared with native extracts. Ig<span class="hlt">G</span>-antibodies <span class="hlt">induced</span> by Dpg-Pol effectively inhibit human-IgE binding to allergens which may be part of the mechanism of action of SIT.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27157131','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27157131"><span>House dust mite-<span class="hlt">induced</span> asthma causes oxidative damage and DNA double-strand <span class="hlt">breaks</span> in the lungs.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chan, Tze Khee; Loh, Xin Yi; Peh, Hong Yong; Tan, W N Felicia; Tan, W S Daniel; Li, Na; Tay, Ian J J; Wong, W S Fred; Engelward, Bevin P</p> <p>2016-07-01</p> <p>Asthma is related to airway inflammation and oxidative stress. High levels of reactive oxygen and nitrogen species can <span class="hlt">induce</span> cytotoxic DNA damage. Nevertheless, little is known about the possible role of allergen-<span class="hlt">induced</span> DNA damage and DNA repair as modulators of asthma-associated pathology. We sought to study DNA damage and DNA damage responses <span class="hlt">induced</span> by house dust mite (HDM) in vivo and in vitro. We measured DNA double-strand <span class="hlt">breaks</span> (DSBs), DNA repair proteins, and apoptosis in an HDM-<span class="hlt">induced</span> allergic asthma model and in lung samples from asthmatic patients. To study DNA repair, we treated mice with the DSB repair inhibitor NU7441. To study the direct DNA-damaging effect of HDM on human bronchial epithelial cells, we exposed BEAS-<span class="hlt">2</span>B cells to HDM and measured DNA damage and reactive oxygen species levels. HDM challenge increased lung levels of oxidative damage to proteins (3-nitrotyrosine), lipids (8-isoprostane), and nucleic acid (8-oxoguanine). Immunohistochemical evidence for HDM-<span class="hlt">induced</span> DNA DSBs was revealed by increased levels of the DSB marker γ Histone <span class="hlt">2</span>AX (H<span class="hlt">2</span>AX) foci in bronchial epithelium. BEAS-<span class="hlt">2</span>B cells exposed to HDM showed enhanced DNA damage, as measured by using the comet assay and γH<span class="hlt">2</span>AX staining. In lung tissue from human patients with asthma, we observed increased levels of DNA repair proteins and apoptosis, as shown by caspase-3 cleavage, caspase-activated DNase levels, and terminal deoxynucleotidyl transferase-mediated dUTP nick end-labeling staining. Notably, NU7441 augmented DNA damage and cytokine production in the bronchial epithelium and apoptosis in the allergic airway, implicating DSBs as an underlying driver of asthma pathophysiology. This work calls attention to reactive oxygen and nitrogen species and HDM-<span class="hlt">induced</span> cytotoxicity and to a potential role for DNA repair as a modulator of asthma-associated pathophysiology. Copyright © 2016 American Academy of Allergy, Asthma & Immunology. Published by Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3471751','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3471751"><span>Induction of Cardiac Fibrosis by β-Blocker in <span class="hlt">G</span> Protein-independent and <span class="hlt">G</span> Protein-coupled Receptor Kinase 5/β-Arrestin<span class="hlt">2</span>-dependent Signaling Pathways*</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Nakaya, Michio; Chikura, Satsuki; Watari, Kenji; Mizuno, Natsumi; Mochinaga, Koji; Mangmool, Supachoke; Koyanagi, Satoru; Ohdo, Shigehiro; Sato, Yoji; Ide, Tomomi; Nishida, Motohiro; Kurose, Hitoshi</p> <p>2012-01-01</p> <p><span class="hlt">G</span>-protein coupled receptors (GPCRs) have long been known as receptors that activate <span class="hlt">G</span> protein-dependent cellular signaling pathways. In addition to the <span class="hlt">G</span> protein-dependent pathways, recent reports have revealed that several ligands called “biased ligands” elicit <span class="hlt">G</span> protein-independent and β-arrestin-dependent signaling through GPCRs (biased agonism). Several β-blockers are known as biased ligands. All β-blockers inhibit the binding of agonists to the β-adrenergic receptors. In addition to β-blocking action, some β-blockers are reported to <span class="hlt">induce</span> cellular responses through <span class="hlt">G</span> protein-independent and β-arrestin-dependent signaling pathways. However, the physiological significance <span class="hlt">induced</span> by the β-arrestin-dependent pathway remains much to be clarified in vivo. Here, we demonstrate that metoprolol, a β1-adrenergic receptor-selective blocker, could <span class="hlt">induce</span> cardiac fibrosis through a <span class="hlt">G</span> protein-independent and β-arrestin<span class="hlt">2</span>-dependent pathway. Metoprolol, a β-blocker, increased the expression of fibrotic genes responsible for cardiac fibrosis in cardiomyocytes. Furthermore, metoprolol <span class="hlt">induced</span> the interaction between β1-adrenergic receptor and β-arrestin<span class="hlt">2</span>, but not β-arrestin1. The interaction between β1-adrenergic receptor and β-arrestin<span class="hlt">2</span> by metoprolol was impaired in the <span class="hlt">G</span> protein-coupled receptor kinase 5 (GRK5)-knockdown cells. Metoprolol-<span class="hlt">induced</span> cardiac fibrosis led to cardiac dysfunction. However, the metoprolol-<span class="hlt">induced</span> fibrosis and cardiac dysfunction were not evoked in β-arrestin<span class="hlt">2</span>- or GRK5-knock-out mice. Thus, metoprolol is a biased ligand that selectively activates a <span class="hlt">G</span> protein-independent and GRK5/β-arrestin<span class="hlt">2</span>-dependent pathway, and <span class="hlt">induces</span> cardiac fibrosis. This study demonstrates the physiological importance of biased agonism, and suggests that <span class="hlt">G</span> protein-independent and β-arrestin-dependent signaling is a reason for the diversity of the effectiveness of β-blockers. PMID:22888001</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23276591','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23276591"><span>The generation of oxidative stress-<span class="hlt">induced</span> rearrangements in Saccharomyces cerevisiae mtDNA is dependent on the Nuc1 (Endo<span class="hlt">G/ExoG</span>) nuclease and is enhanced by inactivation of the MRX complex.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Dzierzbicki, Piotr; Kaniak-Golik, Aneta; Malc, Ewa; Mieczkowski, Piotr; Ciesla, Zygmunt</p> <p>2012-12-01</p> <p>Oxidative stress is known to enhance the frequency of two major types of alterations in the mitochondrial genome of Saccharomyces cerevisiae: point mutations and large deletions resulting in the generation of respiration-deficient petite rhō mutants. We investigated the effect of antimycin A, a well-known agent <span class="hlt">inducing</span> oxidative stress, on the stability of mtDNA. We show that antimycin enhances exclusively the generation of respiration-deficient petite mutants and this is accompanied by a significant increase in the level of reactive oxygen species (ROS) and in a marked drop of cellular ATP. Whole mitochondrial genome sequencing revealed that mtDNAs of antimycin-<span class="hlt">induced</span> petite mutants are deleted for most of the wild-type sequence and usually contain one of the active origins of mtDNA replication: ori1, ori<span class="hlt">2</span> ori3 or ori5. We show that the frequency of antimycin-<span class="hlt">induced</span> rhō mutants is significantly elevated in mutants deleted either for the RAD50 or XRS<span class="hlt">2</span> gene, both encoding the components of the MRX complex, which is known to be involved in the repair of double strand <span class="hlt">breaks</span> (DSBs) in DNA. Furthermore, enhanced frequency of rhō mutants in cultures of antimycin-treated cells lacking Rad50 was further increased by the simultaneous absence of the Ogg1 glycosylase, an important enzyme functioning in mtBER. We demonstrate also that rad50Δ and xrs<span class="hlt">2</span>Δ deletion mutants display a considerable reduction in the frequency of allelic mitochondrial recombination, suggesting that it is the deficiency in homologous recombination which is responsible for enhanced rearrangements of mtDNA in antimycin-treated cells of these mutants. Finally, we show that the generation of large-scale mtDNA deletions <span class="hlt">induced</span> by antimycin is markedly decreased in a nuc1Δ mutant lacking the activity of the Nuc1 nuclease, an ortholog of the mammalian mitochondrial nucleases Endo<span class="hlt">G</span> and Exo<span class="hlt">G</span>. This result indicates that the nuclease plays an important role in processing of oxidative stress-<span class="hlt">induced</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016cosp...41E.795H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016cosp...41E.795H"><span>Simulated-microgravity <span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span>/M arrest in zebrafish embryonic cell is regulated by dre-miR-22a and its target cep135</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hang, Xiaoming; Sun, Yeqing; Wu, Di; Li, Yixiao; Wang, Ruonan</p> <p>2016-07-01</p> <p>Microgravity has been recognized as a major environmental factor that can <span class="hlt">induce</span> a number of adverse effects such as bone loss, skeletal muscle atrophy, cardiovascular problems and immune system dysregulation, etc. The underlying mechanisms are not absolutely identified yet. Our previous study demonstrated centrosomal protein of 135 kDa (CEP135) might be a microgravity sensitive molecule. In this study, the expression and regulation of CEP135 and its possible roles in cell cycle regulation under simulated microgravity (SMG) condition were investigated. SMG can <span class="hlt">induce</span> significant increasing of cep135 in zebrafish embryos, detected by both in situ hybridization and RT-qPCR, while CEP135 protein level was significantly decreased, tested by western blot. The similar results were also obtained in zebrafish embryonic cells (ZF4) exposed to SMG. Accordingly, the expression level of dre-miR-22a, which might be the potential miRNA for targeting cep135, was significantly increased in SMG exposed ZF4 cells. By combining the results obtained from transfection and dual luciferase reporter assay, we firstly confirmed that dre-miR-22a regulated the expression of cep135 in ZF4 cells. Further investigation on cell cycle demonstrated SMG <span class="hlt">induced</span> a significant arrest in <span class="hlt">G</span><span class="hlt">2</span>/M phase. Transfection of dre-miR-22a also <span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span>/M arrest in ZF4 cells. These results suggest that SMG <span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span>/M arrest in ZF4 cells is via cep135, while dre-miR-22a plays a key role in modulating this effect. Key Words: Simulated-microgravity; cep135; dre-miR-22a; <span class="hlt">G</span><span class="hlt">2</span>/M arrest; zebrafish embryonic cell</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li class="active"><span>19</span></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_19 --> <div id="page_20" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li class="active"><span>20</span></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="381"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4933758','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4933758"><span>Tubeimoside-1 <span class="hlt">induces</span> oxidative stress-mediated apoptosis and <span class="hlt">G</span>0/<span class="hlt">G</span>1 phase arrest in human prostate carcinoma cells in vitro</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Yang, Jing-bo; Khan, Muhammad; He, Yang-yang; Yao, Min; Li, Yong-ming; Gao, Hong-wen; Ma, Tong-hui</p> <p>2016-01-01</p> <p>Aim: Tubeimoside-1 (TBMS1), a triterpenoid saponin extracted from the Chinese herbal medicine Bolbostemma paniculatum (Maxim) Franquet (Cucurbitaceae), has shown anticancer activities in various cancer cell lines. The aim of this study was to investigate the anticancer activity and molecular targets of TBMS1 in human prostate cancer cells in vitro. Methods: DU145 and P3 human prostate cancer cells were treated with TBMS1. Cell viability and apoptosis were detected. ROS generation, mitochondrial membrane potential and cell cycle profile were examined. Western blotting was used to measure the expression of relevant proteins in the cells. Results: TBMS1 (5–100 μmol/L) significantly suppressed the viability of DU145 and P3 cells with IC50 values of approximately 10 and 20 μmol/L, respectively. Furthermore, TBMS1 dose-dependently <span class="hlt">induced</span> apoptosis and cell cycle arrest at <span class="hlt">G</span>0/<span class="hlt">G</span>1 phase in DU145 and P3 cells. In DU145 cells, TBMS1 <span class="hlt">induced</span> mitochondrial apoptosis, evidenced by ROS generation, mitochondrial dysfunction, endoplasmic reticulum stress, modulated Bcl-<span class="hlt">2</span> family protein and cleaved caspase-3, and activated ASK-1 and its downstream targets p38 and JNK. The <span class="hlt">G</span>0/<span class="hlt">G</span>1 phase arrest was linked to increased expression of p53 and p21 and decreased expression of cyclin E and cdk<span class="hlt">2</span>. Co-treatment with Z-VAD-FMK (pan-caspase inhibitor) could attenuate TBMS1-<span class="hlt">induced</span> apoptosis but did not prevent <span class="hlt">G</span>0/<span class="hlt">G</span>1 arrest. Moreover, co-treatment with NAC (ROS scavenger), SB203580 (p38 inhibitor), SP600125 (JNK inhibitor) or salubrinal (ER stress inhibitor) significantly attenuated TBMS1-<span class="hlt">induced</span> apoptosis. Conclusion: TBMS1 <span class="hlt">induces</span> oxidative stress-mediated apoptosis in DU145 human prostate cancer cells in vitro via the mitochondrial pathway. PMID:27292614</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2139799','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2139799"><span>Thrombopoietin-<span class="hlt">induced</span> Polyploidization of Bone Marrow Megakaryocytes Is Due to a Unique Regulatory Mechanism in Late Mitosis</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Nagata, Yuka; Muro, Yoshinao; Todokoro, Kazuo</p> <p>1997-01-01</p> <p>Megakaryocytes undergo a unique differentiation program, becoming polyploid through repeated cycles of DNA synthesis without concomitant cell division. However, the mechanism underlying this polyploidization remains totally unknown. It has been postulated that polyploidization is due to a skipping of mitosis after each round of DNA replication. We carried out immunohistochemical studies on mouse bone marrow megakaryocytes during thrombopoietin- <span class="hlt">induced</span> polyploidization and found that during this process megakaryocytes indeed enter mitosis and progress through normal prophase, prometaphase, metaphase, and up to anaphase A, but not to anaphase B, telophase, or cytokinesis. It was clearly observed that multiple spindle poles were formed as the polyploid megakaryocytes entered mitosis; the nuclear membrane broke down during prophase; the sister <span class="hlt">chromatids</span> were aligned on a multifaced plate, and the centrosomes were symmetrically located on either side of each face of the plate at metaphase; and a set of sister <span class="hlt">chromatids</span> moved into the multiple centrosomes during anaphase A. We further noted that the pair of spindle poles in anaphase were located in close proximity to each other, probably because of the lack of outward movement of spindle poles during anaphase B. Thus, the reassembling nuclear envelope may enclose all the sister <span class="hlt">chromatids</span> in a single nucleus at anaphase and then skip telophase and cytokinesis. These observations clearly indicate that polyploidization of megakaryocytes is not simply due to a skipping of mitosis, and that the megakaryocytes must have a unique regulatory mechanism in anaphase, e.<span class="hlt">g</span>., factors regulating anaphase such as microtubule motor proteins might be involved in this polyploidization process. PMID:9334347</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29183320','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29183320"><span>Fungal 7-epi-10-deacetyltaxol produced by an endophytic Pestalotiopsis microspora <span class="hlt">induces</span> apoptosis in human hepatocellular carcinoma cell line (Hep<span class="hlt">G</span><span class="hlt">2</span>).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Subban, Kamalraj; Singh, Satpal; Subramani, Ramesh; Johnpaul, Muthumary; Chelliah, Jayabaskaran</p> <p>2017-11-28</p> <p>Paclitaxel (taxol) is a potent anticancer drug that is used in the treatment of a wide variety of cancerous. In the present study, we identified a taxol derivative named 7-epi-10-deacetyltaxol (EDT) from the culture of an endophytic fungus Pestalotiopsis microspora isolated from the bark of Taxodium mucronatum. This study was carried out to investigate the effects of fungal EDT on cell proliferation, the induction of apoptosis and the molecular mechanisms of apoptosis in human hepatoma Hep<span class="hlt">G</span><span class="hlt">2</span> cells in vitro. The endophytic fungus was identified by traditional and molecular taxonomical characterization and the fungal EDT was purified using column chromatography and confirmed by various spectroscopic and chromatographic comparisons with authentic paclitaxel. We studied the in vitro effects of EDT on Hep<span class="hlt">G</span><span class="hlt">2</span> cells for parameters such as cell cycle distribution, DNA fragmentation, reactive oxygen species (ROS) generation and nuclear morphology. Further, western blot analysis was used to evaluate Bcl-<span class="hlt">2</span>-associated X protein (Bax), B-cell lymphoma <span class="hlt">2</span> (Bcl-<span class="hlt">2</span>), p38-mitogen activated protein kinase (MAPK) and poly [ADP-ribose] polymerase (PARP) expression. We demonstrate that the fungal EDT exhibited significant in vitro cytotoxicity in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. We investigated cytotoxicity mechanism of EDT in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. The results showed nuclear condensation and DNA fragmentation were observed in cells treated with fungal EDT. Besides, the fungal EDT arrested Hep<span class="hlt">G</span><span class="hlt">2</span> cells at <span class="hlt">G</span><span class="hlt">2</span>/M phase of cell cycle. Furthermore, fungal EDT <span class="hlt">induced</span> apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells in a dose-dependent manner associated with ROS generation and increased Bax/Bcl-<span class="hlt">2</span> ratio, p38 MAPKs and PARP cleavage. Our data show that EDT <span class="hlt">induced</span> apoptotic cell death in Hep<span class="hlt">G</span><span class="hlt">2</span> cells occurs through intrinsic pathway by generation of ROS mediated and activation of MAPK pathway. This is the first report for 7-epi-10-deacetyltaxol (EDT) isolated from a microbial source.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20672644','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20672644"><span>[XPS analysis of beads formed by fuse <span class="hlt">breaking</span> of electric copper wire].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wu, Ying; Meng, Qing-Shan; Wang, Xin-Ming; Gao, Wei; Di, Man</p> <p>2010-05-01</p> <p>The in-depth composition of beads formed by fuse <span class="hlt">breaking</span> of the electric copper wire in different circumstances was studied by XPS with Ar+ ion sputtering. In addition, the measured Auger spectra and the calculated Auger parameters were compared for differentiation of the substances of Cu and Cu<span class="hlt">2</span>O. Corresponding to the sputtering depth, the molten product on a bead <span class="hlt">induced</span> directly by fuse <span class="hlt">breaking</span> of the copper wire without cover may be distinguished as three portions: surface layer with a drastic decrease in carbon content; intermediate layer with a gentle change in oxygen content and gradually diminished carbon peak, and consisting of Cu<span class="hlt">2</span>O; transition layer without Cu<span class="hlt">2</span>O and with a rapid decrease in oxygen content. While the molten product on a bead formed by fuse <span class="hlt">breaking</span> of the copper wire after its insulating cover had been burned out may be distinguished as two portions: surface layer with carbon content decreasing quickly; subsurface layer without Cu<span class="hlt">2</span>O and with carbon and oxygen content decreasing gradually. Thus, it can be seen that there was an obvious interface between the layered surface product and the substrate for the first type of bead, while as to the second type of bead there was no interface. As a result, the presence of Cu<span class="hlt">2</span>O and the quantitative results can be used to identify the molten product on a bead <span class="hlt">induced</span> directly by fuse <span class="hlt">breaking</span> of the copper wire without cover and the molten product on a bead formed by fuse <span class="hlt">breaking</span> of the cupper wire after its insulating cover had been burned out, as a complementary technique for the judgments of fire cause.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JChPh.147h4306K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JChPh.147h4306K"><span>Potential energy and dipole moment surfaces of the triplet states of the O<span class="hlt">2</span>(X3Σ<span class="hlt">g</span>-) - O<span class="hlt">2</span>(X3Σ<span class="hlt">g</span>-,a1Δ<span class="hlt">g</span>,b1Σ<span class="hlt">g</span>+) complex</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Karman, Tijs; van der Avoird, Ad; Groenenboom, Gerrit C.</p> <p>2017-08-01</p> <p>We compute four-dimensional diabatic potential energy surfaces and transition dipole moment surfaces of O<span class="hlt">2</span>-O<span class="hlt">2</span>, relevant for the theoretical description of collision-<span class="hlt">induced</span> absorption in the forbidden X3Σ<span class="hlt">g</span>- → a1Δ<span class="hlt">g</span> and X3Σ<span class="hlt">g</span>- → b1Σ<span class="hlt">g</span>+ bands at 7883 cm-1 and 13 122 cm-1, respectively. We compute potentials at the multi-reference configuration interaction (MRCI) level and dipole surfaces at the MRCI and complete active space self-consistent field (CASSCF) levels of theory. Potentials and dipole surfaces are transformed to a diabatic basis using a recent multiple-property-based diabatization algorithm. We discuss the angular expansion of these surfaces, derive the symmetry constraints on the expansion coefficients, and present working equations for determining the expansion coefficients by numerical integration over the angles. We also present an interpolation scheme with exponential extrapolation to both short and large separations, which is used for representing the O<span class="hlt">2</span>-O<span class="hlt">2</span> distance dependence of the angular expansion coefficients. For the triplet ground state of the complex, the potential energy surface is in reasonable agreement with previous calculations, whereas global excited state potentials are reported here for the first time. The transition dipole moment surfaces are strongly dependent on the level of theory at which they are calculated, as is also shown here by benchmark calculations at high symmetry geometries. Therefore, ab initio calculations of the collision-<span class="hlt">induced</span> absorption spectra cannot become quantitatively predictive unless more accurate transition dipole surfaces can be computed. This is left as an open question for method development in electronic structure theory. The calculated potential energy and transition dipole moment surfaces are employed in quantum dynamical calculations of collision-<span class="hlt">induced</span> absorption spectra reported in Paper II [T. Karman et al., J. Chem. Phys. 147, 084307 (2017)].</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28863529','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28863529"><span>Potential energy and dipole moment surfaces of the triplet states of the O<span class="hlt">2</span>(X3Σ<span class="hlt">g</span>-) - O<span class="hlt">2</span>(X3Σ<span class="hlt">g</span>-,a1Δ<span class="hlt">g</span>,b1Σ<span class="hlt">g</span>+) complex.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Karman, Tijs; van der Avoird, Ad; Groenenboom, Gerrit C</p> <p>2017-08-28</p> <p>We compute four-dimensional diabatic potential energy surfaces and transition dipole moment surfaces of O <span class="hlt">2</span> -O <span class="hlt">2</span> , relevant for the theoretical description of collision-<span class="hlt">induced</span> absorption in the forbidden X 3 Σ <span class="hlt">g</span> -  → a 1 Δ <span class="hlt">g</span> and X 3 Σ <span class="hlt">g</span> -  → b 1 Σ <span class="hlt">g</span> + bands at 7883 cm -1 and 13 122 cm -1 , respectively. We compute potentials at the multi-reference configuration interaction (MRCI) level and dipole surfaces at the MRCI and complete active space self-consistent field (CASSCF) levels of theory. Potentials and dipole surfaces are transformed to a diabatic basis using a recent multiple-property-based diabatization algorithm. We discuss the angular expansion of these surfaces, derive the symmetry constraints on the expansion coefficients, and present working equations for determining the expansion coefficients by numerical integration over the angles. We also present an interpolation scheme with exponential extrapolation to both short and large separations, which is used for representing the O <span class="hlt">2</span> -O <span class="hlt">2</span> distance dependence of the angular expansion coefficients. For the triplet ground state of the complex, the potential energy surface is in reasonable agreement with previous calculations, whereas global excited state potentials are reported here for the first time. The transition dipole moment surfaces are strongly dependent on the level of theory at which they are calculated, as is also shown here by benchmark calculations at high symmetry geometries. Therefore, ab initio calculations of the collision-<span class="hlt">induced</span> absorption spectra cannot become quantitatively predictive unless more accurate transition dipole surfaces can be computed. This is left as an open question for method development in electronic structure theory. The calculated potential energy and transition dipole moment surfaces are employed in quantum dynamical calculations of collision-<span class="hlt">induced</span> absorption spectra reported in Paper II [T. Karman et al., J. Chem. Phys. 147, 084307 (2017)].</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19910018758&hterms=anemia&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Danemia','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19910018758&hterms=anemia&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Danemia"><span>X ray sensitivity of diploid skin fibroblasts from patients with Fanconi's anemia</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kale, Ranjini</p> <p>1989-01-01</p> <p>Experiments were performed on Fanconi's anemia and normal human fibroblast cell lines growing in culture in an attempt to correlate cell cycle kinetics with genomic damage and determine their bearing on the mechanism of chromosome aberration induction. FA fibroblasts showed a significantly increased susceptibility to chromosomal breakage by x rays in the <span class="hlt">G</span><span class="hlt">2</span> phase of the cell cycle. No such response was observed in fibroblasts irradiated in the <span class="hlt">G</span>0 phase. The observed increases in achromatic lesions and in <span class="hlt">chromatid</span> deletions in FA cells as compared with normal cells appear to indicate that FA cells are deficient in strand <span class="hlt">break</span> repair and also possibly in base damage excision repair. Experiments are now in progress to further elucidate the mechanisms involved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3826130','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3826130"><span>Microcystin-LR and Cylindrospermopsin <span class="hlt">Induced</span> Alterations in Chromatin Organization of Plant Cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Máthé, Csaba; M-Hamvas, Márta; Vasas, Gábor</p> <p>2013-01-01</p> <p>Cyanobacteria produce metabolites with diverse bioactivities, structures and pharmacological properties. The effects of microcystins (MCYs), a family of peptide type protein-phosphatase inhibitors and cylindrospermopsin (CYN), an alkaloid type of protein synthesis blocker will be discussed in this review. We are focusing mainly on cyanotoxin-<span class="hlt">induced</span> changes of chromatin organization and their possible cellular mechanisms. The particularities of plant cells explain the importance of such studies. Preprophase bands (PPBs) are premitotic cytoskeletal structures important in the determination of plant cell division plane. Phragmoplasts are cytoskeletal structures involved in plant cytokinesis. Both cyanotoxins <span class="hlt">induce</span> the formation of multipolar spindles and disrupted phragmoplasts, leading to abnormal sister <span class="hlt">chromatid</span> segregation during mitosis. Thus, MCY and CYN are probably <span class="hlt">inducing</span> alterations of chromosome number. MCY <span class="hlt">induces</span> programmed cell death: chromatin condensation, nucleus fragmentation, necrosis, alterations of nuclease and protease enzyme activities and patterns. The above effects may be related to elevated reactive oxygen species (ROS) and/or disfunctioning of microtubule associated proteins. Specific effects: MCY-LR <span class="hlt">induces</span> histone H3 hyperphosphorylation leading to incomplete <span class="hlt">chromatid</span> segregation and the formation of micronuclei. CYN <span class="hlt">induces</span> the formation of split or double PPB directly related to protein synthesis inhibition. Cyanotoxins are powerful tools in the study of plant cell organization. PMID:24084787</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4252111','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4252111"><span>Restriction Endonucleases from Invasive Neisseria gonorrhoeae Cause Double-Strand <span class="hlt">Breaks</span> and Distort Mitosis in Epithelial Cells during Infection</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Weyler, Linda; Engelbrecht, Mattias; Mata Forsberg, Manuel; Brehwens, Karl; Vare, Daniel; Vielfort, Katarina; Wojcik, Andrzej; Aro, Helena</p> <p>2014-01-01</p> <p>The host epithelium is both a barrier against, and the target for microbial infections. Maintaining regulated cell growth ensures an intact protective layer towards microbial-<span class="hlt">induced</span> cellular damage. Neisseria gonorrhoeae infections disrupt host cell cycle regulation machinery and the infection causes DNA double strand <span class="hlt">breaks</span> that delay progression through the <span class="hlt">G</span><span class="hlt">2</span>/M phase. We show that intracellular gonococci upregulate and release restriction endonucleases that enter the nucleus and damage human chromosomal DNA. Bacterial lysates containing restriction endonucleases were able to fragment genomic DNA as detected by PFGE. Lysates were also microinjected into the cytoplasm of cells in interphase and after 20 h, DNA double strand <span class="hlt">breaks</span> were identified by 53BP1 staining. In addition, by using live-cell microscopy and NHS-ester stained live gonococci we visualized the subcellular location of the bacteria upon mitosis. Infected cells show dysregulation of the spindle assembly checkpoint proteins MAD1 and MAD<span class="hlt">2</span>, impaired and prolonged M-phase, nuclear swelling, micronuclei formation and chromosomal instability. These data highlight basic molecular functions of how gonococcal infections affect host cell cycle regulation, cause DNA double strand <span class="hlt">breaks</span> and predispose cellular malignancies. PMID:25460012</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25460012','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25460012"><span>Restriction endonucleases from invasive Neisseria gonorrhoeae cause double-strand <span class="hlt">breaks</span> and distort mitosis in epithelial cells during infection.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Weyler, Linda; Engelbrecht, Mattias; Mata Forsberg, Manuel; Brehwens, Karl; Vare, Daniel; Vielfort, Katarina; Wojcik, Andrzej; Aro, Helena</p> <p>2014-01-01</p> <p>The host epithelium is both a barrier against, and the target for microbial infections. Maintaining regulated cell growth ensures an intact protective layer towards microbial-<span class="hlt">induced</span> cellular damage. Neisseria gonorrhoeae infections disrupt host cell cycle regulation machinery and the infection causes DNA double strand <span class="hlt">breaks</span> that delay progression through the <span class="hlt">G</span><span class="hlt">2</span>/M phase. We show that intracellular gonococci upregulate and release restriction endonucleases that enter the nucleus and damage human chromosomal DNA. Bacterial lysates containing restriction endonucleases were able to fragment genomic DNA as detected by PFGE. Lysates were also microinjected into the cytoplasm of cells in interphase and after 20 h, DNA double strand <span class="hlt">breaks</span> were identified by 53BP1 staining. In addition, by using live-cell microscopy and NHS-ester stained live gonococci we visualized the subcellular location of the bacteria upon mitosis. Infected cells show dysregulation of the spindle assembly checkpoint proteins MAD1 and MAD<span class="hlt">2</span>, impaired and prolonged M-phase, nuclear swelling, micronuclei formation and chromosomal instability. These data highlight basic molecular functions of how gonococcal infections affect host cell cycle regulation, cause DNA double strand <span class="hlt">breaks</span> and predispose cellular malignancies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5703705','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5703705"><span>Immunoglobulins <span class="hlt">G</span> from Sera of Amyotrophic Lateral Sclerosis Patients <span class="hlt">Induce</span> Oxidative Stress and Upregulation of Antioxidative System in BV-<span class="hlt">2</span> Microglial Cell Line</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Milošević, Milena; Milićević, Katarina; Božić, Iva; Lavrnja, Irena; Stevanović, Ivana; Bijelić, Dunja; Dubaić, Marija; Živković, Irena; Stević, Zorica; Giniatullin, Rashid; Andjus, Pavle</p> <p>2017-01-01</p> <p>Amyotrophic lateral sclerosis (ALS) is a neurodegenerative disorder with a very fast progression, no diagnostic tool for the presymptomatic phase, and still no effective treatment of the disease. Although ALS affects motor neurons, the overall pathophysiological condition points out to the non-cell autonomous mechanisms, where astrocytes and microglia play crucial roles in the disease progression. We have already shown that Ig<span class="hlt">G</span> from sera of ALS patients (ALS Ig<span class="hlt">G</span>) <span class="hlt">induce</span> calcium transients and an increase in the mobility of acidic vesicles in cultured rat astrocytes. Having in mind the role of microglia in neurodegeneration, and a well-documented fact that oxidative stress is one of the many components contributing to the disease, we decided to examine the effect of ALS Ig<span class="hlt">G</span> on activation, oxidative stress and antioxidative system of BV-<span class="hlt">2</span> microglia, and to evaluate their acute effect on cytosolic peroxide, pH, and on reactive oxygen species (ROS) generation. All tested ALS IgGs (compared to control Ig<span class="hlt">G</span>) <span class="hlt">induced</span> oxidative stress (rise in nitric oxide and the index of lipid peroxidation) followed by release of TNF-α and higher antioxidative defense (elevation of Mn- and CuZn-superoxide dismutase, catalase, and glutathione reductase with a decrease of glutathione peroxidase and glutathione) after 24 h treatment. Both ALS Ig<span class="hlt">G</span> and control Ig<span class="hlt">G</span> showed same localization on the membrane of BV-<span class="hlt">2</span> cells following 24 h treatment. Cytosolic peroxide and pH alteration were evaluated with fluorescent probes HyPer and SypHer, respectively, having in mind that HyPer also reacts to pH changes. Out of 11 tested IgGs from ALS patients, 4 <span class="hlt">induced</span> slow exponential rise of HyPer signal, with maximal normalized fluorescence in the range 0.<span class="hlt">2</span>–0.5, also <span class="hlt">inducing</span> similar increase of SypHer intensity, but of a lower amplitude. None of the control IgGs <span class="hlt">induced</span> changes with neither of the indicators. Acute ROS generation was detected in one out of three tested ALS samples with carboxy-H<span class="hlt">2</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29628444','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29628444"><span>PH domain leucine-rich repeat protein phosphatase <span class="hlt">2</span> (PHLPP<span class="hlt">2</span>) regulates <span class="hlt">G</span>-protein-coupled receptor kinase 5 (GRK5)-<span class="hlt">induced</span> cardiac hypertrophy in vitro.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yeh, Szu-Tsen; Zambrano, Cristina M; Koch, Walter J; Purcell, Nicole H</p> <p>2018-05-25</p> <p>PH domain leucine-rich repeat protein phosphatase (PHLPP) is a serine/threonine phosphatase that has been shown to regulate cell growth and survival through dephosphorylation of several members of the AGC family of kinases. <span class="hlt">G</span>-protein-coupled receptor kinase 5 (GRK5) is an AGC kinase that regulates phenylephrine (PE)-<span class="hlt">induced</span> cardiac hypertrophy through its noncanonical function of directly targeting proteins to the nucleus to regulate transcription. Here we investigated the possibility that the PHLPP<span class="hlt">2</span> isoform can regulate GRK5-<span class="hlt">induced</span> cardiomyocyte hypertrophy in neonatal rat ventricular myocytes (NRVMs). We show that removal of PHLPP<span class="hlt">2</span> by siRNA <span class="hlt">induces</span> hypertrophic growth of NRVMs as measured by cell size changes at baseline, potentiated PE-<span class="hlt">induced</span> cell size changes, and re-expression of fetal genes atrial natriuretic factor and brain natriuretic peptide. Endogenous GRK5 and PHLPP<span class="hlt">2</span> were found to interact in NRVMs, and PE-<span class="hlt">induced</span> nuclear accumulation of GRK5 was enhanced upon down-regulation of PHLPP<span class="hlt">2</span>. Conversely, overexpression of PHLPP<span class="hlt">2</span> blocked PE-<span class="hlt">induced</span> hypertrophic growth, re-expression of fetal genes, and nuclear accumulation of GRK5, which depended on its phosphatase activity. Finally, using siRNA against GRK5, we found that GRK5 was necessary for the hypertrophic response <span class="hlt">induced</span> by PHLPP<span class="hlt">2</span> knockdown. Our findings demonstrate for the first time a novel regulation of GRK5 by the phosphatase PHLPP<span class="hlt">2</span>, which modulates hypertrophic growth. Understanding the signaling pathways affected by PHLPP<span class="hlt">2</span> has potential for new therapeutic targets in the treatment of cardiac hypertrophy and failure. © 2018 by The American Society for Biochemistry and Molecular Biology, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24953651','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24953651"><span>MOF phosphorylation by ATM regulates 53BP1-mediated double-strand <span class="hlt">break</span> repair pathway choice.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gupta, Arun; Hunt, Clayton R; Hegde, Muralidhar L; Chakraborty, Sharmistha; Chakraborty, Sharmistha; Udayakumar, Durga; Horikoshi, Nobuo; Singh, Mayank; Ramnarain, Deepti B; Hittelman, Walter N; Namjoshi, Sarita; Asaithamby, Aroumougame; Hazra, Tapas K; Ludwig, Thomas; Pandita, Raj K; Tyler, Jessica K; Pandita, Tej K</p> <p>2014-07-10</p> <p>Cell-cycle phase is a critical determinant of the choice between DNA damage repair by nonhomologous end-joining (NHEJ) or homologous recombination (HR). Here, we report that double-strand <span class="hlt">breaks</span> (DSBs) <span class="hlt">induce</span> ATM-dependent MOF (a histone H4 acetyl-transferase) phosphorylation (p-T392-MOF) and that phosphorylated MOF colocalizes with γ-H<span class="hlt">2</span>AX, ATM, and 53BP1 foci. Mutation of the phosphorylation site (MOF-T392A) impedes DNA repair in S and <span class="hlt">G</span><span class="hlt">2</span> phase but not <span class="hlt">G</span>1 phase cells. Expression of MOF-T392A also blocks the reduction in DSB-associated 53BP1 seen in wild-type S/<span class="hlt">G</span><span class="hlt">2</span> phase cells, resulting in enhanced 53BP1 and reduced BRCA1 association. Decreased BRCA1 levels at DSB sites correlates with defective repairosome formation, reduced HR repair, and decreased cell survival following irradiation. These data support a model whereby ATM-mediated MOF-T392 phosphorylation modulates 53BP1 function to facilitate the subsequent recruitment of HR repair proteins, uncovering a regulatory role for MOF in DSB repair pathway choice during S/<span class="hlt">G</span><span class="hlt">2</span> phase. Copyright © 2014 The Authors. Published by Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15488639','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15488639"><span>Physalis peruviana extract <span class="hlt">induces</span> apoptosis in human Hep <span class="hlt">G</span><span class="hlt">2</span> cells through CD95/CD95L system and the mitochondrial signaling transduction pathway.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wu, Shu-Jing; Ng, Lean-Teik; Lin, Doung-Liang; Huang, Shan-Ney; Wang, Shyh-Shyan; Lin, Chun-Ching</p> <p>2004-11-25</p> <p>Physalis species is a popular folk medicine used for treating cancer, leukemia, hepatitis and other diseases. Studies have shown that the ethanol extract of Physalis peruviana (EEPP) inhibits growth and <span class="hlt">induces</span> apoptotic death of human Hep <span class="hlt">G</span><span class="hlt">2</span> cells in culture, whereas proliferation of the mouse BALB/C normal liver cells was not affected. In this study, we performed detailed studies to define the molecular mechanism of EEPP-<span class="hlt">induced</span> apoptosis in Hep <span class="hlt">G</span><span class="hlt">2</span> cells. The results further confirmed that EEPP inhibited cell proliferation in a dose- and time-dependent manner. At 50 microg/ml, EEPP significantly increased the accumulation of the sub-<span class="hlt">G</span>1 peak (hypoploid) and the portion of apoptotic annexin V positive cells. EEPP was found to trigger apoptosis through the release of cytochrome c, Smac/DIABLO and Omi/HtrA<span class="hlt">2</span> from mitochondria to cytosol and consequently resulted in caspase-3 activation. Pre-treatment with a general caspase inhibitor (z-VAD-fmk) prevented cytochrome c release. After 48 h of EEPP treatment, the apoptosis of Hep <span class="hlt">G</span><span class="hlt">2</span> cells was found to associate with an elevated p53, and CD95 and CD95L proteins expression. Furthermore, a marked down-regulation of the expression of the Bcl-<span class="hlt">2</span>, Bcl-XL and XIAP, and up-regulation of the Bax and Bad proteins were noted. Taken together, the present results suggest that EEPP-<span class="hlt">induced</span> Hep <span class="hlt">G</span><span class="hlt">2</span> cell apoptosis was possibly mediated through the CD95/CD95L system and the mitochondrial signaling transduction pathway.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25518891','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25518891"><span>Rice bran protein hydrolysates prevented interleukin-6- and high glucose-<span class="hlt">induced</span> insulin resistance in Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Boonloh, Kampeebhorn; Kukongviriyapan, Upa; Pannangpetch, Patchareewan; Kongyingyoes, Bunkerd; Senggunprai, Laddawan; Prawan, Auemduan; Thawornchinsombut, Supawan; Kukongviriyapan, Veerapol</p> <p>2015-02-01</p> <p>Rice bran, which is a byproduct of rice milling process, contains various nutrients and biologically active compounds. Rice bran protein hydrolysates have various pharmacological activities such as antidiabetic and antidyslipidemic effects. However, there are limited studies about the mechanisms of rice bran protein hydrolysates (RBP) on insulin resistance and lipid metabolism. RBP used in this study were prepared from Thai Jasmine rice. When Hep<span class="hlt">G</span><span class="hlt">2</span> cells were treated with IL-6, the IRS-1 expression and Akt phosphorylation were suppressed. This effect of IL-6 was prevented by RBP in association with inhibition of STAT3 phosphorylation and SOCS3 expression. RBP could increase the phospho-AMPK levels and inhibit IL-6- or high glucose-<span class="hlt">induced</span> suppression of AMPK and Akt activation. High glucose-<span class="hlt">induced</span> dysregulation of the expression of lipogenic genes, including SREBP-1c, FASN and CPT-1, was normalized by RBP treatment. Moreover, impaired glucose utilization in insulin resistant Hep<span class="hlt">G</span><span class="hlt">2</span> cells was significantly alleviated by concurrent treatment with RBP. Our results suggested that RBP suppresses inflammatory cytokine signaling and activates AMPK, and thereby these effects may underlie the insulin sensitizing effect.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28639221','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28639221"><span>HBV X Protein <span class="hlt">induces</span> overexpression of HERV-W env through NF-κB in Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liu, Cong; Liu, Lijuan; Wang, Xiuling; Liu, Youyi; Wang, Miao; Zhu, Fan</p> <p>2017-12-01</p> <p>Human endogenous retrovirus W family (HERV-W) envelope (env) at chromosome 7 is highly expressed in the placenta and possesses fusogenic activity in trophoblast development. HERV-W env has been found to be overexpressed in some cancers and immune diseases. Viral transactivators can <span class="hlt">induce</span> the overexpression of HERV-W env in human cell lines. Hepatitis B virus X protein (HBx) is believed to be a multifunctional oncogenic protein. Here, we reported that HBx could increase the promoter activity of HERV-W env and upregulate the mRNA levels of non-spliced and spliced HERV-W env and also its protein in human hepatoma Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Interestingly, we found that the inhibition of nuclear factor κB (NF-κB) using shRNA targeting NF-κB/p65 or PDTC (an inhibitor of NF-κB) could attenuate the upregulation of HERV-W env <span class="hlt">induced</span> by HBx. These suggested that HBx might upregulate the expression of HERV-W env through NF-κB in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. This study might provide a new insight in HBV-associated liver diseases including HCC.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28373429','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28373429"><span>DNA Double-strand <span class="hlt">Breaks</span> <span class="hlt">Induced</span> byFractionated Neutron Beam Irradiation for Boron Neutron Capture Therapy.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kinashi, Yuko; Yokomizo, Natsuya; Takahashi, Sentaro</p> <p>2017-04-01</p> <p>To use the 53BP1 foci assay to detect DNA double-strand <span class="hlt">breaks</span> <span class="hlt">induced</span> by fractionated neutron beam irradiation of normal cells. The Kyoto University Research Reactor heavy-water facility and gamma-ray irradiation system were used as experimental radiation sources. After fixation of Chinese Hamster Ovary cells with 3.6% formalin, immunofluorescence staining was performed. Number and size of foci were analyzed using ImageJ software. Fractionated neutron irradiation <span class="hlt">induced</span> 25% fewer 53BP1 foci than single irradiation at the same dose. By contrast, gamma irradiation <span class="hlt">induced</span> 30% fewer 53BP1 foci than single irradiation at the same dose. Fractionated neutron irradiation <span class="hlt">induced</span> larger foci than gamma irradiation, raising the possibility that persistent unrepaired DNA damage was amplified due to the high linear energy transfer component in the neutron beam. Unrepaired cluster DNA damage was more prevalent after fractionated neutron irradiation than after gamma irradiation. Copyright© 2017, International Institute of Anticancer Research (Dr. George J. Delinasios), All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25320297','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25320297"><span>Topical herpes simplex virus <span class="hlt">2</span> (HSV-<span class="hlt">2</span>) vaccination with human papillomavirus vectors expressing <span class="hlt">gB/g</span>D ectodomains <span class="hlt">induces</span> genital-tissue-resident memory CD8+ T cells and reduces genital disease and viral shedding after HSV-<span class="hlt">2</span> challenge.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Çuburu, Nicolas; Wang, Kening; Goodman, Kyle N; Pang, Yuk Ying; Thompson, Cynthia D; Lowy, Douglas R; Cohen, Jeffrey I; Schiller, John T</p> <p>2015-01-01</p> <p>No herpes simplex virus <span class="hlt">2</span> (HSV-<span class="hlt">2</span>) vaccine has been licensed for use in humans. HSV-<span class="hlt">2</span> glycoproteins B (<span class="hlt">g</span>B) and D (<span class="hlt">g</span>D) are targets of neutralizing antibodies and T cells, but clinical trials involving intramuscular (i.m.) injection of HSV-<span class="hlt">2</span> <span class="hlt">g</span>B and <span class="hlt">g</span>D in adjuvants have not been effective. Here we evaluated intravaginal (ivag) genetic immunization of C57BL/6 mice with a replication-defective human papillomavirus pseudovirus (HPV PsV) expressing HSV-<span class="hlt">2</span> <span class="hlt">g</span>B (HPV-<span class="hlt">g</span>B) or <span class="hlt">g</span>D (HPV-<span class="hlt">g</span>D) constructs to target different subcellular compartments. HPV PsV expressing a secreted ectodomain of <span class="hlt">g</span>B (<span class="hlt">g</span>Bsec) or <span class="hlt">g</span>D (<span class="hlt">g</span>Dsec), but not PsV expressing a cytoplasmic or membrane-bound form, <span class="hlt">induced</span> circulating and intravaginal-tissue-resident memory CD8(+) T cells that were able to secrete gamma interferon (IFN-γ) and tumor necrosis factor alpha (TNF-α) as well as moderate levels of serum HSV neutralizing antibodies. Combined immunization with HPV-<span class="hlt">g</span>Bsec and HPV-<span class="hlt">g</span>Dsec (HPV-<span class="hlt">gBsec/g</span>Dsec) vaccines conferred longer survival after vaginal challenge with HSV-<span class="hlt">2</span> than immunization with HPV-<span class="hlt">g</span>Bsec or HPV-<span class="hlt">g</span>Dsec alone. HPV-<span class="hlt">gBsec/g</span>Dsec ivag vaccination was associated with a reduced severity of genital lesions and lower levels of viral shedding in the genital tract after HSV-<span class="hlt">2</span> challenge. In contrast, intramuscular vaccination with a soluble truncated <span class="hlt">g</span>D protein (<span class="hlt">g</span>D<span class="hlt">2</span>t) in alum and monophosphoryl lipid A (MPL) elicited high neutralizing antibody titers and improved survival but did not reduce genital lesions and viral shedding. Vaccination combining ivag HPV-<span class="hlt">gBsec/g</span>Dsec and i.m. <span class="hlt">g</span>D<span class="hlt">2</span>t-alum-MPL improved survival and reduced genital lesions and viral shedding. Finally, high levels of circulating HSV-<span class="hlt">2</span>-specific CD8(+) T cells, but not serum antibodies, correlated with reduced viral shedding. Taken together, our data underscore the potential of HPV PsV as a platform for a topical mucosal vaccine to control local manifestations of primary HSV-<span class="hlt">2</span> infection. Genital herpes is a highly prevalent chronic disease caused by</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4301134','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4301134"><span>Topical Herpes Simplex Virus <span class="hlt">2</span> (HSV-<span class="hlt">2</span>) Vaccination with Human Papillomavirus Vectors Expressing <span class="hlt">gB/g</span>D Ectodomains <span class="hlt">Induces</span> Genital-Tissue-Resident Memory CD8+ T Cells and Reduces Genital Disease and Viral Shedding after HSV-<span class="hlt">2</span> Challenge</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Çuburu, Nicolas; Wang, Kening; Goodman, Kyle N.; Pang, Yuk Ying; Thompson, Cynthia D.; Lowy, Douglas R.; Cohen, Jeffrey I.</p> <p>2014-01-01</p> <p>ABSTRACT No herpes simplex virus <span class="hlt">2</span> (HSV-<span class="hlt">2</span>) vaccine has been licensed for use in humans. HSV-<span class="hlt">2</span> glycoproteins B (<span class="hlt">g</span>B) and D (<span class="hlt">g</span>D) are targets of neutralizing antibodies and T cells, but clinical trials involving intramuscular (i.m.) injection of HSV-<span class="hlt">2</span> <span class="hlt">g</span>B and <span class="hlt">g</span>D in adjuvants have not been effective. Here we evaluated intravaginal (ivag) genetic immunization of C57BL/6 mice with a replication-defective human papillomavirus pseudovirus (HPV PsV) expressing HSV-<span class="hlt">2</span> <span class="hlt">g</span>B (HPV-<span class="hlt">g</span>B) or <span class="hlt">g</span>D (HPV-<span class="hlt">g</span>D) constructs to target different subcellular compartments. HPV PsV expressing a secreted ectodomain of <span class="hlt">g</span>B (<span class="hlt">g</span>Bsec) or <span class="hlt">g</span>D (<span class="hlt">g</span>Dsec), but not PsV expressing a cytoplasmic or membrane-bound form, <span class="hlt">induced</span> circulating and intravaginal-tissue-resident memory CD8+ T cells that were able to secrete gamma interferon (IFN-γ) and tumor necrosis factor alpha (TNF-α) as well as moderate levels of serum HSV neutralizing antibodies. Combined immunization with HPV-<span class="hlt">g</span>Bsec and HPV-<span class="hlt">g</span>Dsec (HPV-<span class="hlt">gBsec/g</span>Dsec) vaccines conferred longer survival after vaginal challenge with HSV-<span class="hlt">2</span> than immunization with HPV-<span class="hlt">g</span>Bsec or HPV-<span class="hlt">g</span>Dsec alone. HPV-<span class="hlt">gBsec/g</span>Dsec ivag vaccination was associated with a reduced severity of genital lesions and lower levels of viral shedding in the genital tract after HSV-<span class="hlt">2</span> challenge. In contrast, intramuscular vaccination with a soluble truncated <span class="hlt">g</span>D protein (<span class="hlt">g</span>D<span class="hlt">2</span>t) in alum and monophosphoryl lipid A (MPL) elicited high neutralizing antibody titers and improved survival but did not reduce genital lesions and viral shedding. Vaccination combining ivag HPV-<span class="hlt">gBsec/g</span>Dsec and i.m. <span class="hlt">g</span>D<span class="hlt">2</span>t-alum-MPL improved survival and reduced genital lesions and viral shedding. Finally, high levels of circulating HSV-<span class="hlt">2</span>-specific CD8+ T cells, but not serum antibodies, correlated with reduced viral shedding. Taken together, our data underscore the potential of HPV PsV as a platform for a topical mucosal vaccine to control local manifestations of primary HSV-<span class="hlt">2</span> infection. IMPORTANCE Genital herpes is a highly prevalent chronic</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=109829','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=109829"><span>Rotavirus <span class="hlt">2</span>/6 Viruslike Particles Administered Intranasally with Cholera Toxin, Escherichia coli Heat-Labile Toxin (LT), and LT-R192<span class="hlt">G</span> <span class="hlt">Induce</span> Protection from Rotavirus Challenge</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>O’Neal, Christine M.; Clements, John D.; Estes, Mary K.; Conner, Margaret E.</p> <p>1998-01-01</p> <p>We have shown that rotavirus <span class="hlt">2</span>/6 viruslike particles composed of proteins VP<span class="hlt">2</span> and VP6 (<span class="hlt">2</span>/6-VLPs) administered to mice intranasally with cholera toxin (CT) <span class="hlt">induced</span> protection from rotavirus challenge, as measured by virus shedding. Since it is unclear if CT will be approved for human use, we evaluated the adjuvanticity of Escherichia coli heat-labile toxin (LT) and LT-R192<span class="hlt">G</span>. Mice were inoculated intranasally with 10 μ<span class="hlt">g</span> of <span class="hlt">2</span>/6-VLPs combined with CT, LT, or LT-R192<span class="hlt">G</span>. All three adjuvants <span class="hlt">induced</span> equivalent geometric mean titers of rotavirus-specific serum antibody and intestinal immunoglobulin <span class="hlt">G</span> (Ig<span class="hlt">G</span>). Mice inoculated with <span class="hlt">2</span>/6-VLPs with LT produced significantly higher titers of intestinal IgA than mice given CT as the adjuvant. All mice inoculated with <span class="hlt">2</span>/6-VLPs mixed with LT and LT-R192<span class="hlt">G</span> were totally protected (100%) from rotavirus challenge, while mice inoculated with <span class="hlt">2</span>/6-VLPs mixed with CT showed a mean 91% protection from challenge. The availability of a safe, effective mucosal adjuvant such as LT-R192<span class="hlt">G</span> will increase the practicality of administering recombinant vaccines mucosally. PMID:9525668</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li class="active"><span>20</span></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_20 --> <div id="page_21" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li class="active"><span>21</span></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="401"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5615224','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5615224"><span>Cnidium officinale Makino extract <span class="hlt">induces</span> apoptosis through activation of caspase-3 and p53 in human liver cancer Hep<span class="hlt">G</span><span class="hlt">2</span> cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Hong, Heeok; An, Jeong Cheol; de La Cruz, Joseph F.; Hwang, Seong-Gu</p> <p>2017-01-01</p> <p>A number of diverse studies have reported the anticancer properties of Cnidium officinale Makino (CO). However, the apoptotic effect of this traditional medicinal herb in human hepatocellular carcinoma cells (Hep<span class="hlt">G</span><span class="hlt">2</span>) remains to be elucidated. Therefore, the present study investigated the ability of CO to reduce cell viability through apoptotic pathways. Cell viability was determined using the <span class="hlt">2</span>,3-bis [<span class="hlt">2</span>-methyloxy-4-nitro-5-sulfophenyl]-<span class="hlt">2</span>H-tetrazolium-5-carboxanilide assay. CO extract-<span class="hlt">induced</span> apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells was assessed by Hoechst 33258 staining. The cell cycle was monitored using fluorescence-activated cell sorting analysis with propidium iodide staining. Furthermore, the present study explored whether various signaling molecules associated with Hep<span class="hlt">G</span><span class="hlt">2</span> cell death were affected by CO treatment, including caspase-3, B-cell lymphoma <span class="hlt">2</span> (Bcl-<span class="hlt">2</span>), tumor protein p53 (p53), cyclin-dependent kinase 4 (CDK4) and cyclin D. The expression levels of these genes were examined by reverse-transcription polymerase chain reaction and western blotting. The expression levels of caspase-3 and p53 were upregulated with CO extract treatment, whereas those of Bcl-<span class="hlt">2</span>, CDK4 and cyclin D were significantly downregulated. Cleaved caspase-3 expression was upregulated following treatment with CO extract in a dose-dependent manner. Collectively, the data suggest that CO extract has the potential to <span class="hlt">induce</span> apoptosis of Hep<span class="hlt">G</span><span class="hlt">2</span> cells and may act by suppressing the cell cycle, which leads to caspase-3 cleavage and p53 signaling. PMID:28966688</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=33412','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=33412"><span>Agonist-<span class="hlt">induced</span> conformational changes in the <span class="hlt">G</span>-protein-coupling domain of the β<span class="hlt">2</span> adrenergic receptor</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Ghanouni, Pejman; Steenhuis, Jacqueline J.; Farrens, David L.; Kobilka, Brian K.</p> <p>2001-01-01</p> <p>The majority of extracellular physiologic signaling molecules act by stimulating GTP-binding protein (<span class="hlt">G</span>-protein)-coupled receptors (GPCRs). To monitor directly the formation of the active state of a prototypical GPCR, we devised a method to site specifically attach fluorescein to an endogenous cysteine (Cys-265) at the cytoplasmic end of transmembrane 6 (TM6) of the β<span class="hlt">2</span> adrenergic receptor (β<span class="hlt">2</span>AR), adjacent to the <span class="hlt">G</span>-protein-coupling domain. We demonstrate that this tag reports agonist-<span class="hlt">induced</span> conformational changes in the receptor, with agonists causing a decline in the fluorescence intensity of fluorescein-β<span class="hlt">2</span>AR that is proportional to the biological efficacy of the agonist. We also find that agonists alter the interaction between the fluorescein at Cys-265 and fluorescence-quenching reagents localized to different molecular environments of the receptor. These observations are consistent with a rotation and/or tilting of TM6 on agonist activation. Our studies, when compared with studies of activation in rhodopsin, indicate a general mechanism for GPCR activation; however, a notable difference is the relatively slow kinetics of the conformational changes in the β<span class="hlt">2</span>AR, which may reflect the different energetics of activation by diffusible ligands. PMID:11353823</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4518503','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4518503"><span>Fasting-<span class="hlt">induced</span> <span class="hlt">G</span>0/<span class="hlt">G</span>1 switch gene <span class="hlt">2</span> and FGF21 expression in the liver are under regulation of adipose tissue derived fatty acids</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Jaeger, Doris; Schoiswohl, Gabriele; Hofer, Peter; Schreiber, Renate; Schweiger, Martina; Eichmann, Thomas O.; Pollak, Nina M.; Poecher, Nadja; Grabner, Gernot F.; Zierler, Kathrin A.; Eder, Sandra; Kolb, Dagmar; Radner, Franz P.W.; Preiss-Landl, Karina; Lass, Achim; Zechner, Rudolf; Kershaw, Erin E.; Haemmerle, Guenter</p> <p>2015-01-01</p> <p>Background & Aims Adipose tissue (AT)-derived fatty acids (FAs) are utilized for hepatic triacylglycerol (TG) generation upon fasting. However, their potential impact as signaling molecules is not established. Herein we examined the role of exogenous AT-derived FAs in the regulation of hepatic gene expression by investigating mice with a defect in AT-derived FA supply to the liver. Methods Plasma FA levels, tissue TG hydrolytic activities and lipid content were determined in mice lacking the lipase co-activator comparative gene identification-58 (CGI-58) selectively in AT (CGI-58-ATko) applying standard protocols. Hepatic expression of lipases, FA oxidative genes, transcription factors, ER stress markers, hormones and cytokines were determined by qRT-PCR, Western blotting and ELISA. Results Impaired AT-derived FA supply upon fasting of CGI-58-ATko mice causes a marked defect in liver PPARα-signaling and nuclear CREBH translocation. This severely reduced the expression of respective target genes such as the ATGL inhibitor <span class="hlt">G</span>0/<span class="hlt">G</span>1 switch gene-<span class="hlt">2</span> (<span class="hlt">G</span>0S<span class="hlt">2</span>) and the endocrine metabolic regulator FGF21. These changes could be reversed by lipid administration and raising plasma FA levels. Impaired AT-lipolysis failed to <span class="hlt">induce</span> hepatic <span class="hlt">G</span>0S<span class="hlt">2</span> expression in fasted CGI-58-ATko mice leading to enhanced ATGL-mediated TG-breakdown strongly reducing hepatic TG deposition. On high fat diet, impaired AT-lipolysis counteracts hepatic TG accumulation and liver stress linked to improved systemic insulin sensitivity. Conclusions AT-derived FAs are a critical regulator of hepatic fasting gene expression required for the induction of <span class="hlt">G</span>0S<span class="hlt">2</span>-expression in the liver to control hepatic TG-breakdown. Interfering with AT-lipolysis or hepatic <span class="hlt">G</span>0S<span class="hlt">2</span> expression represents an effective strategy for the treatment of hepatic steatosis. PMID:25733154</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=326991','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=326991"><span>Isolation and characterization of rabbit anti-m3 <span class="hlt">2,2</span>,7<span class="hlt">G</span> antibodies.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Luhrmann, R; Appel, B; Bringmann, P; Rinke, J; Reuter, R; Rothe, S; Bald, R</p> <p>1982-01-01</p> <p>Antibodies specific for intact <span class="hlt">2,2</span>,7-trimethylguanosine (m3 <span class="hlt">2,2</span>,7<span class="hlt">G</span>) were <span class="hlt">induced</span> by immunization of rabbits with a nucleoside-human serum albumen (HSA) conjugate. Competition radioimmunoassay showed that the antibody distinguishes well between intact m3 <span class="hlt">2,2</span>,7<span class="hlt">G</span> and its alkali-hydrolysed form (m3 <span class="hlt">2,2</span>,7<span class="hlt">G</span>*). Antibody specificity is largely dependent on the presence of all three methyl groups in m3 <span class="hlt">2,2</span>,7<span class="hlt">G</span>: none of the less extensively methylated nucleosides m7<span class="hlt">G</span>, m<span class="hlt">2</span><span class="hlt">G</span> and m<span class="hlt">2</span> <span class="hlt">2,2</span><span class="hlt">G</span> is able to compete efficiently with the homologous hapten. Little or no competition was observed with m1<span class="hlt">G</span>, m1A, m6A, m5U and each of the four unmodified ribonucleosides. Binding studies with nucleoplasmic RNAs from Ehrlich ascites cells suggest that the antibody reacts specifically with the m3 <span class="hlt">2,2</span>,7<span class="hlt">G</span>-containing cap structure of the small nuclear U-RNAs (U-snRNAs). Thus the antibody should be a valuable tool for studying the role of the 5'-terminal regions of the U-snRNAs of eucaryotic cells. Images PMID:7155893</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19780040992&hterms=incubation+period&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dincubation%2Bperiod','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19780040992&hterms=incubation+period&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dincubation%2Bperiod"><span>The effect of NaCl/<span class="hlt">g</span>/ on the Na<span class="hlt">2</span>SO4-<span class="hlt">induced</span> hot corrosion of NiAl</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Smeggil, J. G.; Bornstein, N. S.; Decrescente, M. A.</p> <p>1977-01-01</p> <p>Studies have been performed to examine the effect of NaCl vapor on the Na<span class="hlt">2</span>SO4-<span class="hlt">induced</span> hot corrosion of the alumina former NiAl. In the incubation period associated with such hot corrosion, NaCl(<span class="hlt">g</span>) has been shown to be effective in removing aluminum from below the protective alumina scale and redepositing it as Al<span class="hlt">2</span>O3 whiskers on the surface of the Na<span class="hlt">2</span>SO4-coated sample. Similar effects seen in simple oxidation are associated with isothermal rupturing of the protective alumina scale.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4794602','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4794602"><span>DNA Strand <span class="hlt">Breaks</span> in Mitotic Germ Cells of Caenorhabditis elegans Evaluated by Comet Assay</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Park, Sojin; Choi, Seoyun; Ahn, Byungchan</p> <p>2016-01-01</p> <p>DNA damage responses are important for the maintenance of genome stability and the survival of organisms. Such responses are activated in the presence of DNA damage and lead to cell cycle arrest, apoptosis, and DNA repair. In Caenorhabditis elegans, double-strand <span class="hlt">breaks</span> <span class="hlt">induced</span> by DNA damaging agents have been detected indirectly by antibodies against DSB recognizing proteins. In this study we used a comet assay to detect DNA strand <span class="hlt">breaks</span> and to measure the elimination of DNA strand <span class="hlt">breaks</span> in mitotic germline nuclei of C. elegans. We found that C. elegans brc-1 mutants were more sensitive to ionizing radiation and camptothecin than the N<span class="hlt">2</span> wild-type strain and repaired DNA strand <span class="hlt">breaks</span> less efficiently than N<span class="hlt">2</span>. This study is the first demonstration of direct measurement of DNA strand <span class="hlt">breaks</span> in mitotic germline nuclei of C. elegans. This newly developed assay can be applied to detect DNA strand <span class="hlt">breaks</span> in different C. elegans mutants that are sensitive to DNA damaging agents. PMID:26903030</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3816019','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3816019"><span>Depigmented Allergoids Reveal New Epitopes with Capacity to <span class="hlt">Induce</span> Ig<span class="hlt">G</span> Blocking Antibodies</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>López-Matas, M. Angeles; Gallego, Mayte; Iraola, Víctor; Robinson, Douglas; Carnés, Jerónimo</p> <p>2013-01-01</p> <p>Background. The synthesis of allergen-specific blocking IgGs that interact with IgE after allergen immunotherapy (SIT) has been related to clinical efficacy. The objectives were to investigate the epitope specificity of Ig<span class="hlt">G</span>-antibodies <span class="hlt">induced</span> by depigmented-polymerized (Dpg-Pol) allergoids and unmodified allergen extracts, and examine IgE-blocking activity of <span class="hlt">induced</span> Ig<span class="hlt">G</span>-antibodies. Methods. Rabbits were immunized with native and Dpg-Pol extracts of birch pollen, and serum samples were obtained. Recognition of linear Ig<span class="hlt">G</span>-epitopes of Bet v 1 and Bet v <span class="hlt">2</span> and the capacity of these Ig<span class="hlt">G</span>-antibodies to block binding of human-IgE was determined. Results. Serum from rabbits immunized with native extracts recognised 11 linear epitopes from Bet v 1, while that from Dpg-Pol-immunized animals recognised 8. For Bet v <span class="hlt">2</span>, 8 epitopes were recognized by Ig<span class="hlt">G</span> from native immunized animals, and 9 from Dpg-Pol immunized one. Dpg-Pol and native immunized serum did not always recognise the same epitopes, but specific-Ig<span class="hlt">G</span> from both could block human-IgE binding sites for native extract. Conclusions. Depigmented-polymerized birch extract stimulates the synthesis of specific Ig<span class="hlt">G</span>-antibodies which recognize common but also novel epitopes compared with native extracts. Ig<span class="hlt">G</span>-antibodies <span class="hlt">induced</span> by Dpg-Pol effectively inhibit human-IgE binding to allergens which may be part of the mechanism of action of SIT. PMID:24222901</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20050000409&hterms=leukemia&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dleukemia','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20050000409&hterms=leukemia&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dleukemia"><span>A computational approach to the relationship between radiation <span class="hlt">induced</span> double strand <span class="hlt">breaks</span> and translocations</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Holley, W. R.; Chatterjee, A.</p> <p>1994-01-01</p> <p>A theoretical framework is presented which provides a quantitative analysis of radiation <span class="hlt">induced</span> translocations between the ab1 oncogene on CH9q34 and a breakpoint cluster region, bcr, on CH 22q11. Such translocations are associated frequently with chronic myelogenous leukemia. The theory is based on the assumption that incorrect or unfaithful rejoining of initial double strand <span class="hlt">breaks</span> produced concurrently within the 200 kbp intron region upstream of the second abl exon, and the 16.5 kbp region between bcr exon <span class="hlt">2</span> and exon 6 interact with each other, resulting in a fusion gene. for an x-ray dose of 100 Gy, there is good agreement between the theoretical estimate and the one available experimental result. The theory has been extended to provide dose response curves for these types of translocations. These curves are quadratic at low doses and become linear at high doses.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23370448','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23370448"><span>Black soybean seed coat polyphenols prevent B(a)P-<span class="hlt">induced</span> DNA damage through modulating drug-metabolizing enzymes in Hep<span class="hlt">G</span><span class="hlt">2</span> cells and ICR mice.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Tianshun; Jiang, Songyan; He, Chao; Kimura, Yuki; Yamashita, Yoko; Ashida, Hitoshi</p> <p>2013-04-15</p> <p>Black soybean seed coat is a rich source of polyphenols that have been reported to have various physiological functions. The present study investigated the potential protective effects of polyphenolic extracts from black soybean seed coat on DNA damage in human hepatoma Hep<span class="hlt">G</span><span class="hlt">2</span> cells and ICR mice. The results from micronucleus (MN) assay revealed that black soybean seed coat extract (BE) at concentrations up to 25μ<span class="hlt">g</span>/mL was non-genotoxic. It is noteworthy that BE (at 4.85μ<span class="hlt">g</span>/mL) and its main components, procyanidins (PCs) and cyanidin 3-glucoside (C3<span class="hlt">G</span>), at 10μM significantly reduced the genotoxic effect <span class="hlt">induced</span> by benzo[a]pyrene [B(a)P]. To obtain insights into the underlying mechanism, we investigated BE and its main components on drug-metabolizing enzyme expression. The results of this study demonstrate that BE and its main components, PCs and C3<span class="hlt">G</span>, down-regulated B(a)P-<span class="hlt">induced</span> cytochrome P4501A1 (CYP1A1) expression by inhibiting the transformation of aryl hydrocarbon receptor. Moreover, they increased expression of detoxifying defense enzymes, glutathione S-transferases (GSTs) via increasing the binding of nuclear factor-erythroid-<span class="hlt">2</span>-related factor <span class="hlt">2</span> to antioxidant response elements. Collectively, we found that PCs and C3<span class="hlt">G</span>, which are the main active compounds of BE, down-regulated CYP1A1 and up-regulated GST expression to protect B(a)P-<span class="hlt">induced</span> DNA damage in Hep<span class="hlt">G</span><span class="hlt">2</span> cells and ICR mice effectively. Copyright © 2013 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27459171','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27459171"><span>Serum PAI-1 and PAI-1 4<span class="hlt">G</span>/5<span class="hlt">G</span> Polymorphism in Hepatitis C Virus-<span class="hlt">Induced</span> Cirrhosis and Hepatitis C Virus-<span class="hlt">Induced</span> Hepatocellular Carcinoma Patients.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>El Edel, Rawhia H; Essa, Enas Said; Essa, Abdallah S; Hegazy, Sara A; El Rowedy, Dalia I</p> <p>2016-11-01</p> <p>Association between variable agent-<span class="hlt">induced</span> hepatocellular carcinoma (HCC) and both PAI-1 4<span class="hlt">G</span>/5<span class="hlt">G</span> polymorphism and plasminogen activator inhibitor (PAI-1) levels compared to healthy controls have been reported in earlier studies. We aimed to assess serum PAI-1 and PAI-1 4<span class="hlt">G</span>/5<span class="hlt">G</span> polymorphism in hepatitis C virus (HCV)-<span class="hlt">induced</span> HCC, HCV-<span class="hlt">induced</span> liver cirrhosis, and viral infection-free apparently healthy control subjects. Forty nine HCC, 52 cirrhosis, and 105 controls were genotyped for PAI-1 4<span class="hlt">G</span>/5<span class="hlt">G</span> using an allele-specific polymerase chain reaction analysis. In addition, for 31 HCC, 24 cirrhosis, and 28 controls, serum PAI-1 level was measured by enzyme-linked immunosorbent assay (ELISA). There was no significant difference in PAI-1 4<span class="hlt">G</span>/5<span class="hlt">G</span> genotype distribution between cirrhosis and controls (p = 0.33, p = 0.15, and p = 0.38 for the codominant, dominant, and recessive models, respectively) or between HCC and cirrhosis (p = 0.5, p = 0.24, and p = 0.69 for the codominant, dominant, and recessive models, respectively). Serum PAI-1 was significantly higher in cirrhosis than controls and significantly lower in HCC than cirrhosis (p < 0.001 for both). Serum PAI-1 did not differ significantly among the three PAI-1 4<span class="hlt">G</span>/5<span class="hlt">G</span> genotypes in controls, cirrhosis, and HCC (p = 0.29, p = 0.28, and p = 0.73 respectively). We documented higher serum PAI-1 in HCV-<span class="hlt">induced</span> HCC than viral infection-free controls, but interestingly, lower than HCV-<span class="hlt">induced</span> liver cirrhosis patients. This was not genotype related. Further studies will be needed to clearly elucidate the underlying mechanism.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28735729','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28735729"><span>Protective effects of flavonoids isolated from Korean milk thistle Cirsium japonicum var. maackii (Maxim.) Matsum on tert-butyl hydroperoxide-<span class="hlt">induced</span> hepatotoxicity in Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jung, Hyun Ah; Abdul, Qudeer Ahmed; Byun, Jeong Su; Joung, Eun-Ji; Gwon, Wi-Gyeong; Lee, Min-Sup; Kim, Hyeung-Rak; Choi, Jae Sue</p> <p>2017-09-14</p> <p>Milk thistle leaves and flowers have been traditionally used as herbal remedy to alleviate liver diseases for decades. Korean milk thistle, Cirsium japonicum var. maackii (Maxim.) Matsum has been employed in traditional folk medicine as diuretic, antiphlogistic, hemostatic, and detoxifying agents. The aim of current investigation was to evaluate hepatoprotective properties of the MeOH extract of the roots, stems, leaves and flowers of Korean milk thistle as well as four isolated flavonoids, luteolin, luteolin 5-O-glucoside, apigenin and apigenin 7-O-glucuronide during t-BHP-<span class="hlt">induced</span> oxidative stress in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Hepatoprotective potential of the MeOH extracts and flavonoids derived from Korean milk thistle against t-BHP-<span class="hlt">induced</span> oxidative stress in Hep<span class="hlt">G</span><span class="hlt">2</span> cells were evaluated following MTT method. Incubating Hep<span class="hlt">G</span><span class="hlt">2</span> cells with t-BHP markedly decreased the cell viability and increased the intracellular ROS generation accompanied by depleted GSH levels. Protein expression of heme oxygenase (HO-1) and nuclear factor-E<span class="hlt">2</span>-related factor <span class="hlt">2</span> (Nrf-<span class="hlt">2</span>) was determined by Western blot. Our findings revealed that pretreating Hep<span class="hlt">G</span><span class="hlt">2</span> cells with MeOH extracts and bioactive flavonoids significantly attenuated the t-BHP-<span class="hlt">induced</span> oxidative damage, followed by increased cell viability in a dose-dependent manner. The results illustrate that excess ROS generation was reduced and GSH levels increased dose-dependently when Hep<span class="hlt">G</span><span class="hlt">2</span> cells were pretreated with four flavonoids. Moreover, Western blotting analysis demonstrated that protein expressions of Nrf-<span class="hlt">2</span> and HO-1 were also up-regulated by flavonoids treatment. These results clearly demonstrate that the MeOH extracts and flavonoids from Korean milk thistle protected Hep<span class="hlt">G</span><span class="hlt">2</span> cells against oxidative damage triggered by t-BHP principally by modulating ROS generation and restoring depleted GSH levels in addition to the increased Nrf-<span class="hlt">2</span>/HO-1 signaling cascade. These flavonoids are potential natural antioxidative biomarkers against oxidative stress-<span class="hlt">induced</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20364911','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20364911"><span>Regulating strain states by using the recovery potential of lunch <span class="hlt">breaks</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Krajewski, Jarek; Wieland, Rainer; Sauerland, Martin</p> <p>2010-04-01</p> <p>The aim of the worksite study is to elucidate the strain reducing impact of different forms of spending lunch <span class="hlt">breaks</span>. With the help of the so-called silent room cabin concept, it was possible to <span class="hlt">induce</span> a lunch-<span class="hlt">break</span> relaxation opportunity that provided visual and territorial privacy. To evaluate the proposed effects, 14 call center agents were assigned to either 20 min progressive muscle relaxation (PMR) or small-talk (ST) <span class="hlt">break</span> groups. We analyzed the data in a controlled trial for a period of 6 months (every <span class="hlt">2</span> months four measurements a day at 12:00, 13:00, 16:00, 20:00) using independent observer and self-report ratings of emotional, mental, motivational, and physical strain. Results indicated that only the PMR <span class="hlt">break</span> reduced postlunchtime and afternoon strain. Although further intervention research is required, our results suggest that PMR lunch <span class="hlt">break</span> may sustainable reduce strain states in real worksite settings. Copyright 2010 APA, all rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24722377','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24722377"><span>Chalcones suppress fatty acid-<span class="hlt">induced</span> lipid accumulation through a LKB1/AMPK signaling pathway in Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Tianshun; Yamamoto, Norio; Ashida, Hitoshi</p> <p>2014-06-01</p> <p>Excessive lipid accumulation in the liver has been proposed to cause hyperlipidemia, diabetes and fatty liver disease. 4-Hydroxyderricin (4HD), xanthoangelol (XAG), cardamonin (CAR) and flavokawain B (FKB) are chalcones that have exhibited various biological effects against obesity, inflammation, and diabetes; however, little is known about the inhibitory effects of these chalcones on fatty liver disease. In the present study, we investigated the ability of 4HD, XAG, CAR, and FKB to reduce lipid accumulation in hepatocytes. When Hep<span class="hlt">G</span><span class="hlt">2</span> cells were treated with a mixture of fatty acids (FAs; palmitic acid : oleic acid = 1 : <span class="hlt">2</span> ratio), significant lipid accumulation was observed. Under the same experimental conditions, addition of chalcones at 5 μM significantly suppressed the FA-<span class="hlt">induced</span> lipid accumulation. We found that the expression of sterol regulatory element-binding protein-1 (SREBP-1), a key molecule involved in lipogenesis, was decreased in these chalcone-treated cells. We also found that these chalcones increased the expression of peroxisome proliferator-activated receptor α (PPARα), which is involved in FA oxidation. Moreover, these chalcones increased phosphorylation of AMP-activated protein kinase (AMPK) and liver kinase B1 (LKB1), upstream regulators of SREBP-1 and PPARα. We confirmed that an AMPK inhibitor, compound C, reversed chalcone-<span class="hlt">induced</span> changes in SREBP-1 and PPARα expression in the Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Collectively, we found that 4HD, XAG, CAR, and XAG attenuated lipid accumulation through activation of the LKB1/AMPK signaling pathway in Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12492371','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12492371"><span>Influence of environmental pH on <span class="hlt">G</span><span class="hlt">2</span>-phase arrest caused by ionizing radiation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Park, Heon Joo; Lee, Sang Hwa; Chung, HyunSook; Rhee, Yun Hee; Lim, Byung Uk; Ha, Sung Whan; Griffin, Robert J; Lee, Hyung Sik; Song, Chang Won; Choi, Eun Kyung</p> <p>2003-01-01</p> <p>We investigated the effects of an acidic environment on the <span class="hlt">G</span><span class="hlt">2</span>/M-phase arrest, apoptosis, clonogenic death, and changes in cyclin B1-CDC<span class="hlt">2</span> kinase activity caused by a 4-Gy irradiation in RKO.C human colorectal cancer cells in vitro. The time to reach peak <span class="hlt">G</span><span class="hlt">2</span>/M-phase arrest after irradiation was delayed in pH 6.6 medium compared to that in pH 7.5 medium. Furthermore, the radiation-<span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span>/M-phase arrest decayed more slowly in pH 6.6 medium than in pH 7.5 medium. Finally, there was less radiation-<span class="hlt">induced</span> apoptosis and clonogenic cell death in pH 6.6 medium than in pH 7.5 medium. It appeared that the prolongation of <span class="hlt">G</span><span class="hlt">2</span>-phase arrest after irradiation in the acidic environment allowed for greater repair of radiation-<span class="hlt">induced</span> DNA damage, thereby decreasing the radiation-<span class="hlt">induced</span> cell death. The prolongation of <span class="hlt">G</span><span class="hlt">2</span>-phase arrest after irradiation in the acidic pH environment appeared to be related at least in part to a prolongation of the phosphorylation of CDC<span class="hlt">2</span>, which inhibited cyclin B1-CDC<span class="hlt">2</span> kinase activity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26792947','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26792947"><span>DNA strand <span class="hlt">breaks</span> and crosslinks <span class="hlt">induced</span> by transient anions in the range <span class="hlt">2</span>-20 eV.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Luo, Xinglan; Zheng, Yi; Sanche, Léon</p> <p>2014-04-15</p> <p>The energy dependence of the yields of single and double strand <span class="hlt">breaks</span> (SSB and DSB) and crosslinks <span class="hlt">induced</span> by electron impact on plasmid DNA films is measured in the <span class="hlt">2</span>-20 eV range. The yield functions exhibit two strong maxima, which are interpreted to result from the formation of core-excited resonances (i.e., transient anions) of the bases, and their decay into the autoionization channel, resulting in π → π * electronic transitions of the bases followed by electron transfer to the C-O σ * bond in the phosphate group. Occupancy of the σ * orbital ruptures the C-O bond of the backbone via dissociative electron attachment, producing a SSB. From a comparison of our results with those of other works, including theoretical calculations and electron-energy-loss spectra of the bases, the 4.6 eV peak in the SSB yield function is attributed to the resonance decay into the lowest electronically excited states of the bases; in particular, those resulting from the transitions 1 3 A'( π <span class="hlt">2</span> → π 3 *) and 1 3 A″(n <span class="hlt">2</span> → π 3 *) of thymine and 1 3 A'( π → π *) of cytosine. The strongest peak at 9.6 eV in the SSB yield function is also associated with electron captured by excited states of the bases, resulting mostly from a multitude of higher-energy π → π * transitions. The DSB yield function exhibits strong maxima at 6.1 and 9.6 eV. The peak at 9.6 eV is probably related to the same resonance manifold as that leading to SSB, but the other at 6.1 eV may be more restricted to decay into the electronic state 1 3 A' ( π → π *) of cytosine via autoionization. The yield function of crosslinks is dominated by a broad peak extending over the 3.6-11.6 eV range with a sharper one at 17.6 eV. The different line shape of the latter function, compared to that of SSB and DSB, appears to be due to the formation of reactive radical sites in the initial supercoiled configuration of the plasmid, which react with the circular form (i.e., DNA with a SSB) to produce a</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5732777','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5732777"><span>PARP inhibition causes premature loss of cohesion in cancer cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Kukolj, Eva; Kaufmann, Tanja; Dick, Amalie E.; Zeillinger, Robert; Gerlich, Daniel W.; Slade, Dea</p> <p>2017-01-01</p> <p>Poly(ADP-ribose) polymerases (PARPs) regulate various aspects of cellular function including mitotic progression. Although PARP inhibitors have been undergoing various clinical trials and the PARP1/<span class="hlt">2</span> inhibitor olaparib was approved as monotherapy for BRCA-mutated ovarian cancer, their mode of action in killing tumour cells is not fully understood. We investigated the effect of PARP inhibition on mitosis in cancerous (cervical, ovary, breast and osteosarcoma) and non-cancerous cells by live-cell imaging. The clinically relevant inhibitor olaparib <span class="hlt">induced</span> strong perturbations in mitosis, including problems with chromosome alignment at the metaphase plate, anaphase delay, and premature loss of cohesion (cohesion fatigue) after a prolonged metaphase arrest, resulting in sister <span class="hlt">chromatid</span> scattering. PARP1 and PARP<span class="hlt">2</span> depletion suppressed the phenotype while PARP<span class="hlt">2</span> overexpression enhanced it, suggesting that olaparib-bound PARP1 and PARP<span class="hlt">2</span> rather than the lack of catalytic activity causes this phenotype. Olaparib-<span class="hlt">induced</span> mitotic <span class="hlt">chromatid</span> scattering was observed in various cancer cell lines with increased protein levels of PARP1 and PARP<span class="hlt">2</span>, but not in non-cancer or cancer cell lines that expressed lower levels of PARP1 or PARP<span class="hlt">2</span>. Interestingly, the sister <span class="hlt">chromatid</span> scattering phenotype occurred only when olaparib was added during the S-phase preceding mitosis, suggesting that PARP1 and PARP<span class="hlt">2</span> entrapment at replication forks impairs sister <span class="hlt">chromatid</span> cohesion. Clinically relevant DNA-damaging agents that impair replication progression such as topoisomerase inhibitors and cisplatin were also found to <span class="hlt">induce</span> sister <span class="hlt">chromatid</span> scattering and metaphase plate alignment problems, suggesting that these mitotic phenotypes are a common outcome of replication perturbation. PMID:29262611</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/2009589','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/2009589"><span>Sister <span class="hlt">chromatid</span> exchange in children of Seventh-Day Adventists and matched controls.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hermansen, R; Waksvik, H; Fønnebø, V</p> <p>1991-03-01</p> <p>The low risk of cancer in Seventh-Day Adventists (SDAs) has been suggested to be due to genetic selection. To investigate this claim we examined the sister <span class="hlt">chromatid</span> exchange (SCE) frequency in peripheral blood lymphocytes in 16 SDA children in Tromsø, all aged 0.5-8 years and 16 controls matched for sex and age. In 12 of 16 pairs, the SDA children had a lower SCE frequency than the controls. The mean difference was 4.06 (95% confidence interval -17.02-8.89, P = 0.51). There was no sex difference, and no correlation between age and SCE frequency. The genetic starting point with regard to SCE frequency seems to be the same for SDA children and controls.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..MAR.V1005O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..MAR.V1005O"><span>Optical probes of symmetry <span class="hlt">breaking</span> in magnetic and superconducting BaFe<span class="hlt">2</span>(As1-xPx)<span class="hlt">2</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Orenstein, Joseph</p> <p></p> <p>The discovery of iron pnictide superconductors has opened promising new directions in the effort to fully understand the phenomenon of high-Tc, with a focus on the connections between superconductivity, magnetism, and electronic nematicity. The BaFe<span class="hlt">2</span>(As1-xPx)<span class="hlt">2</span> (P:Ba122) system in particular has received attention because isovalent substitution of As for P generates less disorder than doping on the Fe site. The phase diagram of P:Ba122 is characterized by a line of simultaneous antiferromagnetic (AF) and tetragonal-to-orthorhombic transitions, Ts (x) , that penetrates the superconducting dome at x =0.28, just below optimal doping (xopt = 0.30). In this work, we use spatially-resolved optical polarimetry and photomodulated reflectance to detect linear birefringence and therefore <span class="hlt">breaking</span> of 4-fold rotational (C4) symmetry. In underdoped (x<0.28) samples, birefringence appears at T>Tsand grows continuously with decreasing T . The birefringence is unidirectional in a large (300 μm x300 μm) field of view, suggesting that C4 <span class="hlt">breaking</span> in this range of T is caused by residual strain that couples to a diverging nematic susceptibility. Birefringence maps just below Ts (x) show the appearance of domains, indicating the onset of spontaneous symmetry <span class="hlt">breaking</span> to an AF ground state. Surprisingly, in samples with x>0.28, in which the low T phase is superconducting/ tetragonal rather than AF/orthorhombic, C4 <span class="hlt">breaking</span> is observed as well, with an abrupt onset and domain formation at 55 K. We tentatively associate these features with a transition to an AF phase <span class="hlt">induced</span> by residual strain, as previously proposed [H.-H. Kuo et al. Phys. Rev. B86, 134507 (2012)] to account for structure in resistivity vs. T. Time-resolved photomodulation allow us to follow the amplitude of the AF order with time following pulsed photoexcitation. Below Tc the AF order at first weakens , but then strengthens in response to the photoinduced weakening of superconductivity. This complex time evolution is</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5356696','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5356696"><span>Gremlin inhibits UV-<span class="hlt">induced</span> skin cell damages via activating VEGFR<span class="hlt">2</span>-Nrf<span class="hlt">2</span> signaling</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Xu, Qiu-yun; Zhang, Jing; Lin, Meng-ting; Tu, Ying; He, Li; Bi, Zhi-gang; Cheng, Bo</p> <p>2016-01-01</p> <p>Ultra Violet (UV) radiation <span class="hlt">induces</span> reactive oxygen species (ROS) production, DNA oxidation and single strand <span class="hlt">breaks</span> (SSBs), which will eventually lead to skin cell damages or even skin cancer. Here, we tested the potential activity of gremlin, a novel vascular endothelial growth factor (VEGF) receptor <span class="hlt">2</span> (VEGFR<span class="hlt">2</span>) agonist, against UV-<span class="hlt">induced</span> skin cell damages. We show that gremlin activated VEGFR<span class="hlt">2</span> and significantly inhibited UV-<span class="hlt">induced</span> death and apoptosis of skin keratinocytes and fibroblasts. Pharmacological inhibition or shRNA-mediated knockdown of VEGFR<span class="hlt">2</span> almost abolished gremlin-mediated cytoprotection against UV in the skin cells. Further studies showed that gremlin activated VEGFR<span class="hlt">2</span> downstream NF-E<span class="hlt">2</span>-related factor <span class="hlt">2</span> (Nrf<span class="hlt">2</span>) signaling, which appeared required for subsequent skin cell protection. Nrf<span class="hlt">2</span> shRNA knockdown or S40T dominant negative mutation largely inhibited gremlin-mediated skin cell protection against UV. At last, we show that gremlin dramatically inhibited UV-<span class="hlt">induced</span> ROS production and DNA SSB formation in skin keratinocytes and fibroblasts. We conclude that gremlin protects skin cells from UV damages via activating VEGFR<span class="hlt">2</span>-Nrf<span class="hlt">2</span> signaling. Gremlin could be further tested as a novel anti-UV skin protectant. PMID:27713170</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27713170','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27713170"><span>Gremlin inhibits UV-<span class="hlt">induced</span> skin cell damages via activating VEGFR<span class="hlt">2</span>-Nrf<span class="hlt">2</span> signaling.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ji, Chao; Huang, Jin-Wen; Xu, Qiu-Yun; Zhang, Jing; Lin, Meng-Ting; Tu, Ying; He, Li; Bi, Zhi-Gang; Cheng, Bo</p> <p>2016-12-20</p> <p>Ultra Violet (UV) radiation <span class="hlt">induces</span> reactive oxygen species (ROS) production, DNA oxidation and single strand <span class="hlt">breaks</span> (SSBs), which will eventually lead to skin cell damages or even skin cancer. Here, we tested the potential activity of gremlin, a novel vascular endothelial growth factor (VEGF) receptor <span class="hlt">2</span> (VEGFR<span class="hlt">2</span>) agonist, against UV-<span class="hlt">induced</span> skin cell damages. We show that gremlin activated VEGFR<span class="hlt">2</span> and significantly inhibited UV-<span class="hlt">induced</span> death and apoptosis of skin keratinocytes and fibroblasts. Pharmacological inhibition or shRNA-mediated knockdown of VEGFR<span class="hlt">2</span> almost abolished gremlin-mediated cytoprotection against UV in the skin cells. Further studies showed that gremlin activated VEGFR<span class="hlt">2</span> downstream NF-E<span class="hlt">2</span>-related factor <span class="hlt">2</span> (Nrf<span class="hlt">2</span>) signaling, which appeared required for subsequent skin cell protection. Nrf<span class="hlt">2</span> shRNA knockdown or S40T dominant negative mutation largely inhibited gremlin-mediated skin cell protection against UV. At last, we show that gremlin dramatically inhibited UV-<span class="hlt">induced</span> ROS production and DNA SSB formation in skin keratinocytes and fibroblasts. We conclude that gremlin protects skin cells from UV damages via activating VEGFR<span class="hlt">2</span>-Nrf<span class="hlt">2</span> signaling. Gremlin could be further tested as a novel anti-UV skin protectant.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li class="active"><span>21</span></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_21 --> <div id="page_22" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li class="active"><span>22</span></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="421"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/5998375-chromosomal-aberrations-delays-cell-progression-induced-rays-tradescantia-clone-meristems','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/5998375-chromosomal-aberrations-delays-cell-progression-induced-rays-tradescantia-clone-meristems"><span>Chromosomal aberrations and delays in cell progression <span class="hlt">induced</span> by x-rays in Tradescantia clone 02 meristems</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Geard, C.R.</p> <p>1983-01-01</p> <p>In root meristems of Tradescantia clone 02 (developed by Sparrow and his colleagues for mutation studies), X-rays interfere with the progression of cells through the cell cycle and <span class="hlt">induce</span> chromosomal aberrations in a dose-dependent manner consistent with linear-quadratic kinetics. Sequential mitotic cell accumulations after irradiation indicate that sensitivity to aberration induction is probably greatest in cells from late S to early <span class="hlt">G</span><span class="hlt">2</span>, with <span class="hlt">chromatid</span> interchanges the most frequent aberration type and all aberrations consistent with initiation from the interaction between two lesions. The ratio of the coefficients in the linear (..cap alpha..) and the quadratic (..beta..) terms (..cap alpha../..beta..) ismore » equal to the dose average of specific energy produced by individual particles in the site where interaction takes place. The ratio ..cap alpha../..beta.. for chromosomal aberrations is similar to that previously found for X-ray-<span class="hlt">induced</span> mutation in Tradescantia stamen hairs, supporting the proposal that radiation-<span class="hlt">induced</span> mutational events are due to chromosomal aberrations with interaction distances of about 1..mu..m. Abrahamson and co-workers have noted that both ..cap alpha../..beta.. ratios appear to be related to nuclear target size and are similar for chromosomal and mutational endpoints in the same organism. These findings support this concept; however, it is apparent that any situation which diminishes yield at high doses (e.<span class="hlt">g</span>., mitotic delay) will probably affect the ..beta.. component. 23 references, 5 figures, <span class="hlt">2</span> tables.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23926120','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23926120"><span>5-demethyltangeretin inhibits human nonsmall cell lung cancer cell growth by <span class="hlt">inducing</span> <span class="hlt">G</span><span class="hlt">2</span>/M cell cycle arrest and apoptosis.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Charoensinphon, Noppawat; Qiu, Peiju; Dong, Ping; Zheng, Jinkai; Ngauv, Pearline; Cao, Yong; Li, Shiming; Ho, Chi-Tang; Xiao, Hang</p> <p>2013-12-01</p> <p>Tangeretin (TAN) and 5-demethyltangeretin (5DT) are two closely related polymethoxyflavones found in citrus fruits. We investigated growth inhibitory effects on three human nonsmall cell lung cancer (NSCLC) cells. Cell viability assay demonstrated that 5DT inhibited NSCLC cell growth in a time- and dose-dependent manner, and IC50 s of 5DT were 79-fold, 57-fold, and 56-fold lower than those of TAN in A549, H460, and H1299 cells, respectively. Flow cytometry analysis showed that 5DT <span class="hlt">induced</span> extensive <span class="hlt">G</span><span class="hlt">2</span>/M cell cycle arrest and apoptosis in NSCLC cells, while TAN at tenfold higher concentrations did not. The apoptosis <span class="hlt">induced</span> by 5DT was further confirmed by activation of caspase-3 and cleavage of PARP. Moreover, 5DT dose-dependently upregulated p53 and p21(Cip1/Waf1), and downregulated Cdc-<span class="hlt">2</span> (Cdk-1) and cyclin B1. HPLC analysis revealed that the intracellular levels of 5DT in NSCLC cells were <span class="hlt">2</span>.7-4.9 fold higher than those of TAN after the cells were treated with 5DT or TAN at the same concentration. Our results demonstrated that 5DT inhibited NSCLC cell growth by <span class="hlt">inducing</span> <span class="hlt">G</span><span class="hlt">2</span>/M cell cycle arrest and apoptosis. These effects were much stronger than those produced by TAN, which is partially due to the higher intracellular uptake of 5DT than TAN. © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27607345','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27607345"><span>Protective Role of GPER Agonist <span class="hlt">G</span>-1 on Cardiotoxicity <span class="hlt">Induced</span> by Doxorubicin.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>De Francesco, Ernestina M; Rocca, Carmine; Scavello, Francesco; Amelio, Daniela; Pasqua, Teresa; Rigiracciolo, Damiano C; Scarpelli, Andrea; Avino, Silvia; Cirillo, Francesca; Amodio, Nicola; Cerra, Maria C; Maggiolini, Marcello; Angelone, Tommaso</p> <p>2017-07-01</p> <p>The use of Doxorubicin (Dox), a frontline drug for many cancers, is often complicated by dose-limiting cardiotoxicity in approximately 20% of patients. The <span class="hlt">G</span>-protein estrogen receptor GPER/GPR30 mediates estrogen action as the cardioprotection under certain stressful conditions. For instance, GPER activation by the selective agonist <span class="hlt">G</span>-1 reduced myocardial inflammation, improved immunosuppression, triggered pro-survival signaling cascades, improved myocardial mechanical performance, and reduced infarct size after ischemia/reperfusion (I/R) injury. Hence, we evaluated whether ligand-activated GPER may exert cardioprotection in male rats chronically treated with Dox. 1 week of <span class="hlt">G</span>-1 (50 μ<span class="hlt">g</span>/kg/day) intraperitoneal administration mitigated Dox (3 mg/kg/day) adverse effects, as revealed by reduced TNF-α, IL-1β, LDH, and ROS levels. Western blotting analysis of cardiac homogenates indicated that <span class="hlt">G</span>-1 prevents the increase in p-c-jun, BAX, CTGF, iNOS, and COX<span class="hlt">2</span> expression <span class="hlt">induced</span> by Dox. Moreover, the activation of GPER rescued the inhibitory action elicited by Dox on the expression of BCL<span class="hlt">2</span>, pERK, and pAKT. TUNEL assay indicated that GPER activation may also attenuate the cardiomyocyte apoptosis upon Dox exposure. Using ex vivo Langendorff perfused heart technique, we also found an increased systolic recovery and a reduction of both infarct size and LDH levels in rats treated with <span class="hlt">G</span>-1 in combination with Dox respect to animals treated with Dox alone. Accordingly, the beneficial effects <span class="hlt">induced</span> by <span class="hlt">G</span>-1 were abrogated in the presence of the GPER selective antagonist <span class="hlt">G</span>15. These data suggest that GPER activation mitigates Dox-<span class="hlt">induced</span> cardiotoxicity, thus proposing GPER as a novel pharmacological target to limit the detrimental cardiac effects of Dox treatment. J. Cell. Physiol. 232: 1640-1649, 2017. © 2016 Wiley Periodicals, Inc. © 2016 Wiley Periodicals, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27680669','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27680669"><span>Nucleolar Reorganization Upon Site-Specific Double-Strand <span class="hlt">Break</span> Induction.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Franek, Michal; Kovaříková, Alena; Bártová, Eva; Kozubek, Stanislav</p> <p>2016-11-01</p> <p>DNA damage response (DDR) in ribosomal genes and mechanisms of DNA repair in embryonic stem cells (ESCs) are less explored nuclear events. DDR in ESCs should be unique due to their high proliferation rate, expression of pluripotency factors, and specific chromatin signature. Given short population doubling time and fast progress through <span class="hlt">G</span>1 phase, ESCs require a sustained production of rRNA, which leads to the formation of large and prominent nucleoli. Although transcription of rRNA in the nucleolus is relatively well understood, little is known about DDR in this nuclear compartment. Here, we directed formation of double-strand <span class="hlt">breaks</span> in rRNA genes with I- PpoI endonuclease, and we studied nucleolar morphology, DDR, and chromatin modifications. We observed a pronounced formation of I- PpoI-<span class="hlt">induced</span> nucleolar caps, positive on BRCA1, NBS1, MDC1, γH<span class="hlt">2</span>AX, and UBF1 proteins. We showed interaction of nucleolar protein TCOF1 with HDAC1 and TCOF1 with CARM1 after DNA injury. Moreover, H3R17me<span class="hlt">2</span>a modification mediated by CARM1 was found in I- PpoI-<span class="hlt">induced</span> nucleolar caps. Finally, we report that heterochromatin protein 1 is not involved in DNA repair of nucleolar caps.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5084524','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5084524"><span>Nucleolar Reorganization Upon Site-Specific Double-Strand <span class="hlt">Break</span> Induction</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Franek, Michal; Kovaříková, Alena; Bártová, Eva; Kozubek, Stanislav</p> <p>2016-01-01</p> <p>DNA damage response (DDR) in ribosomal genes and mechanisms of DNA repair in embryonic stem cells (ESCs) are less explored nuclear events. DDR in ESCs should be unique due to their high proliferation rate, expression of pluripotency factors, and specific chromatin signature. Given short population doubling time and fast progress through <span class="hlt">G</span>1 phase, ESCs require a sustained production of rRNA, which leads to the formation of large and prominent nucleoli. Although transcription of rRNA in the nucleolus is relatively well understood, little is known about DDR in this nuclear compartment. Here, we directed formation of double-strand <span class="hlt">breaks</span> in rRNA genes with I-PpoI endonuclease, and we studied nucleolar morphology, DDR, and chromatin modifications. We observed a pronounced formation of I-PpoI-<span class="hlt">induced</span> nucleolar caps, positive on BRCA1, NBS1, MDC1, γH<span class="hlt">2</span>AX, and UBF1 proteins. We showed interaction of nucleolar protein TCOF1 with HDAC1 and TCOF1 with CARM1 after DNA injury. Moreover, H3R17me<span class="hlt">2</span>a modification mediated by CARM1 was found in I-PpoI-<span class="hlt">induced</span> nucleolar caps. Finally, we report that heterochromatin protein 1 is not involved in DNA repair of nucleolar caps. PMID:27680669</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3375024','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3375024"><span>Restoration of C/EBPα in dedifferentiated liposarcoma <span class="hlt">induces</span> <span class="hlt">G</span><span class="hlt">2</span>/M cell cycle arrest and apoptosis</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Wu, Yuhsin V.; Okada, Tomoyo; DeCarolis, Penelope; Socci, Nicholas; O’Connor, Rachael; Geha, Rula C.; Somberg, C. Joy; Antonescu, Cristina; Singer, Samuel</p> <p>2012-01-01</p> <p>Well differentiated liposarcoma (WDLS) and dedifferentiated liposarcoma (DDLS) represent the most common biological group of liposarcoma, and there is a pressing need to develop targeted therapies for patients with advanced disease. To identify potential therapeutic targets, we sought to identify differences in the adipogenic pathways between DDLS, WDLS, and normal adipose tissue. In a microarray analysis of DDLS (n=84), WDLS (n=79), and normal fat (n=23), C/EBPα, a transcription factor involved in cell cycle regulation and differentiation, was underexpressed in DDLS compared to both WDLS and normal fat (15.<span class="hlt">2</span> fold and 27.8 fold, respectively). In normal adipose-derived stem cells, C/EBPα expression was strongly <span class="hlt">induced</span> when cells were cultured in differentiation media, but in three DDLS cell lines, this induction was nearly absent. We restored C/EBPα expression in one of the cell lines (DDLS8817) by transfection of an <span class="hlt">inducible</span> C/EBα expression vector. <span class="hlt">Inducing</span> C/EBPα expression reduced proliferation and caused cells to accumulate in <span class="hlt">G</span><span class="hlt">2</span>/M. Under differentiation conditions, the cell proliferation was reduced further, and 66% of the DDLS cells containing the <span class="hlt">inducible</span> C/EBPα expression vector underwent apoptosis as demonstrated by annexin V staining. These cells in differentiation conditions expressed early adipocyte-specific mRNAs such as LPL and FABP4, but they failed to accumulate intracellular lipid droplets, a characteristic of mature adipocytes. These results demonstrate that loss of C/EBPα is an important factor in suppressing apoptosis and maintaining the dedifferentiated state in DDLS. Restoring C/EBPα may be a useful therapeutic approach for dedifferentiated liposarcomas. PMID:22170698</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvD..97e4504C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvD..97e4504C"><span>Nonperturbative study of dynamical SUSY <span class="hlt">breaking</span> in N =(<span class="hlt">2</span> ,<span class="hlt">2</span> ) Yang-Mills theory</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Catterall, Simon; Jha, Raghav G.; Joseph, Anosh</p> <p>2018-03-01</p> <p>We examine the possibility of dynamical supersymmetry <span class="hlt">breaking</span> in two-dimensional N =(<span class="hlt">2</span> ,<span class="hlt">2</span> ) supersymmetric Yang-Mills theory. The theory is discretized on a Euclidean spacetime lattice using a supersymmetric lattice action. We compute the vacuum energy of the theory at finite temperature and take the zero-temperature limit. Supersymmetry will be spontaneously broken in this theory if the measured ground-state energy is nonzero. By performing simulations on a range of lattices up to 96 ×96 we are able to perform a careful extrapolation to the continuum limit for a wide range of temperatures. Subsequent extrapolations to the zero-temperature limit yield an upper bound on the ground-state energy density. We find the energy density to be statistically consistent with zero in agreement with the absence of dynamical supersymmetry <span class="hlt">breaking</span> in this theory.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17168312','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17168312"><span>[Construction and expression of HSV-<span class="hlt">2</span><span class="hlt">g</span>D-Hsp70 fusion protein gene].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Fan, Jian-Yong; Yang, Hui-Lan; Wang, Ying; Guan, Lei</p> <p>2006-11-01</p> <p>To construct and express Hsp70-HSV<span class="hlt">2</span><span class="hlt">g</span>D fusion protein. Genes of Hsp70 and HSV-<span class="hlt">2</span><span class="hlt">g</span>D were subcloned into vectors pGEX-4T-1 respectively. After confirmed by DNA sequence analysis, the recombinant plasmids pGEX-4T-HSP-<span class="hlt">g</span>D was transformed into E. coli DH5alpha and <span class="hlt">induced</span> to express with IPTG. The expressed protein was characterized by SDS-PAGE and Western blot after purified. BALB/c mice were immunized with fusion proteins respectively via intra-m uscular injection. The proliferation of spleen lymphocytes, the level of y-IFN in culture and anti-HSV-<span class="hlt">2</span><span class="hlt">g</span>D Ig<span class="hlt">G</span> antibody in serum was detected was detected. The expressed protein was analyzed by SDS-PAGE after <span class="hlt">induced</span> with IPTG, which showed a new band with an apparent molecular mass corresponding to the predicted size (118 kD). Western Blotting analysis demonstrates that the purified Hsp70-HSV<span class="hlt">2</span><span class="hlt">g</span>D fusion protein had specific binding activity. The stimulation indexes of spleen lymphocytes, the level of gamma-IFN in culture and anti-HSV-<span class="hlt">2</span><span class="hlt">g</span>D Ig<span class="hlt">G</span> antibody in serum of GST-Hsp70-<span class="hlt">g</span>D group was obviously higher than that of other groups (P < 0.05 respectively). The successful expression of the Hsp70-HSV<span class="hlt">2</span><span class="hlt">g</span>D fusion protein, which can <span class="hlt">induce</span> immune responses, laid a solid foundation for its further research.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFMEP33A1056C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMEP33A1056C"><span>Investigation of River Seismic Signal <span class="hlt">Induced</span> by Sediment Transport and Water Flow: Controlled Dam <span class="hlt">Breaking</span> Experiments</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chen, H. Y.; Chen, S. C.; Chao, W. A.</p> <p>2015-12-01</p> <p>Natural river's bedload often hard to measure, which leads numerous uncertainties for us to predict the landscape evolution. However, the measurement of bedload flux has its certain importance to estimate the river hazard. Thus, we use seismometer to receive the seismic signal <span class="hlt">induced</span> by bedload for partially fill the gap of field measurement capabilities. Our research conducted a controlled dam <span class="hlt">breaking</span> experiments at Landao River, Huisun Forest since it has advantage to well constraining the spatial and temporal variation of bedload transport. We set continuous bedload trap at downstream riverbed of dam to trap the transport bedload after dam <span class="hlt">breaking</span> so as to analyze its grain size distribution and transport behavior. In the meantime we cooperate with two portable velocity seismometers (Guralp CMG6TD) along the river to explore the relationship between bedload transport and seismic signal. Bedload trap was divided into three layers, bottom, middle, and top respectively. After the experiment, we analyzed the grain size and found out the median particle size from bottom to top is 88.664mm, 129.601mm, and 214.801mm individually. The median particle size of top layer is similar with the upstream riverbed before the experiment which median particle size is 230.683mm. This phenomena indicated that as the river flow become stronger after dam <span class="hlt">breaking</span>, the sediment size will thereupon become larger, which meant the sediment from upstream will be carried down by the water flow and turned into bedload. Furthermore, we may tell apart the seismic signal <span class="hlt">induced</span> by water flow and bedload by means of two different position seismometers. Eventually, we may estimate the probable error band of bedload quantity via accurately control of water depth, time-lapse photography, 3D LiDAR and other hydrology parameters.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015NPPP..265...81R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015NPPP..265...81R"><span>Spontaneous Symmetry <span class="hlt">Breaking</span> in Supernova Neutrinos</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Raffelt, Georg G.</p> <p>2015-08-01</p> <p>Some recent developments in supernova neutrino physics are introduced where spontaneous symmetry <span class="hlt">breaking</span> is a common theme. The physics of self-<span class="hlt">induced</span> flavor conversion has acquired a new complication in that a new class of instabilities <span class="hlt">breaks</span> axial symmetry of a neutrino stream, the multi-azimuth angle (MAA) instability. A completely different new phenomenon, discovered in the first realistic three-dimensional (3D) simulations, is the Lepton-Emission Self-sustained Asymmetry (LESA) during the accretion phase. Here, a neutrino-hydrodynamical instability <span class="hlt">breaks</span> global spherical symmetry in that the lepton-number flux (νe minus ν‾e) develops a stable dipole pattern such that the lepton flux is almost exclusively emitted in one hemisphere.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22273915-pressure-effects-magnetic-pair-breaking-mn-eu-substituted-bafe-sub-sub','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22273915-pressure-effects-magnetic-pair-breaking-mn-eu-substituted-bafe-sub-sub"><span>Pressure effects on magnetic pair-<span class="hlt">breaking</span> in Mn- and Eu-substituted BaFe{sub <span class="hlt">2</span>}As{sub <span class="hlt">2</span>}</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Rosa, P. F. S., E-mail: ferrari@ifi.unicamp.br; University of California, Irvine, California 92697-4574; Garitezi, T. M.</p> <p>2014-05-07</p> <p>We report a combined study of hydrostatic pressure (P ≤ 25 kbar) and chemical substitution on the magnetic pair-<span class="hlt">breaking</span> effect in Eu- and Mn-substituted BaFe{sub <span class="hlt">2</span>}As{sub <span class="hlt">2</span>} single crystals. At ambient pressure, both substitutions suppress the superconducting (SC) transition temperature (T{sub c}) of BaFe{sub <span class="hlt">2</span>–x}Co{sub x}As{sub <span class="hlt">2</span>} samples slightly under the optimally doped region, indicating the presence of a pair-<span class="hlt">breaking</span> effect. At low pressures, an increase of T{sub c} is observed for all studied compounds followed by an expected decrease at higher pressures. However, in the Eu dilute system, T{sub c} further increases at higher pressure along with a narrowingmore » of the SC transition, suggesting that a pair-<span class="hlt">breaking</span> mechanism reminiscent of the Eu Kondo single impurity regime is being suppressed by pressure. Furthermore, Electron Spin Resonance (ESR) measurements indicate the presence of Mn{sup <span class="hlt">2</span>+} and Eu{sup <span class="hlt">2</span>+} local moments and the microscopic parameters extracted from the ESR analysis reveal that the Abrikosov–Gor'kov expression for magnetic pair-<span class="hlt">breaking</span> in a conventional sign-preserving superconducting state cannot describe the observed reduction of T{sub c}.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1424726-time-reversal-symmetry-breaking-superconductivity-epitaxial-bismuth-nickel-bilayers','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1424726-time-reversal-symmetry-breaking-superconductivity-epitaxial-bismuth-nickel-bilayers"><span>Time-reversal symmetry-<span class="hlt">breaking</span> superconductivity in epitaxial bismuth/nickel bilayers</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Gong, Xinxin; Kargarian, Mehdi; Stern, Alex; ...</p> <p>2017-03-31</p> <p>Superconductivity that spontaneously <span class="hlt">breaks</span> time-reversal symmetry (TRS) has been found, so far, only in a handful of three-dimensional (3D) crystals with bulk inversion symmetry. We report an observation of spontaneous TRS <span class="hlt">breaking</span> in a <span class="hlt">2</span>D superconducting system without inversion symmetry: the epitaxial bilayer films of bismuth and nickel. The evidence comes from the onset of the polar Kerr effect at the superconducting transition in the absence of an external magnetic field, detected by the ultrasensitive loop-less fiber-optic Sagnac interferometer. Because of strong spin-orbit interaction and lack of inversion symmetry in a Bi/Ni bilayer, superconducting pairing cannot be classified as singletmore » or triplet.We propose a theoretical model where magnetic fluctuations in Ni <span class="hlt">induce</span> the superconducting pairing of the d xy ± id x<span class="hlt">2</span>-y<span class="hlt">2</span> orbital symmetry between the electrons in Bi. In this model, the order parameter spontaneously <span class="hlt">breaks</span> the TRS and has a nonzero phase winding number around the Fermi surface, thus making it a rare example of a <span class="hlt">2</span>D topological superconductor.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/7508088','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/7508088"><span>Chromosome damage in lymphocytes of stainless steel welders related to past and current exposure to manual metal arc welding fumes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jelmert, O; Hansteen, I L; Langård, S</p> <p>1994-02-01</p> <p>Cytogenetic damage was studied in lymphocytes from 42 welders using the manual metal arc (MMA) method on stainless steel (SS). A detailed characterization of previous exposure by job interviews, and for current exposure with personal air sampling and biological monitoring of chromium (Cr) and nickel (Ni) in blood and urine, was done for 32 of these welders. A subgroup of 20 welders was studied before and after 1-4 months of MMA/SS welding. A matched reference group I, and a larger reference group II were established for comparison. A significant increase in <span class="hlt">chromatid</span> <span class="hlt">breaks</span> (1.4 vs. 0.9 and 0.8 for group I and II) and for cells with aberrations (<span class="hlt">2.2</span> vs. 1.6 in group II) was found in the welders. An even larger difference was found when comparing non-smoking welders with their non-smoking referents. No synergistic effect between smoking and MMA/SS welding fumes was observed for any type of aberrations. Current welding fume exposure during the week before sampling was not associated with increases in any type of cytogenetic damage. The results indicated that the increase in <span class="hlt">chromatid</span> <span class="hlt">breaks</span> was associated with cumulated welding fume exposure for more than a year, and with not using respirators. Exposure to MMA/SS welding fumes for up to 4 months gave a slight, but significant increase in <span class="hlt">chromatid</span> <span class="hlt">breaks</span> when using the welders as their own referents. However, when using matched referents in the study after exposure, no difference was found between these welders and their matched referents. No differences between the groups were observed in the DNA synthesis and repair-inhibited cultures or for SCE.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26559846','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26559846"><span>A Herpes Simplex Virus <span class="hlt">2</span> (HSV-<span class="hlt">2</span>) <span class="hlt">g</span>D Mutant Impaired for Neural Tropism Is Superior to an HSV-<span class="hlt">2</span> <span class="hlt">g</span>D Subunit Vaccine To Protect Animals from Challenge with HSV-<span class="hlt">2</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Kening; Goodman, Kyle N; Li, Daniel Y; Raffeld, Mark; Chavez, Mayra; Cohen, Jeffrey I</p> <p>2016-01-01</p> <p>A recent phase 3 trial with soluble herpes simplex virus <span class="hlt">2</span> (HSV-<span class="hlt">2</span>) glycoprotein D (<span class="hlt">g</span>D<span class="hlt">2</span>t) in adjuvant failed to show protection against genital herpes. We postulated that live attenuated HSV-<span class="hlt">2</span> would provide more HSV antigens for induction of virus-specific antibodies and cellular immunity than would <span class="hlt">g</span>D<span class="hlt">2</span>t. We previously reported an HSV-<span class="hlt">2</span> mutant, HSV<span class="hlt">2</span>-<span class="hlt">g</span>D27, in which the nectin-1 binding domain of <span class="hlt">g</span>D<span class="hlt">2</span> is altered so that the virus is impaired for infecting neural cells, but not epithelial cells, in vitro and is impaired for infecting dorsal root ganglia in mice (K. Wang, J. D. Kappel, C. Canders, W. F. Davila, D. Sayre, M. Chavez, L. Pesnicak, and J. I. Cohen, J Virol 86:12891-12902, 2012, doi:10.1128/JVI.01055-12). Here we report that the mutations in HSV<span class="hlt">2</span>-<span class="hlt">g</span>D27 were stable when the virus was passaged in cell culture and during acute infection of mice. HSV<span class="hlt">2</span>-<span class="hlt">g</span>D27 was attenuated in mice when it was inoculated onto the cornea, intramuscularly (i.m.), intravaginally, and intracranially. Vaccination of mice i.m. with HSV<span class="hlt">2</span>-<span class="hlt">g</span>D27 provided better inhibition of challenge virus replication in the vagina than when the virus was used to vaccinate mice intranasally or subcutaneously. Comparison of i.m. vaccinations with HSV<span class="hlt">2</span>-<span class="hlt">g</span>D27 versus <span class="hlt">g</span>D<span class="hlt">2</span>t in adjuvant showed that HSV<span class="hlt">2</span>-<span class="hlt">g</span>D27 <span class="hlt">induced</span> larger reductions of challenge virus replication in the vagina and reduced latent viral loads in dorsal root ganglia but <span class="hlt">induced</span> lower serum neutralizing antibody titers than those obtained with <span class="hlt">g</span>D<span class="hlt">2</span>t in adjuvant. Taken together, our data indicate that a live attenuated HSV-<span class="hlt">2</span> vaccine impaired for infection of neurons provides better protection from vaginal challenge with HSV-<span class="hlt">2</span> than that obtained with a subunit vaccine, despite <span class="hlt">inducing</span> lower titers of HSV-<span class="hlt">2</span> neutralizing antibodies in the serum. Genital herpes simplex is one of the most prevalent sexually transmitted diseases. Though HSV-<span class="hlt">2</span> disease is usually mild, it can be life threatening in neonates and immunocompromised persons. In addition, genital</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11835677','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11835677"><span>Heavy ion-<span class="hlt">induced</span> DNA double-strand <span class="hlt">breaks</span> in yeast.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kiefer, Jürgen; Egenolf, Ralf; Ikpeme, Samuel</p> <p>2002-02-01</p> <p>Induction of DSBs in the diploid yeast, Saccharomyces cerevisiae, was measured by pulsed-field gel electrophoresis (PFGE) after the cells had been exposed on membrane filters to a variety of energetic heavy ions with values of linear energy transfer (LET) ranging from about <span class="hlt">2</span> to 11,500 keV/microm, (241)Am alpha particles, and 80 keV X rays. After irradiation, the cells were lysed, and the chromosomes were separated by PFGE. The gels were stained with ethidium bromide, placed on a UV transilluminator, and analyzed using a computer-coupled camera. The fluorescence intensities of the larger bands were found to decrease exponentially with dose or particle fluence. The slope of this line corresponds to the cross section for at least one double-strand <span class="hlt">break</span> (DSB), but closely spaced multiple <span class="hlt">breaks</span> cannot be discriminated. Based on the known size of the native DNA molecules, breakage cross sections per base pair were calculated. They increased with LET until they reached a transient plateau value of about 6 x 10(-7) microm(<span class="hlt">2</span>) at about 300-2000 keV/microm; they then rose for the higher LETs, probably reflecting the influence of delta electrons. The relative biological effectiveness for DNA breakage displays a maximum of about <span class="hlt">2</span>.5 around 100-200 keV/microm and falls below unity for LET values above 10(3) keV/microm. For these yeast cells, comparison of the derived breakage cross sections with the corresponding cross section for inactivation derived from the terminal slope of the survival curves shows a strong linear relationship between these cross sections, extending over several orders of magnitude.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21255959-lupeol-induces-p53-cyclin-mediated-g2-arrest-targets-apoptosis-through-activation-caspase-mouse-skin','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21255959-lupeol-induces-p53-cyclin-mediated-g2-arrest-targets-apoptosis-through-activation-caspase-mouse-skin"><span>Lupeol <span class="hlt">induces</span> p53 and cyclin-B-mediated <span class="hlt">G</span><span class="hlt">2</span>/M arrest and targets apoptosis through activation of caspase in mouse skin</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Nigam, Nidhi; Prasad, Sahdeo; George, Jasmine</p> <p>2009-04-03</p> <p>Lupeol, present in fruits and medicinal plants, is a biologically active compound that has been shown to have various pharmacological properties in experimental studies. In the present study, we demonstrated the modulatory effect of lupeol on 7,12-dimethylbenz[a]anthracene (DMBA)-<span class="hlt">induced</span> alterations on cell proliferation in the skin of Swiss albino mice. Lupeol treatment showed significant (p < 0.05) preventive effects with marked inhibition at 48, 72, and 96 h against DMBA-mediated neoplastic events. Cell-cycle analysis showed that lupeol-<span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span>/M-phase arrest (16-37%) until 72 h, and these inhibitory effects were mediated through inhibition of the cyclin-B-regulated signaling pathway involving p53, p21/WAF1, cdc25C, cdc<span class="hlt">2</span>,more » and cyclin-B gene expression. Further lupeol-<span class="hlt">induced</span> apoptosis was observed, as shown by an increased sub-<span class="hlt">G</span>1 peak (28%) at 96 h, with upregulation of bax and caspase-3 genes and downregulation of anti-apoptotic bcl-<span class="hlt">2</span> and survivin genes. Thus, our results indicate that lupeol has novel anti-proliferative and apoptotic potential that may be helpful in designing strategies to fight skin cancer.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1009291','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1009291"><span>Chromosome analysis from peripheral blood lymphocytes of workers after an acute exposure to benzene.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Clare, M G; Yardley-Jones, A; Maclean, A C; Dean, B J</p> <p>1984-01-01</p> <p>A spillage of about 1200 gallons of benzene occurred during the loading of a ship, and 10 workers on a single shift were exposed to benzene. Shortly afterwards, an assay of the urine of these individuals showed that substantial amounts of phenol were being excreted. About three months after the incident samples of venous blood were taken from 10 individuals exposed to benzene and 11 men on a comparable shift who acted as controls. The lymphocytes were stimulated to divide in short term cultures. For each subject, 200 cells at metaphase were examined for chromosome damage using 48 h cultures, and sister <span class="hlt">chromatid</span> exchanges (SCE) were analysed from about 30 cells in their second division, using 72 h cultures. The most frequent types of aberrations in all the individuals were <span class="hlt">chromatid</span> gaps, with occasional <span class="hlt">breaks</span> of <span class="hlt">chromatids</span> and chromosomes. There were few exchanges within or between the arms of <span class="hlt">chromatids</span> or chromosomes. More cells in the control than in the exposed group showed damage, an effect that was especially noticeable for <span class="hlt">chromatid</span> gaps. All values, however, were considered to be within a normal range. There were slightly more SCE in some of the exposed individuals than in the controls and there was a trend towards a positive association between the frequency of SCE recorded for each individual and the maximum value for the excretion of phenol in the urine on the day after the incident. There is no evidence to indicate that benzene <span class="hlt">induced</span> any type of lasting chromosome damage in the lymphocytes of the 10 exposed workers when cells were examined about three months after the incident. PMID:6722051</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22285401-methoxyphenyl-trimethoxyphenyl-methanone-inhibits-tubulin-polymerization-induces-sub-arrest-triggers-apoptosis-human-leukemia-hl-cells','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22285401-methoxyphenyl-trimethoxyphenyl-methanone-inhibits-tubulin-polymerization-induces-sub-arrest-triggers-apoptosis-human-leukemia-hl-cells"><span>(4-Methoxyphenyl)(3,4,5-trimethoxyphenyl)methanone inhibits tubulin polymerization, <span class="hlt">induces</span> <span class="hlt">G</span>{sub <span class="hlt">2</span>}/M arrest, and triggers apoptosis in human leukemia HL-60 cells</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Magalhães, Hemerson I.F.; Centro de Ciências da Saúde, Departamento de Ciências Farmacêuticas, Universidade Federal da Paraíba, João Pessoa, Paraíba; Wilke, Diego V.</p> <p>2013-10-01</p> <p>(4-Methoxyphenyl)(3,4,5-trimethoxyphenyl)methanone (PHT) is a known cytotoxic compound belonging to the phenstatin family. However, the exact mechanism of action of PHT-<span class="hlt">induced</span> cell death remains to be determined. The aim of this study was to investigate the mechanisms underlying PHT-<span class="hlt">induced</span> cytotoxicity. We found that PHT displayed potent cytotoxicity in different tumor cell lines, showing IC{sub 50} values in the nanomolar range. Cell cycle arrest in <span class="hlt">G</span>{sub <span class="hlt">2</span>}/M phase along with the augmented metaphase cells was found. Cells treated with PHT also showed typical hallmarks of apoptosis such as cell shrinkage, chromatin condensation, phosphatidylserine exposure, increase of the caspase 3/7 and 8 activation,more » loss of mitochondrial membrane potential, and internucleosomal DNA fragmentation without affecting membrane integrity. Studies conducted with isolated tubulin and docking models confirmed that PHT binds to the colchicine site and interferes in the polymerization of microtubules. These results demonstrated that PHT inhibits tubulin polymerization, arrests cancer cells in <span class="hlt">G</span>{sub <span class="hlt">2</span>}/M phase of the cell cycle, and <span class="hlt">induces</span> their apoptosis, exhibiting promising anticancer therapeutic potential. - Highlights: • PHT inhibits tubulin polymerization. • PHT arrests cancer cells in <span class="hlt">G</span>{sub <span class="hlt">2</span>}/M phase of the cell cycle. • PHT <span class="hlt">induces</span> caspase-dependent apoptosis.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/443837-adaptation-tdt-assay-semi-quantitative-flow-cytometric-detection-dna-strand-breaks','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/443837-adaptation-tdt-assay-semi-quantitative-flow-cytometric-detection-dna-strand-breaks"><span>Adaptation of the TdT assay for semi-quantitative flow cytometric detection of DNA strand <span class="hlt">breaks</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Bromidge, T.J.; Howe, D.J.; Johnson, S.A.</p> <p></p> <p>The enzyme Terminal Deoxynucleotidyl Transferase (TdT) is a DNA polymerase which can be used to label DNA strand <span class="hlt">breaks</span> by the incorporation of a labelled nucleotide followed by a fluorescent detection step. The amount of label incorporated can then be assessed by flow cytometry. The mechanism of action of TdT, however, will allow the addition of varying numbers of nucleotides to the free 3{prime} termini produced by DNA strand <span class="hlt">breaks</span>. The substitution of Digoxigenin (DIG){trademark} labelled dideoxynucleotides for labelled deoxy-nucleotides in the TdT assay will limit the addition of label to a DNA <span class="hlt">break</span> to a single nucleotide, thus ensuringmore » a direct relationship between an increase in DNA strand <span class="hlt">breaks</span> and an increase in fluorescence. We have used this adaptation of the TdT assay to evaluate DNA damage incurred in lymphocytes, from patients with Chronic Lymphocytic Leukemia (CLL), on exposure to UV irradiation and apoptosis-<span class="hlt">inducing</span> drugs, fludarabine and <span class="hlt">2</span>-Chloro-<span class="hlt">2</span>{prime}-deoxyadenosine (<span class="hlt">2</span>-CdA). This technique may give a good indication of the susceptibility of CLL patients to apoptosis <span class="hlt">inducing</span> drugs, and hence an indication of the likely response to these therapies. 7 refs., <span class="hlt">2</span> figs., <span class="hlt">2</span> tabs.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22666058-gigaelectronvolt-counterpart-ver-j2019+407-northern-shell-supernova-remnant-g78-cygni','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22666058-gigaelectronvolt-counterpart-ver-j2019+407-northern-shell-supernova-remnant-g78-cygni"><span>THE GIGAELECTRONVOLT COUNTERPART OF VER J2019+407 IN THE NORTHERN SHELL OF THE SUPERNOVA REMNANT <span class="hlt">G</span>78.<span class="hlt">2+2</span>.1 ( γ Cygni)</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Fraija, N.; Araya, M., E-mail: nifraija@astro.unam.mx, E-mail: miguel.araya@ucr.ac.cr</p> <p>2016-07-20</p> <p>Analysis of gamma-ray emission from the supernova remnant <span class="hlt">G</span>78.<span class="hlt">2+2</span>.1 ( γ Cygni) with 7.<span class="hlt">2</span> years of cumulative data from the Fermi Large Area Telescope shows a distinct hard, bright, and extended component to the north of the shell coincident with the known teraelectronvolt source VER J2019+407. In the gigaelectronvolt to teraelectronvolt (GeV–TeV) energy range, its spectrum is best described by a broken power law with indices 1.8 below a <span class="hlt">break</span> energy of 71 GeV and <span class="hlt">2</span>.5 above the <span class="hlt">break</span>. A broadband spectral energy distribution is assembled, and different scenarios for the origin of the gamma rays are explored. Both hadronicmore » and leptonic mechanisms are able to account for the GeV–TeV observations. In the leptonic framework, a superposition of inverse Compton and nonthermal bremsstrahlung emissions is needed, whereas the hadronic scenario requires a cosmic-ray population described by a broken power-law distribution with a relatively hard spectral index of ∼1.8 below a <span class="hlt">break</span> particle energy of 0.45 TeV. In addition, the neutrino flux expected from cosmic-ray interactions is calculated.« less</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li class="active"><span>22</span></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_22 --> <div id="page_23" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li class="active"><span>23</span></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="441"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15766318','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15766318"><span>Different susceptibility of cells of porcine skin and internal organs to ultraviolet A-<span class="hlt">induced</span> <span class="hlt">breaking</span> of nuclear DNA.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Brozyna, Anna; Chwirot, Barbara W</p> <p>2005-01-01</p> <p>There is a continuously growing interest in medical applications of ultraviolet radiation (UV-A and long-wavelength UV-B) especially for laser surgery, phototherapy and photodiagnostics of human internal organs. UV-B and UV-A radiation is potentially mutagenic, however, there has been very little information published to date concerning the significance of possible deleterious action of such photons on cells of internal tissues. The aim of this study is to compare the sensitivities of skin cells to those of internal organs upon exposure to UV-A. To assess this sensitivity we have determined the UV-A dose-dependent frequency of nuclear DNA <span class="hlt">breaks</span> detected with the terminal deoxynucleotidyl transferase-mediated deoxyuridine triphosphate-biotin nick end-labeling (TUNEL) technique. The materials for the study were macroscopic samples of porcine skin, colon and esophagus. The UV-A dose ranged from 0.1 to 1000 mJ/cm<span class="hlt">2</span>, which is similar to doses received by cells in regions examined with laser-<span class="hlt">induced</span> fluorescence or by cells surrounding areas subject to a laser ablation. To reduce the influence of DNA repair processes the tissue samples were kept at a low temperature during the irradiation and were deep frozen immediately after completing the irradiation procedure. The cells of the internal organs are much more susceptible to UV-A-<span class="hlt">induced</span> <span class="hlt">breaking</span> of DNA than the skin cells. The percentage fractions and the spatial distributions of the damaged cells and the characteristics of the UV-A dose dependence seem to vary by type of internal organ.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28387973','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28387973"><span>Fusaric Acid <span class="hlt">Induces</span> DNA Damage and Post-Translational Modifications of p53 in Human Hepatocellular Carcinoma (Hep<span class="hlt">G</span><span class="hlt">2</span> ) Cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ghazi, Terisha; Nagiah, Savania; Tiloke, Charlette; Sheik Abdul, Naeem; Chuturgoon, Anil A</p> <p>2017-11-01</p> <p>Fusaric acid (FA), a common fungal contaminant of maize, is known to mediate toxicity in plants and animals; however, its mechanism of action is unclear. p53 is a tumor suppressor protein that is activated in response to cellular stress. The function of p53 is regulated by post-translational modifications-ubiquitination, phosphorylation, and acetylation. This study investigated a possible mechanism of FA <span class="hlt">induced</span> toxicity in the human hepatocellular carcinoma (Hep<span class="hlt">G</span> <span class="hlt">2</span> ) cell line. The effect of FA on DNA integrity and post-translational modifications of p53 were investigated. Methods included: (a) culture and treatment of Hep<span class="hlt">G</span> <span class="hlt">2</span> cells with FA (IC 50 : 580.32 μM, 24 h); (b) comet assay (DNA damage); (c) Western blots (protein expression of p53, MDM<span class="hlt">2</span>, p-Ser-15-p53, a-K382-p53, a-CBP (K1535)/p300 (K1499), HDAC1 and p-Ser-47-Sirt1); and (d) Hoechst 33342 assay (apoptosis analysis). FA caused DNA damage in Hep<span class="hlt">G</span> <span class="hlt">2</span> cells relative to the control (P < 0.0001). FA decreased the protein expression of p53 (0.24-fold, P = 0.0004) and increased the expression of p-Ser-15-p53 (12.74-fold, P = 0.0126) and a-K382-p53 (<span class="hlt">2</span>.24-fold, P = 0.0096). This occurred despite the significant decrease in the histone acetyltransferase, a-CBP (K1535)/p300 (K1499) (0.42-fold, P = 0.0023) and increase in the histone deacetylase, p-Ser-47-Sirt1 (1.22-fold, P = 0.0020). The expression of MDM<span class="hlt">2</span>, a negative regulator of p53, was elevated in the FA treatment compared to the control (1.83-fold, P < 0.0001). FA also inhibited cell proliferation and <span class="hlt">induced</span> apoptosis in Hep<span class="hlt">G</span> <span class="hlt">2</span> cells as evidenced by the Hoechst assay. Together, these results indicate that FA is genotoxic and post-translationally modified p53 leading to Hep<span class="hlt">G</span> <span class="hlt">2</span> cell death. J. Cell. Biochem. 118: 3866-3874, 2017. © 2017 Wiley Periodicals, Inc. © 2017 Wiley Periodicals, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18817800','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18817800"><span>Biguanide-<span class="hlt">induced</span> mitochondrial dysfunction yields increased lactate production and cytotoxicity of aerobically-poised Hep<span class="hlt">G</span><span class="hlt">2</span> cells and human hepatocytes in vitro.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Dykens, James A; Jamieson, Joseph; Marroquin, Lisa; Nadanaciva, Sashi; Billis, Puja A; Will, Yvonne</p> <p>2008-12-01</p> <p>As a class, the biguanides <span class="hlt">induce</span> lactic acidosis, a hallmark of mitochondrial impairment. To assess potential mitochondrial impairment, we evaluated the effects of metformin, buformin and phenformin on: 1) viability of Hep<span class="hlt">G</span><span class="hlt">2</span> cells grown in galactose, <span class="hlt">2</span>) respiration by isolated mitochondria, 3) metabolic poise of Hep<span class="hlt">G</span><span class="hlt">2</span> and primary human hepatocytes, 4) activities of immunocaptured respiratory complexes, and 5) mitochondrial membrane potential and redox status in primary human hepatocytes. Phenformin was the most cytotoxic of the three with buformin showing moderate toxicity, and metformin toxicity only at mM concentrations. Importantly, Hep<span class="hlt">G</span><span class="hlt">2</span> cells grown in galactose are markedly more susceptible to biguanide toxicity compared to cells grown in glucose, indicating mitochondrial toxicity as a primary mode of action. The same rank order of potency was observed for isolated mitochondrial respiration where preincubation (40 min) exacerbated respiratory impairment, and was required to reveal inhibition by metformin, suggesting intramitochondrial bio-accumulation. Metabolic profiling of intact cells corroborated respiratory inhibition, but also revealed compensatory increases in lactate production from accelerated glycolysis. High (mM) concentrations of the drugs were needed to inhibit immunocaptured respiratory complexes, supporting the contention that bioaccumulation is involved. The same rank order was found when monitoring mitochondrial membrane potential, ROS production, and glutathione levels in primary human hepatocytes. In toto, these data indicate that biguanide-<span class="hlt">induced</span> lactic acidosis can be attributed to acceleration of glycolysis in response to mitochondrial impairment. Indeed, the desired clinical outcome, viz., decreased blood glucose, could be due to increased glucose uptake and glycolytic flux in response to drug-<span class="hlt">induced</span> mitochondrial dysfunction.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018CoTPh..69..585L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018CoTPh..69..585L"><span>New Double-Periodic Soliton Solutions for the (<span class="hlt">2</span>+1)-Dimensional <span class="hlt">Breaking</span> Soliton Equation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Liu, Jian-Guo; Tian, Yu</p> <p>2018-05-01</p> <p>Under investigation is the (<span class="hlt">2</span>+1)-dimensional <span class="hlt">breaking</span> soliton equation. Based on a special ansätz functions and the bilinear form, some entirely new double-periodic soliton solutions for the (<span class="hlt">2</span>+1)-dimensional <span class="hlt">breaking</span> soliton equation are presented. With the help of symbolic computation software Mathematica, many important and interesting properties for these obtained solutions are revealed with some figures. Supported by National Natural Science Foundation of China under Grant No. 61377067</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3391032','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3391032"><span>Comparisons of Serum Total IgE, Ig<span class="hlt">G</span>, and Ig<span class="hlt">G</span>1 Levels in Patients with and without Echinococcosis-<span class="hlt">Induced</span> Anaphylactic Shock</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Li, Yimei; Zheng, Hong; Gu, Meilin; Cao, Xinghua; Wen, Hao; Liu, Zaoling; Liu, Tao</p> <p>2012-01-01</p> <p>We investigated serum total immunoglobulin E (IgE), Ig<span class="hlt">G</span>, and Ig<span class="hlt">G</span>1 levels in patients with and without echinococcosis-<span class="hlt">induced</span> anaphylactic shock. This was a case-control study of 11 patients with echinococcosis-<span class="hlt">induced</span> anaphylactic shock and 22 echinococcosis patients with cyst rupture but without anaphylactic shock. Blood was collected before surgery (T0), at the time of cyst rupture (T1), and shock (Tx), 1 h (T<span class="hlt">2</span>), 1 day (T3), and 1 week (T4) after cyst rupture. Serum IgE, Ig<span class="hlt">G</span>, and Ig<span class="hlt">G</span>1 were determined by enzyme-linked immunosorbent assay. Serum IgE, Ig<span class="hlt">G</span>, and Ig<span class="hlt">G</span>1 levels were significantly higher in patients who developed anaphylactic shock at all time points. Increased pre-surgical Ig<span class="hlt">G</span> and Ig<span class="hlt">G</span>1 levels were identified to be a significant risk factors for developing anaphylactic shock. The results showed that a serum Ig<span class="hlt">G</span> concentration of 312.25 μ<span class="hlt">g</span>/mL could be used as a cut-off point to predict whether an echinococcosis patient would develop anaphylactic shock. PMID:22764299</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18405831','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18405831"><span>The 20-hydroxyecdysone-<span class="hlt">induced</span> signalling pathway in <span class="hlt">G</span><span class="hlt">2</span>/M arrest of Plodia interpunctella imaginal wing cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Siaussat, David; Bozzolan, Françoise; Porcheron, Patrick; Debernard, Stéphane</p> <p>2008-05-01</p> <p>The mechanisms involved in the control of cellular proliferation by the steroid hormone 20-hydroxyecdysone (20E) in insects are not known. We dissected the 20E signalling pathway responsible for <span class="hlt">G</span><span class="hlt">2</span>/M arrest of imaginal cells from the IAL-PID<span class="hlt">2</span> cells of the Indian meal moth Plodia interpunctella. We first used a 5'-3' RACE-based strategy to clone a 4479bp cDNA encoding a putative P. interpunctella HR3 transcription factor named PiHR3. The deduced amino acid sequence of PiHR3 was highly similar to those of HR3 proteins from other lepidopterans, e.<span class="hlt">g</span>. Manduca sexta and Bombyx mori. Using double-stranded RNA-mediated interference (dsRNAi), we then succeeded in blocking the ability of 20E to <span class="hlt">induce</span> the expression of PiEcR-B1, PiUSP-<span class="hlt">2</span> and PiHR3 genes that encode the P. interpunctella ecdysone receptor B1-isoform, Ultraspiracle-<span class="hlt">2</span> isoform, the insect homologue of the vertebrate retinoid X receptor, and the HR3 transcription factor. We showed that inhibiting the 20E induction of PiEcR-B1, PiUSP-<span class="hlt">2</span> and PiHR3 mRNAs prevented the decreased expression of B cyclin and consequently the <span class="hlt">G</span><span class="hlt">2</span>/M arrest of IAL-PID<span class="hlt">2</span> cells. Using this functional approach, we revealed the participation of EcR, USP and HR3 in a 20E signalling pathway that controls the proliferation of imaginal cells by regulating the expression of B cyclin.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4287329','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4287329"><span>Protective Effects of Black Rice Extracts on Oxidative Stress <span class="hlt">Induced</span> by tert-Butyl Hydroperoxide in Hep<span class="hlt">G</span><span class="hlt">2</span> Cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Lee, Seon-Mi; Choi, Youngmin; Sung, Jeehye; Kim, Younghwa; Jeong, Heon-Sang; Lee, Junsoo</p> <p>2014-01-01</p> <p>Black rice contains many biologically active compounds. The aim of this study was to investigate the protective effects of black rice extracts (whole grain extract, WGE and rice bran extract, RBE) on tert-butyl hydroperoxide (TBHP)-<span class="hlt">induced</span> oxidative injury in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Cellular reactive oxygen species (ROS), antioxidant enzyme activities, malondialdehyde (MDA) and glutathione (GSH) concentrations were evaluated as biomarkers of cellular oxidative status. Cells pretreated with 50 and 100 μ<span class="hlt">g</span>/mL of WGE or RBE were more resistant to oxidative stress in a dose-dependent manner. The highest WGE and BRE concentrations enhanced GSH concentrations and modulated antioxidant enzyme activities (glutathione reductase, glutathione-S-transferase, catalase, and superoxide dismutase) compared to TBHP-treated cells. Cells treated with RBE showed higher protective effect compared to cells treated with WGE against oxidative insult. Black rice extracts attenuated oxidative insult by inhibiting cellular ROS and MDA increase and by modulating antioxidant enzyme activities in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. PMID:25580401</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27093810','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27093810"><span>[Stimuli phrases of adductor spasmodic dysphonia phonatory <span class="hlt">break</span> in mandarin Chinese].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ge, Pingjiang; Ren, Qingyi; Chen, Zhipeng; Cheng, Qiuhui; Sheng, Xiaoli; Wang, Ling; Chen, Shaohua; Zhang, Siyi</p> <p>2015-12-01</p> <p>To investigate the characteristics of adductor spasmodic dysphonia phonatory <span class="hlt">break</span> in mandarin Chinese and select the stimuli phrases. Thirty-eight patients with adductor spasmodic dysphonia were involved in this study. Standard phrase " fù mŭ xīn" and a speech corpus in mandarin Chinese with 229 syllables covering all vowel and constant of mandarin Chinese were selected. Every patient read the phrases above twice in normal speed and comfortable voice. Two auditory perpetual speech pathologists marked phonatory <span class="hlt">break</span> syllables respectively. The frequency of phonatory <span class="hlt">break</span> syllables and their located phrases were calculated, rated and described. The phrases including the most phonatory <span class="hlt">break</span> syllables were selected as stimuli phrases, the phonatory <span class="hlt">break</span> frequency of which was also higher than that of standard phrase "fù mŭ xīn". Phonatory <span class="hlt">break</span> happened in the reading of all patients. The average number of phonatory <span class="hlt">break</span> syllables was 14 (3-33). Phonatroy <span class="hlt">break</span> occurred when saying 177 (77.3%) syllables in the speech corpus. The syllables "guŏ, rén, zāng, diàn, chē, <span class="hlt">g</span>è, guăn, a, bā, ne, de" broke in 23.1%-41.0% patients. These syllables belonged to the phrases "pĭng guŏ, huŏ chē, shì de, nĭ shì <span class="hlt">g</span>è hăo rén, wŏ mén shì yŏu zŏng shì bă qĭn shì nong dé hĕn zāng, wŏ mén nà biān yŏu wăng qiú yùn dong chăng, cān <span class="hlt">g</span>ŭan, jiŭ bā hé yī <span class="hlt">g</span>è miàn bāo dìan, tā shì duō me kāng kăi a,wŏ yīng <span class="hlt">g</span>āi zài xìn lĭ xiĕ yī xiē shén mē ne?". Thirty-seven patients (97.3%) had phonatory <span class="hlt">break</span> in above mentioned words. Ratios of these words phonatory <span class="hlt">break</span> also were more than "fù mŭ xīn". Adductor spasmodic dysphonic patients exhibited different degrees of phonatory <span class="hlt">break</span> in mandarine Chinese. The phrases" shì de, pĭng guŏ, huŏ chē, nĭ shì <span class="hlt">g</span>è hăo rén, wŏ mén nà biān yŏu wăng qiú yùn dong chăng, cān <span class="hlt">g</span>ŭan, jiŭ bā hé yī <span class="hlt">g</span>è miàn bāo dìan, tā shì duō me kāng kăi a" were recommended as stimuli</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26461344','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26461344"><span>Investigating free radical generation in Hep<span class="hlt">G</span><span class="hlt">2</span> cells using immuno-spin trapping.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Horinouchi, Yuya; Summers, Fiona A; Ehrenshaft, Marilyn; Kawazoe, Kazuyoshi; Tsuchiya, Koichiro; Tamaki, Toshiaki; Mason, Ronald P</p> <p>2014-10-01</p> <p>Oxidative stress can <span class="hlt">induce</span> the generation of free radicals, which are believed to play an important role in both physiological and pathological processes and a number of diseases such as cancer. Therefore, it is important to identify chemicals which are capable of <span class="hlt">inducing</span> oxidative stress. In this study, we evaluated the ability of four environmental chemicals, aniline, nitrosobenzene (NB), N,N-dimethylaniline (DMA) and N,N-dimethyl-4-nitrosoaniline (DMNA), to <span class="hlt">induce</span> free radicals and cellular damage in the hepatoma cell line Hep<span class="hlt">G</span><span class="hlt">2</span>. Cytotoxicity was assessed using lactate dehydrogenase (LDH) assays and morphological changes were observed using phase contrast microscopy. Free radicals were detected by immuno-spin trapping (IST) in in-cell western experiments or in confocal microscopy experiments to determine the subcellular localization of free radical generation. DMNA <span class="hlt">induced</span> free radical generation, LDH release and morphological changes in Hep<span class="hlt">G</span><span class="hlt">2</span> cells whereas aniline, NB and DMA did not. Confocal microscopy showed that DMNA <span class="hlt">induced</span> free radical generation mainly in the cytosol. Preincubation of Hep<span class="hlt">G</span><span class="hlt">2</span> cells with N-acetylcysteine and <span class="hlt">2,2</span>'-dipyridyl significantly prevented free radical generation upon subsequent incubation with DMNA, whereas preincubation with apocynin and dimethyl sulfoxide did not. These results suggest that DMNA <span class="hlt">induces</span> oxidative stress and that reactive oxygen species, metals and free radical generation play a critical role in DMNA-<span class="hlt">induced</span> cytotoxicity. Copyright © 2014. Published by Elsevier Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22170698','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22170698"><span>Restoration of C/EBPα in dedifferentiated liposarcoma <span class="hlt">induces</span> <span class="hlt">G</span><span class="hlt">2</span>/M cell cycle arrest and apoptosis.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wu, Yuhsin V; Okada, Tomoyo; DeCarolis, Penelope; Socci, Nicholas; O'Connor, Rachael; Geha, Rula C; Joy Somberg, C; Antonescu, Cristina; Singer, Samuel</p> <p>2012-04-01</p> <p>Well-differentiated liposarcoma (WDLS) and dedifferentiated liposarcoma (DDLS) represent the most common biological group of liposarcoma, and there is a pressing need to develop targeted therapies for patients with advanced disease. To identify potential therapeutic targets, we sought to identify differences in the adipogenic pathways between DDLS, WDLS, and normal adipose tissue. In a microarray analysis of DDLS (n = 84), WDLS (n = 79), and normal fat (n = 23), C/EBPα, a transcription factor involved in cell cycle regulation and differentiation, was underexpressed in DDLS when compared to both WDLS and normal fat (15.<span class="hlt">2</span>- and 27.8-fold, respectively). In normal adipose-derived stem cells, C/EBPα expression was strongly <span class="hlt">induced</span> when cells were cultured in differentiation media, but in three DDLS cell lines, this induction was nearly absent. We restored C/EBPα expression in one of the cell lines (DDLS8817) by transfection of an <span class="hlt">inducible</span> C/EBPα expression vector. <span class="hlt">Inducing</span> C/EBPα expression reduced proliferation and caused cells to accumulate in <span class="hlt">G</span><span class="hlt">2</span>/M. Under differentiation conditions, the cell proliferation was reduced further, and 66% of the DDLS cells containing the <span class="hlt">inducible</span> C/EBPα expression vector underwent apoptosis as demonstrated by annexin V staining. These cells in differentiation conditions expressed early adipocyte-specific mRNAs such as LPL and FABP4, but they failed to accumulate intracellular lipid droplets, a characteristic of mature adipocytes. These results demonstrate that loss of C/EBPα is an important factor in suppressing apoptosis and maintaining the dedifferentiated state in DDLS. Restoring C/EBPα may be a useful therapeutic approach for DDLS. Copyright © 2011 Wiley Periodicals, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4426816','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4426816"><span>Sites of instability in the human TCF3 (E<span class="hlt">2</span>A) gene adopt <span class="hlt">G</span>-quadruplex DNA structures in vitro</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Williams, Jonathan D.; Fleetwood, Sara; Berroyer, Alexandra; Kim, Nayun; Larson, Erik D.</p> <p>2015-01-01</p> <p>The formation of highly stable four-stranded DNA, called <span class="hlt">G</span>-quadruplex (<span class="hlt">G</span>4), promotes site-specific genome instability. <span class="hlt">G</span>4 DNA structures fold from repetitive guanine sequences, and increasing experimental evidence connects <span class="hlt">G</span>4 sequence motifs with specific gene rearrangements. The human transcription factor 3 (TCF3) gene (also termed E<span class="hlt">2</span>A) is subject to genetic instability associated with severe disease, most notably a common translocation event t(1;19) associated with acute lymphoblastic leukemia. The sites of instability in TCF3 are not randomly distributed, but focused to certain sequences. We asked if <span class="hlt">G</span>4 DNA formation could explain why TCF3 is prone to recombination and mutagenesis. Here we demonstrate that sequences surrounding the major t(1;19) <span class="hlt">break</span> site and a region associated with copy number variations both contain <span class="hlt">G</span>4 sequence motifs. The motifs identified readily adopt <span class="hlt">G</span>4 DNA structures that are stable enough to interfere with DNA synthesis in physiological salt conditions in vitro. When introduced into the yeast genome, TCF3 <span class="hlt">G</span>4 motifs promoted gross chromosomal rearrangements in a transcription-dependent manner. Our results provide a molecular rationale for the site-specific instability of human TCF3, suggesting that <span class="hlt">G</span>4 DNA structures contribute to oncogenic DNA <span class="hlt">breaks</span> and recombination. PMID:26029241</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001NuPhB.592..164C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001NuPhB.592..164C"><span>RG-invariant sum rule in a generalization of anomaly-mediated SUSY-<span class="hlt">breaking</span> models</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Carena, Marcela; Huitu, Katri; Kobayashi, Tatsuo</p> <p>2001-01-01</p> <p>We study a generalization of anomaly-mediated supersymmetry-<span class="hlt">breaking</span> (AMSB) scenarios, under the assumption that the effects of the high-scale theory do not completely decouple and that D-term type contributions can therefore be present. We investigate the effect of such possible D-term additional contributions to soft scalar masses by requiring that, for non-vanishing, renormalizable Yukawa couplings Yijk, the sum of squared soft supersymmetry <span class="hlt">breaking</span> mass parameters, M<span class="hlt">2</span>ijk≡mi<span class="hlt">2</span>+mj<span class="hlt">2</span>+mk<span class="hlt">2</span>, is RG-invariant, in the sense that it becomes independent of the specific ultraviolet boundary conditions as it occurs in the AMSB models. This type of models can avoid the problem of tachyonic solutions for the slepton mass spectrum present in AMSB scenarios. We implement the electroweak symmetry <span class="hlt">breaking</span> condition and explore the sparticle spectrum associated with this framework. To show the possible diversity of the sparticle spectrum, we consider two examples, one in which the D-terms <span class="hlt">induce</span> a common soft supersymmetry <span class="hlt">breaking</span> mass term for all sfermion masses, and another one in which a light stop can be present in the spectrum.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3681665','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3681665"><span>Fragile DNA Motifs Trigger Mutagenesis at Distant Chromosomal Loci in Saccharomyces cerevisiae</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Saini, Natalie; Zhang, Yu; Nishida, Yuri; Sheng, Ziwei; Choudhury, Shilpa; Mieczkowski, Piotr; Lobachev, Kirill S.</p> <p>2013-01-01</p> <p>DNA sequences capable of adopting non-canonical secondary structures have been associated with gross-chromosomal rearrangements in humans and model organisms. Previously, we have shown that long inverted repeats that form hairpin and cruciform structures and triplex-forming GAA/TTC repeats <span class="hlt">induce</span> the formation of double-strand <span class="hlt">breaks</span> which trigger genome instability in yeast. In this study, we demonstrate that breakage at both inverted repeats and GAA/TTC repeats is augmented by defects in DNA replication. Increased fragility is associated with increased mutation levels in the reporter genes located as far as 8 kb from both sides of the repeats. The increase in mutations was dependent on the presence of inverted or GAA/TTC repeats and activity of the translesion polymerase Polζ. Mutagenesis <span class="hlt">induced</span> by inverted repeats also required Sae<span class="hlt">2</span> which opens hairpin-capped <span class="hlt">breaks</span> and initiates end resection. The amount of breakage at the repeats is an important determinant of mutations as a perfect palindromic sequence with inherently increased fragility was also found to elevate mutation rates even in replication-proficient strains. We hypothesize that the underlying mechanism for mutagenesis <span class="hlt">induced</span> by fragile motifs involves the formation of long single-stranded regions in the broken chromosome, invasion of the undamaged sister <span class="hlt">chromatid</span> for repair, and faulty DNA synthesis employing Polζ. These data demonstrate that repeat-mediated <span class="hlt">breaks</span> pose a dual threat to eukaryotic genome integrity by <span class="hlt">inducing</span> chromosomal aberrations as well as mutations in flanking genes. PMID:23785298</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26301894','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26301894"><span>Recombinant Buckwheat Trypsin Inhibitor <span class="hlt">Induces</span> Mitophagy by Directly Targeting Mitochondria and Causes Mitochondrial Dysfunction in Hep <span class="hlt">G</span><span class="hlt">2</span> Cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Zhuanhua; Li, Shanshan; Ren, Rong; Li, Jiao; Cui, Xiaodong</p> <p>2015-09-09</p> <p>Mitochondria are essential targets for cancer chemotherapy and other disease treatments. Recombinant buckwheat trypsin inhibitor (rBTI), a member of the potato type I proteinase inhibitor family, was derived from tartary buckwheat extracts. Our results showed that rBTI directly targeted mitochondria and <span class="hlt">induced</span> mitochondrial fragmentation and mitophagy. This occurs through enhanced depolarization of the mitochondrial membrane potential, increasing reactive oxygen species (ROS) generation associated with the rise of the superoxide dismutase and catalase activity and glutathione peroxidase (GSH) content, and changes in the GSH/oxidized glutathione ratio. Mild and transient ROS <span class="hlt">induced</span> by rBTI were shown to be important signaling molecules required to <span class="hlt">induce</span> Hep <span class="hlt">G</span><span class="hlt">2</span> mitophagy to remove dysfunctional mitochondria. Furthermore, rBTI could directly <span class="hlt">induce</span> mitochondrial fragmentation. It was also noted that rBTI highly increased colocalization of mitochondria in treated cells compared to nontreated cells. Tom 20, a subunit of the translocase of the mitochondrial outer membrane complex responsible for recognizing mitochondrial presequences, may be the direct target of rBTI.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25601756','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25601756"><span>Interdependence of the rad50 hook and globular domain functions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hohl, Marcel; Kochańczyk, Tomasz; Tous, Cristina; Aguilera, Andrés; Krężel, Artur; Petrini, John H J</p> <p>2015-02-05</p> <p>Rad50 contains a conserved Zn(<span class="hlt">2</span>+) coordination domain (the Rad50 hook) that functions as a homodimerization interface. Hook ablation phenocopies Rad50 deficiency in all respects. Here, we focused on rad50 mutations flanking the Zn(<span class="hlt">2</span>+)-coordinating hook cysteines. These mutants impaired hook-mediated dimerization, but recombination between sister <span class="hlt">chromatids</span> was largely unaffected. This may reflect that cohesin-mediated sister <span class="hlt">chromatid</span> interactions are sufficient for double-strand <span class="hlt">break</span> repair. However, Mre11 complex functions specified by the globular domain, including Tel1 (ATM) activation, nonhomologous end joining, and DNA double-strand <span class="hlt">break</span> end resection were affected, suggesting that dimerization exerts a broad influence on Mre11 complex function. These phenotypes were suppressed by mutations within the coiled-coil and globular ATPase domains, suggesting a model in which conformational changes in the hook and globular domains are transmitted via the extended coils of Rad50. We propose that transmission of spatial information in this manner underlies the regulation of Mre11 complex functions. Copyright © 2015 Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPhA...50.5401A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPhA...50.5401A"><span>Coleman-Weinberg symmetry <span class="hlt">breaking</span> in SU(8) <span class="hlt">induced</span> by a third rank antisymmetric tensor scalar field II: the fermion spectrum</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Adler, Stephen L.</p> <p>2017-07-01</p> <p>We continue our study of Coleman-Weinberg symmetry <span class="hlt">breaking</span> <span class="hlt">induced</span> by a third rank antisymmetric tensor scalar, in the context of the SU(8) model (Adler 2014 Int. J. Mod. Phys. A 29 1450130) we proposed earlier. We focus in this paper on qualitative features that will determine whether the model can make contact with the observed particle spectrum. We discuss the mechanism for giving the spin \\frac{3}{<span class="hlt">2</span>} field a mass by the BEH mechanism, and analyze the remaining massless spin \\frac{1}{<span class="hlt">2</span>} fermions, the global chiral symmetries, and the running couplings after symmetry <span class="hlt">breaking</span>. We note that the smallest gluon mass matrix eigenvalue has an eigenvector suggestive of U(1) B-L , and conjecture that the theory runs to an infrared fixed point at which there is a massless gluon with 3 to  -1 ratios in generator components. Assuming this, we discuss a mechanism for making contact with the standard model, based on a conjectured asymmetric <span class="hlt">breaking</span> of Sp(4) to SU(<span class="hlt">2</span>) subgroups, one of which is the electroweak SU(<span class="hlt">2</span>), and the other of which is a ‘technicolor’ group that binds the original SU(8) model fermions, which play the role of ‘preons’, into composites. Quarks can emerge as 5 preon composites and leptons as 3 preon composites, with consequent stability of the proton against decay to a single lepton plus a meson. A composite Higgs boson can emerge as a two preon composite. Since anomaly matching for the relevant conserved global symmetry current is not obeyed by three fermion families, emergence of three composite families requires formation of a Goldstone boson with quantum numbers matching this current, which can be a light dark matter candidate.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20730658','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20730658"><span>Increased frequency of sister <span class="hlt">chromatid</span> exchanges and decrease in cell viability and proliferation kinetics in human peripheral blood lymphocytes after in vitro exposure to whole bee venom.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gajski, Goran; Garaj-Vrhovac, Vera</p> <p>2010-10-01</p> <p>The present study was aimed to investigate the impact of bee venom on frequency of sister <span class="hlt">chromatid</span> exchanges (SCE) and viability in human peripheral blood lymphocytes in vitro. In addition, the proportion of lymphocytes that undergo one, two or three cell divisions as well as proliferative rate index (PRI) have been determined. Aqueous solution of whole bee venom was added to whole blood samples in concentrations ranging from 0.1 microg/mL to 20 microg/mL in different lengths of time. Results showed that whole bee venom inhibited cell viability, resulting in a 22.86 +/- 1.14% and 51.21 +/- 0.58% reduction of viable cells at 1 hour and 6 hours, respectively. The mean SCE per cell in all the exposed samples was significantly higher than in the corresponding controls. In addition, the percentage of high frequency cells (HFC) for each sample was estimated using the pooled distribution of all SCE measurements. This parameter was also significantly higher compared to the control. Inhibition of proliferation was statistically significant for both exposure times and concentrations and was time and dose dependent. These data indicate that whole bee venom inhibited cell proliferation, resulting in a 36.87 +/- 5.89% and 38.43 +/- 1.96% reduction of proliferation at 1 hour and 6 hours, respectively. In conclusion, this report demonstrated that whole bee venom is capable of <span class="hlt">inducing</span> DNA alterations by virtue of increasing sister <span class="hlt">chromatid</span> exchanges in addition to the cell viability decrease and inhibition of proliferation kinetics in human peripheral blood lymphocytes in vitro.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29280499','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29280499"><span><span class="hlt">G</span>5<span class="hlt">G</span><span class="hlt">2</span>.5 core-shell tecto-dendrimer specifically targets reactive glia in brain ischemia.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Murta, Veronica; Schilrreff, Priscila; Rosciszewski, Gerardo; Morilla, Maria Jose; Ramos, Alberto Javier</p> <p>2018-03-01</p> <p>Secondary neuronal death is a serious stroke complication. This process is facilitated by the conversion of glial cells to the reactive pro-inflammatory phenotype that <span class="hlt">induces</span> neurodegeneration. Therefore, regulation of glial activation is a compelling strategy to reduce brain damage after stroke. However, drugs have difficulties to access the CNS, and to specifically target glial cells. In the present work, we explored the use core-shell polyamidoamine tecto-dendrimer (<span class="hlt">G</span>5<span class="hlt">G</span><span class="hlt">2</span>.5 PAMAM) and studied its ability to target distinct populations of stroke-activated glial cells. We found that <span class="hlt">G</span>5<span class="hlt">G</span><span class="hlt">2</span>.5 tecto-dendrimer is actively engulfed by primary glial cells in a time- and dose-dependent manner showing high cellular selectivity and lysosomal localization. In addition, oxygen-glucose deprivation or lipopolysaccharides exposure in vitro and brain ischemia in vivo increase glial <span class="hlt">G</span>5<span class="hlt">G</span><span class="hlt">2</span>.5 uptake; not being incorporated by neurons or other cell types. We conclude that <span class="hlt">G</span>5<span class="hlt">G</span><span class="hlt">2</span>.5 tecto-dendrimer is a highly suitable carrier for targeted drug delivery to reactive glial cells in vitro and in vivo after brain ischemia. © 2017 International Society for Neurochemistry.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24028616','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24028616"><span>GPER-1 agonist <span class="hlt">G</span>1 <span class="hlt">induces</span> vasorelaxation through activation of epidermal growth factor receptor-dependent signalling pathway.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jang, Eun Jin; Seok, Young Mi; Arterburn, Jeffrey B; Olatunji, Lawrence A; Kim, In Kyeom</p> <p>2013-10-01</p> <p>The <span class="hlt">G</span> protein-coupled oestrogen receptor-1 (GPER-1) agonist <span class="hlt">G</span>1 <span class="hlt">induces</span> endothelium-dependent relaxation. Activation of the epidermal growth factor (EGF) receptor leads to transduction of signals from the plasma membrane for the release of nitric oxide. We tested the hypothesis that <span class="hlt">G</span>1 <span class="hlt">induces</span> endothelium-dependent vasorelaxation through activation of the EGF receptor. Rat aortic rings were mounted in organ baths. After pretreatment with various inhibitors, aortic rings contracted with 11,9-epoxymethano-prostaglandin F<span class="hlt">2</span>α or KCl were subjected to relaxation by <span class="hlt">G</span>1. <span class="hlt">G</span>1 <span class="hlt">induced</span> endothelium-dependent vasorelaxation, which was attenuated by pretreatment with either L -N(ω) -nitroarginine methyl ester (L -NAME), an inhibitor of nitric oxide synthase, or (3aS,4R,9bR)-4-(6-bromo-1,3-benzodioxol-5-yl)-3a,4,5,9b-tetrahydro-3H-cyclopenta[c]quinoline HB-EGF, heparin-binding EGF-like growth factor, a GPER-1 antagonist. Neither a general oestrogen receptor antagonist, ICI 182 780, nor a selective oestrogen receptor-α antagonist, methyl-piperidino-pyrazole dihydrochloride (MPP), had an effect on <span class="hlt">G</span>1-<span class="hlt">induced</span> vasorelaxation. However, pretreatment with EGF receptor blockers, AG1478 or DAPH, resulted in attenuated <span class="hlt">G</span>1-<span class="hlt">induced</span> vasorelaxation. In addition, pretreatment with Src inhibitor 4-amino-3-(4-chlorophenyl)-1-(t-butyl)-1H-pyrazolo[3,4-d]pyrimidine, 4-amino-5-(4-chlorophenyl)-7-(t-butyl)pyrazolo[3,4-d]pyrimidine or Akt inhibitor VIII also resulted in attenuated vascular relaxation <span class="hlt">induced</span> by the cumulative addition of <span class="hlt">G</span>1. However, neither phosphatidylinositol-3 kinase inhibitors LY294002 and wortmannin nor an extracellular signal-regulated kinase inhibitor 1,4-diamino-<span class="hlt">2</span>,3-dicyano-1,4-bis(o-aminophenylmercapto) butadiene monoethanolate had effect on vascular relaxation <span class="hlt">induced</span> by the cumulative addition of <span class="hlt">G</span>1. <span class="hlt">G</span>1 <span class="hlt">induces</span> endothelium-dependent vasorelaxation through Src-mediated activation of the EGF receptor and the Akt pathway in rat aorta. © 2013 Royal Pharmaceutical Society.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3402304','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3402304"><span><span class="hlt">G</span> Protein–Coupled Receptor Kinase <span class="hlt">2</span>, With β-Arrestin <span class="hlt">2</span>, Impairs Insulin-<span class="hlt">Induced</span> Akt/Endothelial Nitric Oxide Synthase Signaling in ob/ob Mouse Aorta</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Taguchi, Kumiko; Matsumoto, Takayuki; Kamata, Katsuo; Kobayashi, Tsuneo</p> <p>2012-01-01</p> <p>In type <span class="hlt">2</span> diabetes, impaired insulin-<span class="hlt">induced</span> Akt/endothelial nitric oxide synthase (eNOS) signaling may decrease the vascular relaxation response. Previously, we reported that this response was negatively regulated by <span class="hlt">G</span> protein–coupled receptor kinase <span class="hlt">2</span> (GRK<span class="hlt">2</span>). In this study, we investigated whether/how in aortas from ob/ob mice (a model of type <span class="hlt">2</span> diabetes) GRK<span class="hlt">2</span> and β-arrestin <span class="hlt">2</span> might regulate insulin-<span class="hlt">induced</span> signaling. Endothelium-dependent relaxation was measured in aortic strips. GRK<span class="hlt">2</span>, β-arrestin <span class="hlt">2</span>, and Akt/eNOS signaling pathway proteins and activities were mainly assayed by Western blotting. In ob/ob (vs. control [Lean]) aortas: 1) insulin-<span class="hlt">induced</span> relaxation was reduced, and this deficit was prevented by GRK<span class="hlt">2</span> inhibitor, anti-GRK<span class="hlt">2</span> antibody, and an siRNA specifically targeting GRK<span class="hlt">2</span>. The Lean aorta relaxation response was reduced to the ob/ob level by pretreatment with an siRNA targeting β-arrestin <span class="hlt">2</span>. <span class="hlt">2</span>) Insulin-stimulated Akt and eNOS phosphorylations were decreased. 3) GRK<span class="hlt">2</span> expression in membranes was elevated, and, upon insulin stimulation, this expression was further increased, but β-arrestin <span class="hlt">2</span> was decreased. In ob/ob aortic membranes under insulin stimulation, the phosphorylations of Akt and eNOS were augmented by GRK<span class="hlt">2</span> inhibitor. In mouse aorta, GRK<span class="hlt">2</span> may be, upon translocation, a key negative regulator of insulin responsiveness and an important regulator of the β-arrestin <span class="hlt">2</span>/Akt/eNOS signaling, which is implicated in diabetic endothelial dysfunction. PMID:22688330</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li class="active"><span>23</span></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_23 --> <div id="page_24" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li class="active"><span>24</span></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="461"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22688330','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22688330"><span><span class="hlt">G</span> protein-coupled receptor kinase <span class="hlt">2</span>, with β-arrestin <span class="hlt">2</span>, impairs insulin-<span class="hlt">induced</span> Akt/endothelial nitric oxide synthase signaling in ob/ob mouse aorta.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Taguchi, Kumiko; Matsumoto, Takayuki; Kamata, Katsuo; Kobayashi, Tsuneo</p> <p>2012-08-01</p> <p>In type <span class="hlt">2</span> diabetes, impaired insulin-<span class="hlt">induced</span> Akt/endothelial nitric oxide synthase (eNOS) signaling may decrease the vascular relaxation response. Previously, we reported that this response was negatively regulated by <span class="hlt">G</span> protein-coupled receptor kinase <span class="hlt">2</span> (GRK<span class="hlt">2</span>). In this study, we investigated whether/how in aortas from ob/ob mice (a model of type <span class="hlt">2</span> diabetes) GRK<span class="hlt">2</span> and β-arrestin <span class="hlt">2</span> might regulate insulin-<span class="hlt">induced</span> signaling. Endothelium-dependent relaxation was measured in aortic strips. GRK<span class="hlt">2</span>, β-arrestin <span class="hlt">2</span>, and Akt/eNOS signaling pathway proteins and activities were mainly assayed by Western blotting. In ob/ob (vs. control [Lean]) aortas: 1) insulin-<span class="hlt">induced</span> relaxation was reduced, and this deficit was prevented by GRK<span class="hlt">2</span> inhibitor, anti-GRK<span class="hlt">2</span> antibody, and an siRNA specifically targeting GRK<span class="hlt">2</span>. The Lean aorta relaxation response was reduced to the ob/ob level by pretreatment with an siRNA targeting β-arrestin <span class="hlt">2</span>. <span class="hlt">2</span>) Insulin-stimulated Akt and eNOS phosphorylations were decreased. 3) GRK<span class="hlt">2</span> expression in membranes was elevated, and, upon insulin stimulation, this expression was further increased, but β-arrestin <span class="hlt">2</span> was decreased. In ob/ob aortic membranes under insulin stimulation, the phosphorylations of Akt and eNOS were augmented by GRK<span class="hlt">2</span> inhibitor. In mouse aorta, GRK<span class="hlt">2</span> may be, upon translocation, a key negative regulator of insulin responsiveness and an important regulator of the β-arrestin <span class="hlt">2</span>/Akt/eNOS signaling, which is implicated in diabetic endothelial dysfunction.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2688491','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2688491"><span>Proteome analysis of Norway maple (Acer platanoides L.) seeds dormancy <span class="hlt">breaking</span> and germination: influence of abscisic and gibberellic acids</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Pawłowski, Tomasz A</p> <p>2009-01-01</p> <p>Background Seed dormancy is controlled by the physiological or structural properties of a seed and the external conditions. It is <span class="hlt">induced</span> as part of the genetic program of seed development and maturation. Seeds with deep physiological embryo dormancy can be stimulated to germinate by a variety of treatments including cold stratification. Hormonal imbalance between germination inhibitors (e.<span class="hlt">g</span>. abscisic acid) and growth promoters (e.<span class="hlt">g</span>. gibberellins) is the main cause of seed dormancy <span class="hlt">breaking</span>. Differences in the status of hormones would affect expression of genes required for germination. Proteomics offers the opportunity to examine simultaneous changes and to classify temporal patterns of protein accumulation occurring during seed dormancy <span class="hlt">breaking</span> and germination. Analysis of the functions of the identified proteins and the related metabolic pathways, in conjunction with the plant hormones implicated in seed dormancy <span class="hlt">breaking</span>, would expand our knowledge about this process. Results A proteomic approach was used to analyse the mechanism of dormancy <span class="hlt">breaking</span> in Norway maple seeds caused by cold stratification, and the participation of the abscisic (ABA) and gibberellic (GA) acids. Forty-four proteins showing significant changes were identified by mass spectrometry. Of these, eight spots were identified as water-responsive, 18 spots were ABA- and nine GA-responsive and nine spots were regulated by both hormones. The classification of proteins showed that most of the proteins associated with dormancy <span class="hlt">breaking</span> in water were involved in protein destination. Most of the ABA- and GA-responsive proteins were involved in protein destination and energy metabolism. Conclusion In this study, ABA was found to mostly down-regulate proteins whereas GA up-regulated proteins abundance. Most of the changes were observed at the end of stratification in the germinated seeds. This is the most active period of dormancy <span class="hlt">breaking</span> when seeds pass from the quiescent state to germination. Seed</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19413897','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19413897"><span>Proteome analysis of Norway maple (Acer platanoides L.) seeds dormancy <span class="hlt">breaking</span> and germination: influence of abscisic and gibberellic acids.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pawłowski, Tomasz A</p> <p>2009-05-04</p> <p>Seed dormancy is controlled by the physiological or structural properties of a seed and the external conditions. It is <span class="hlt">induced</span> as part of the genetic program of seed development and maturation. Seeds with deep physiological embryo dormancy can be stimulated to germinate by a variety of treatments including cold stratification. Hormonal imbalance between germination inhibitors (e.<span class="hlt">g</span>. abscisic acid) and growth promoters (e.<span class="hlt">g</span>. gibberellins) is the main cause of seed dormancy <span class="hlt">breaking</span>. Differences in the status of hormones would affect expression of genes required for germination. Proteomics offers the opportunity to examine simultaneous changes and to classify temporal patterns of protein accumulation occurring during seed dormancy <span class="hlt">breaking</span> and germination. Analysis of the functions of the identified proteins and the related metabolic pathways, in conjunction with the plant hormones implicated in seed dormancy <span class="hlt">breaking</span>, would expand our knowledge about this process. A proteomic approach was used to analyse the mechanism of dormancy <span class="hlt">breaking</span> in Norway maple seeds caused by cold stratification, and the participation of the abscisic (ABA) and gibberellic (GA) acids. Forty-four proteins showing significant changes were identified by mass spectrometry. Of these, eight spots were identified as water-responsive, 18 spots were ABA- and nine GA-responsive and nine spots were regulated by both hormones. The classification of proteins showed that most of the proteins associated with dormancy <span class="hlt">breaking</span> in water were involved in protein destination. Most of the ABA- and GA-responsive proteins were involved in protein destination and energy metabolism. In this study, ABA was found to mostly down-regulate proteins whereas GA up-regulated proteins abundance. Most of the changes were observed at the end of stratification in the germinated seeds. This is the most active period of dormancy <span class="hlt">breaking</span> when seeds pass from the quiescent state to germination. Seed dormancy <span class="hlt">breaking</span> involves</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4260214','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4260214"><span>Sequence dependence of electron-<span class="hlt">induced</span> DNA strand breakage revealed by DNA nanoarrays</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Keller, Adrian; Rackwitz, Jenny; Cauët, Emilie; Liévin, Jacques; Körzdörfer, Thomas; Rotaru, Alexandru; Gothelf, Kurt V.; Besenbacher, Flemming; Bald, Ilko</p> <p>2014-01-01</p> <p>The electronic structure of DNA is determined by its nucleotide sequence, which is for instance exploited in molecular electronics. Here we demonstrate that also the DNA strand breakage <span class="hlt">induced</span> by low-energy electrons (18 eV) depends on the nucleotide sequence. To determine the absolute cross sections for electron <span class="hlt">induced</span> single strand <span class="hlt">breaks</span> in specific 13 mer oligonucleotides we used atomic force microscopy analysis of DNA origami based DNA nanoarrays. We investigated the DNA sequences 5′-TT(XYX)3TT with X = A, <span class="hlt">G</span>, C and Y = T, BrU 5-bromouracil and found absolute strand <span class="hlt">break</span> cross sections between <span class="hlt">2</span>.66 · 10−14 cm<span class="hlt">2</span> and 7.06 · 10−14 cm<span class="hlt">2</span>. The highest cross section was found for 5′-TT(ATA)3TT and 5′-TT(ABrUA)3TT, respectively. BrU is a radiosensitizer, which was discussed to be used in cancer radiation therapy. The replacement of T by BrU into the investigated DNA sequences leads to a slight increase of the absolute strand <span class="hlt">break</span> cross sections resulting in sequence-dependent enhancement factors between 1.14 and 1.66. Nevertheless, the variation of strand <span class="hlt">break</span> cross sections due to the specific nucleotide sequence is considerably higher. Thus, the present results suggest the development of targeted radiosensitizers for cancer radiation therapy. PMID:25487346</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22253133-dna-strand-breaks-crosslinks-induced-transient-anions-range-ev','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22253133-dna-strand-breaks-crosslinks-induced-transient-anions-range-ev"><span>DNA strand <span class="hlt">breaks</span> and crosslinks <span class="hlt">induced</span> by transient anions in the range <span class="hlt">2</span>-20 eV</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Luo, Xinglan; Zheng, Yi, E-mail: Yizheng@fzu.edu.cn; Sanche, Léon</p> <p>2014-04-21</p> <p>The energy dependence of the yields of single and double strand <span class="hlt">breaks</span> (SSB and DSB) and crosslinks <span class="hlt">induced</span> by electron impact on plasmid DNA films is measured in the <span class="hlt">2</span>-20 eV range. The yield functions exhibit two strong maxima, which are interpreted to result from the formation of core-excited resonances (i.e., transient anions) of the bases, and their decay into the autoionization channel, resulting in π → π{sup *} electronic transitions of the bases followed by electron transfer to the C–O σ{sup *} bond in the phosphate group. Occupancy of the σ{sup *} orbital ruptures the C–O bond of themore » backbone via dissociative electron attachment, producing a SSB. From a comparison of our results with those of other works, including theoretical calculations and electron-energy-loss spectra of the bases, the 4.6 eV peak in the SSB yield function is attributed to the resonance decay into the lowest electronically excited states of the bases; in particular, those resulting from the transitions 1{sup 3}A{sup ′} (π{sub <span class="hlt">2</span>} → π{sub 3}{sup *}) and 1{sup 3}A{sup ″} (n{sub <span class="hlt">2</span>} → π{sub 3}{sup *}) of thymine and 1{sup 3}A{sup ′} (π → π{sup *}) of cytosine. The strongest peak at 9.6 eV in the SSB yield function is also associated with electron captured by excited states of the bases, resulting mostly from a multitude of higher-energy π → π{sup *} transitions. The DSB yield function exhibits strong maxima at 6.1 and 9.6 eV. The peak at 9.6 eV is probably related to the same resonance manifold as that leading to SSB, but the other at 6.1 eV may be more restricted to decay into the electronic state 1{sup 3}A{sup ′} (π → π{sup *}) of cytosine via autoionization. The yield function of crosslinks is dominated by a broad peak extending over the 3.6-11.6 eV range with a sharper one at 17.6 eV. The different line shape of the latter function, compared to that of SSB and DSB, appears to be due to the formation of reactive radical sites in the initial</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=303618','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=303618"><span>Oxidant-<span class="hlt">induced</span> DNA damage of target cells.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Schraufstätter, I; Hyslop, P A; Jackson, J H; Cochrane, C G</p> <p>1988-01-01</p> <p>In this study we examined the leukocytic oxidant species that <span class="hlt">induce</span> oxidant damage of DNA in whole cells. H<span class="hlt">2</span>O<span class="hlt">2</span> added extracellularly in micromolar concentrations (10-100 microM) <span class="hlt">induced</span> DNA strand <span class="hlt">breaks</span> in various target cells. The sensitivity of a specific target cell was inversely correlated to its catalase content and the rate of removal of H<span class="hlt">2</span>O<span class="hlt">2</span> by the target cell. Oxidant species produced by xanthine oxidase/purine or phorbol myristate acetate-stimulated monocytes <span class="hlt">induced</span> DNA breakage of target cells in proportion to the amount of H<span class="hlt">2</span>O<span class="hlt">2</span> generated. These DNA strand <span class="hlt">breaks</span> were prevented by extracellular catalase, but not by superoxide dismutase. Cytotoxic doses of HOCl, added to target cells, did not <span class="hlt">induce</span> DNA strand breakage, and myeloperoxidase added extracellularly in the presence of an H<span class="hlt">2</span>O<span class="hlt">2</span>-generating system, prevented the formation of DNA strand <span class="hlt">breaks</span> in proportion to its H<span class="hlt">2</span>O<span class="hlt">2</span> degrading capacity. The studies also indicated that H<span class="hlt">2</span>O<span class="hlt">2</span> formed hydroxyl radical (.OH) intracellularly, which appeared to be the most likely free radical responsible for DNA damage: .OH was detected in cells exposed to H<span class="hlt">2</span>O<span class="hlt">2</span>; the DNA base, deoxyguanosine, was hydroxylated in cells exposed to H<span class="hlt">2</span>O<span class="hlt">2</span>; and intracellular iron was essential for induction of DNA strand <span class="hlt">breaks</span>. PMID:2843565</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010LNCS.6373..133W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010LNCS.6373..133W"><span>An Improved Memetic Algorithm for <span class="hlt">Break</span> Scheduling</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Widl, Magdalena; Musliu, Nysret</p> <p></p> <p>In this paper we consider solving a complex real life <span class="hlt">break</span> scheduling problem. This problem of high practical relevance arises in many working areas, e.<span class="hlt">g</span>. in air traffic control and other fields where supervision personnel is working. The objective is to assign <span class="hlt">breaks</span> to employees such that various constraints reflecting legal demands or ergonomic criteria are satisfied and staffing requirement violations are minimised.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21180486-biguanide-induced-mitochondrial-dysfunction-yields-increased-lactate-production-cytotoxicity-aerobically-poised-hepg2-cells-human-hepatocytes-vitro','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21180486-biguanide-induced-mitochondrial-dysfunction-yields-increased-lactate-production-cytotoxicity-aerobically-poised-hepg2-cells-human-hepatocytes-vitro"><span>Biguanide-<span class="hlt">induced</span> mitochondrial dysfunction yields increased lactate production and cytotoxicity of aerobically-poised Hep<span class="hlt">G</span><span class="hlt">2</span> cells and human hepatocytes in vitro</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Dykens, James A.; Jamieson, Joseph; Marroquin, Lisa</p> <p>2008-12-01</p> <p>As a class, the biguanides <span class="hlt">induce</span> lactic acidosis, a hallmark of mitochondrial impairment. To assess potential mitochondrial impairment, we evaluated the effects of metformin, buformin and phenformin on: 1) viability of Hep<span class="hlt">G</span><span class="hlt">2</span> cells grown in galactose, <span class="hlt">2</span>) respiration by isolated mitochondria, 3) metabolic poise of Hep<span class="hlt">G</span><span class="hlt">2</span> and primary human hepatocytes, 4) activities of immunocaptured respiratory complexes, and 5) mitochondrial membrane potential and redox status in primary human hepatocytes. Phenformin was the most cytotoxic of the three with buformin showing moderate toxicity, and metformin toxicity only at mM concentrations. Importantly, Hep<span class="hlt">G</span><span class="hlt">2</span> cells grown in galactose are markedly more susceptible to biguanidemore » toxicity compared to cells grown in glucose, indicating mitochondrial toxicity as a primary mode of action. The same rank order of potency was observed for isolated mitochondrial respiration where preincubation (40 min) exacerbated respiratory impairment, and was required to reveal inhibition by metformin, suggesting intramitochondrial bio-accumulation. Metabolic profiling of intact cells corroborated respiratory inhibition, but also revealed compensatory increases in lactate production from accelerated glycolysis. High (mM) concentrations of the drugs were needed to inhibit immunocaptured respiratory complexes, supporting the contention that bioaccumulation is involved. The same rank order was found when monitoring mitochondrial membrane potential, ROS production, and glutathione levels in primary human hepatocytes. In toto, these data indicate that biguanide-<span class="hlt">induced</span> lactic acidosis can be attributed to acceleration of glycolysis in response to mitochondrial impairment. Indeed, the desired clinical outcome, viz., decreased blood glucose, could be due to increased glucose uptake and glycolytic flux in response to drug-<span class="hlt">induced</span> mitochondrial dysfunction.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015ApCM...22..141T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015ApCM...22..141T"><span>Fibre <span class="hlt">Break</span> Failure Processes in Unidirectional Composites. Part <span class="hlt">2</span>: Failure and Critical Damage State <span class="hlt">Induced</span> by Sustained Tensile Loading</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Thionnet, A.; Chou, H. Y.; Bunsell, A.</p> <p>2015-04-01</p> <p>The purpose of these three papers is not to just revisit the modelling of unidirectional composites. It is to provide a robust framework based on physical processes that can be used to optimise the design and long term reliability of internally pressurised filament wound structures. The model presented in Part 1 for the case of monotonically loaded unidirectional composites is further developed to consider the effects of the viscoelastic nature of the matrix in determining the kinetics of fibre <span class="hlt">breaks</span> under slow or sustained loading. It is shown that the relaxation of the matrix around fibre <span class="hlt">breaks</span> leads to locally increasing loads on neighbouring fibres and in some cases their delayed failure. Although ultimate failure is similar to the elastic case in that clusters of fibre <span class="hlt">breaks</span> ultimately control composite failure the kinetics of their development varies significantly from the elastic case. Failure loads have been shown to reduce when loading rates are lowered.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18602349','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18602349"><span>Mouse but not human embryonic stem cells are deficient in rejoining of ionizing radiation-<span class="hlt">induced</span> DNA double-strand <span class="hlt">breaks</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bañuelos, C A; Banáth, J P; MacPhail, S H; Zhao, J; Eaves, C A; O'Connor, M D; Lansdorp, P M; Olive, P L</p> <p>2008-09-01</p> <p>Mouse embryonic stem (mES) cells will give rise to all of the cells of the adult mouse, but they failed to rejoin half of the DNA double-strand <span class="hlt">breaks</span> (dsb) produced by high doses of ionizing radiation. A deficiency in DNA-PK(cs) appears to be responsible since mES cells expressed <10% of the level of mouse embryo fibroblasts (MEFs) although Ku70/80 protein levels were higher than MEFs. However, the low level of DNA-PK(cs) found in wild-type cells appeared sufficient to allow rejoining of dsb after doses <20Gy even in <span class="hlt">G</span>1 phase cells. Inhibition of DNA-PK(cs) with wortmannin and NU7026 still sensitized mES cells to radiation confirming the importance of the residual DNA-PK(cs) at low doses. In contrast to wild-type cells, mES cells lacking H<span class="hlt">2</span>AX, a histone protein involved in the DNA damage response, were radiosensitive but they rejoined double-strand <span class="hlt">breaks</span> more rapidly. Consistent with more rapid dsb rejoining, H<span class="hlt">2</span>AX(-/-) mES cells also expressed 6 times more DNA-PK(cs) than wild-type mES cells. Similar results were obtained for ATM(-/-) mES cells. Differentiation of mES cells led to an increase in DNA-PK(cs), an increase in dsb rejoining rate, and a decrease in Ku70/80. Unlike mouse ES, human ES cells were proficient in rejoining of dsb and expressed high levels of DNA-PK(cs). These results confirm the importance of homologous recombination in the accurate repair of double-strand <span class="hlt">breaks</span> in mES cells, they help explain the chromosome abnormalities associated with deficiencies in H<span class="hlt">2</span>AX and ATM, and they add to the growing list of differences in the way rodent and human cells deal with DNA damage.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1995SPIE.2371..279V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1995SPIE.2371..279V"><span>Hep<span class="hlt">G</span><span class="hlt">2</span> human hepatocarcinomas cells sensitization by endogenous porphyrins</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vonarx-Coinsmann, Veronique; Foultier, Marie-Therese; de Brito, Leonor X.; Morlet, Laurent; Patrice, Thierry</p> <p>1995-03-01</p> <p>We assessed the ability of the human hepatocarcinoma cell line Hep<span class="hlt">G</span><span class="hlt">2</span> to synthesize PpIX in vitro from exogenous ALA and analyzed ALA-<span class="hlt">induced</span> toxicity and phototoxicity on this cell line. ALA <span class="hlt">induced</span> a slight dose-dependent dark toxicity, with 79 and 66% cell survival respectively for ALA 50 and 100 mg/ml after 3-h incubation. Whereas the same treatment followed by laser irradiation (l equals 632 nm, 25 J/sq cm) <span class="hlt">induced</span> dose-dependent phototoxicity, with 54 and 19% cell survival 24 h after PDT. Whatever the incubation time with ALA, a 3-h delay before light exposure was found optimal to reach a maximal phototoxicity. Photoproducts <span class="hlt">induced</span> by porphyrin light irradiation absorbed light in the red spectral region at longer wavelengths than did the original porphyrins. The possible enhancement of PDT effects after ALA Hep<span class="hlt">G</span><span class="hlt">2</span> cell incubation was investigated by irradiating cells successively with red light (l equals 632 nm) and light (l equals 650 nm). Total fluence was kept constant at 25 J/sq cm. Phototoxicity was lower when cells were irradiated for increased periods of l equals 650 nm light than with l equals 632 nm light alone. Any photoproducts involved had either a short life or were poorly photoreactive. Hep<span class="hlt">G</span><span class="hlt">2</span> cells, synthesizing enzymes and precursors of endogenous porphyrin synthesis, represent a good in vitro model for experiments using ALA-PpIX-PDT.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27100825','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27100825"><span>A Preliminary Investigation of Caffeinated Alcohol Use During Spring <span class="hlt">Break</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Linden-Carmichael, Ashley N; Lau-Barraco, Cathy</p> <p>2016-06-06</p> <p>Caffeinated alcoholic beverages (e.<span class="hlt">g</span>., Red Bull and vodka) are popular but associated with negative consequences. CABs may be particularly popular during Spring <span class="hlt">Break</span>, a potentially risky social event. We aimed to identify the prevalence of Spring <span class="hlt">Break</span> caffeinated alcohol use, determine how caffeinated alcohol use Spring <span class="hlt">Break</span> drinking habits differ from usual, and examine the association between Spring <span class="hlt">Break</span> caffeinated alcohol use and alcohol-related problems. Data were collected from 95 college students during March of 2013 and 2014. Students completed questionnaires of their alcohol and caffeinated alcohol use before and during Spring <span class="hlt">Break</span> and Spring <span class="hlt">Break</span> alcohol-related problems. Approximately 54% of students used caffeinated alcohol during Spring <span class="hlt">Break</span>. Spring <span class="hlt">Break</span> caffeinated alcohol use was associated with more alcohol-related problems, even after controlling for other alcohol consumed and Spring <span class="hlt">Break</span> vacation status. Caffeinated alcoholic beverages are commonly consumed during Spring <span class="hlt">Break</span> and their use uniquely predicted harms. Prevention efforts placed on caffeinated alcoholic beverage users may be helpful in reducing Spring <span class="hlt">Break</span>-related harms.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21783704','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21783704"><span>Methylglyoxal-bis(guanylhydrazone), a polyamine analogue, sensitized γ-radiation-<span class="hlt">induced</span> cell death in HL-60 leukemia cells Sensitizing effect of MGBG on γ-radiation-<span class="hlt">induced</span> cell death.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kim, Jin Sik; Lee, Jin; Chung, Hai Won; Choi, Han; Paik, Sang Gi; Kim, In Gyu</p> <p>2006-09-01</p> <p>Methylglyoxal-bis(guanylhydrazone) (MGBG), a polyamine analogue, has been known to inhibit the biosynthesis of polyamines, which are important in cell proliferation. We showed that MGBG treatment significantly affected γ-radiation-<span class="hlt">induced</span> cell cycle transition (<span class="hlt">G</span>(1)/<span class="hlt">G</span>(0)→S→<span class="hlt">G</span>(<span class="hlt">2</span>)/M) and thus γ-radiation-<span class="hlt">induced</span> cell death. As determined by micronuclei and comet assay, we showed that it sensitized the cytotoxic effect <span class="hlt">induced</span> by γ-radiation. One of the reasons is that polyamine depletion by MGBG treatment did not effectively protect against the chemical (OH) or physical damage to DNA caused by γ-radiation. Through in vitro experiment, we confirmed that DNA strand <span class="hlt">breaks</span> <span class="hlt">induced</span> by γ-radiation was prevented more effectively in the presence of polyamines (spermine and spermidine) than in the absence of polyamines. MGBG also blocks the cell cycle transition caused by γ-radiation (<span class="hlt">G</span>(<span class="hlt">2</span>) arrest), which helps protect cells by allowing time for DNA repair before entry into mitosis or apoptosis, via the down regulation of cyclin D1, which mediates the transition from <span class="hlt">G</span>(1) to S phase of cell cycle, and ataxia telangiectasia mutated, which is involved in the DNA sensing, repair and cell cycle check point. Therefore, the abrogation of <span class="hlt">G</span>(<span class="hlt">2</span>) arrest sensitizes cells to the effect of γ-radiation. As a result, γ-radiation-<span class="hlt">induced</span> cell death increased by about <span class="hlt">2</span>.5-3.0-fold in cells treated with MGBG. However, exogenous spermidine supplement partially relieved this γ-radiation-<span class="hlt">induced</span> cytotoxicity and cell death. These findings suggest a potentially therapeutic strategy for increasing the cytotoxic efficacy of γ-radiation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22287322','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22287322"><span>Sudan III dye strongly <span class="hlt">induces</span> CYP1A1 mRNA expression in Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ohno, Marumi; Ikenaka, Yoshinori; Ishizuka, Mayumi</p> <p>2012-01-01</p> <p>Sudan dyes possess a high affinity to the aryl hydrocarbon receptor (AHR) and potently <span class="hlt">induce</span> its target genes, such as cytochrome P450 (CYP) 1A1, through unknown mechanisms. We investigated a detailed event occurring in cells after binding of Sudan dye to AHR in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Treatment with 10 µM Sudan III caused rapid translocation of AHR into the nucleus and increased expression levels of human CYP1A1 mRNA by approximately 20-fold after 16 and 24 h. The transactivation was due to the activation of a region located at -1137 to +59 bp from CYP1A1, in particular, four xenobiotic responsive elements (XREs) existing in the region. AHR and the Ah receptor nuclear translocator interacted with XRE sequences in a gel shift assay using nuclear extract from Sudan III--treated Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Moreover, we suggest that constitutive androstane receptor could modify CYP1A1 transactivation by Sudan III. Copyright © 2012 Wiley Periodicals, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19465921','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19465921"><span>The FANC pathway and BLM collaborate during mitosis to prevent micro-nucleation and chromosome abnormalities.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Naim, Valeria; Rosselli, Filippo</p> <p>2009-06-01</p> <p>Loss-of-function of caretaker genes characterizes a group of cancer predisposition diseases that feature cellular hypersensitivity to DNA damage and chromosome fragility; this group includes Fanconi anaemia and Bloom syndrome. The products of the 13 FANC genes (mutated in Fanconi anaemia), which constitute the 'FANC' pathway, and BLM (the RecQ helicase mutated in Bloom syndrome) are thought to collaborate during the S phase of the cell cycle, preventing chromosome instability. Recently, BLM has been implicated in the completion of sister <span class="hlt">chromatid</span> separation during mitosis, a complex process in which precise regulation and execution is crucial to preserve genomic stability. Here we show for the first time a role for the FANC pathway in chromosome segregation during mitotic cell division. FANCD<span class="hlt">2</span>, a key component of the pathway, localizes to discrete spots on mitotic chromosomes. FANCD<span class="hlt">2</span> chromosomal localization is responsive to replicative stress and specifically targets aphidicolin (APH)-<span class="hlt">induced</span> <span class="hlt">chromatid</span> gaps and <span class="hlt">breaks</span>. Our data indicate that the FANC pathway is involved in rescuing abnormal anaphase and telophase (ana-telophase) cells, limiting aneuploidy and reducing chromosome instability in daughter cells. We further address a cooperative role for the FANC pathway and BLM in preventing micronucleation, through FANC-dependent targeting of BLM to non-centromeric abnormal structures <span class="hlt">induced</span> by replicative stress. We reveal new crosstalk between FANC and BLM proteins, extending their interaction beyond the S-phase rescue of damaged DNA to the safeguarding of chromosome stability during mitosis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22552584','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22552584"><span>TLR3 dsRNA agonist inhibits growth and invasion of Hep<span class="hlt">G</span><span class="hlt">2.2</span>.15 HCC cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chen, Li; Xu, Yu-Yin; Zhou, Jia-Ming; Wu, Yuan-Yuan; E, Qun; Zhu, Yuan-Yuan</p> <p>2012-07-01</p> <p>Toll-like receptor 3 (TLR3) is a pattern-recognizing receptor that is involved in immune signaling and plays a crucial role in survival by being able to recognize various viral components including double-stranded RNA (dsRNA). TLR3 expression and function in cancer cells are not well understood. In this study, we investigated whether TLR3 agonist dsRNA (BM-06) can inhibit proliferation and invasion, and promote apoptosis in Hep<span class="hlt">G</span><span class="hlt">2.2</span>.15 cells. Hep<span class="hlt">G</span><span class="hlt">2.2</span>.15 cells secreting hepatitis B virus (HBV) were treated with BM-06 and poly(I:C). Western blot analysis and PCR were employed to determine pharmacodynamic changes in biomarkers relevant to TLR3 signaling. Cell proliferation, invasion and apoptosis were analyzed by CCK-8 assay, transwell assay and flow cytometry. The expression of HBsAg, and HBcAg was observed by immunohistochemistry. Compared with untreated cells, pharmacological NF-κB activity of the TLR3 pathway by BM-06 (1.734-fold) or poly(I:C) (1.377-fold) was <span class="hlt">induced</span>. By western blot analysis, we found that dsRNA <span class="hlt">induced</span> TLR3-activated Hep<span class="hlt">G</span><span class="hlt">2.2</span>.15 cells which expressed NF-κB levels predominantly in the cytoplasmic fraction but fewer signals in the nucleus. BM-06 inhibited the proliferation, invasion and secretion of HBV, and <span class="hlt">induced</span> apoptosis in Hep<span class="hlt">G</span><span class="hlt">2.2</span>.15 cells. In addition, the antitumor effects of BM-06 were superior to poly(I:C). Pharmacological activation of the TLR3 pathway by BM-06 can inhibit Hep<span class="hlt">G</span><span class="hlt">2.2</span>.15 cell growth.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3102573','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3102573"><span>Lipotoxicity in Hep<span class="hlt">G</span><span class="hlt">2</span> cells triggered by free fatty acids</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Yao, Hong-Rui; Liu, Jun; Plumeri, Daniel; Cao, Yong-Bing; He, Ting; Lin, Ling; Li, Yu; Jiang, Yuan-Ying; Li, Ji; Shang, Jing</p> <p>2011-01-01</p> <p>The goal of this study was to investigate the lipid accumulation and lipotoxicity of free fatty acids (FFAs) <span class="hlt">induced</span> in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Hep<span class="hlt">G</span><span class="hlt">2</span> cells were co-incubated with various concentrations of FFAs for 24h and the intracellular lipid contents were observed by Oil Red O and Nile Red staining methods. The lipotoxicity of Hep<span class="hlt">G</span><span class="hlt">2</span> cells were then detected by Hoechest 33342/PI, Annexin V-FITC/PI double-staining and 3-(4,5-dimethylthiazol-<span class="hlt">2</span>-yl)-<span class="hlt">2</span>,5-di phenyltetrazolium bromide (MTT) experiment tests. The experiments showed a lipid accumulation and lipotoxicity by increasing FFA concentration gradients. Through cell morphological observation and quantitative analysis, FFAs have shown to increase in a dose-dependent manner compared with the control group. The data collected from hoechst 33342/PI, annexin V-FITC/PI double staining and also MTT experiments showed that cell apoptosis and necrosis significantly increased with increasing FFA concentrations. Apoptosis was not obvious in the 1 mM FFAs-treated group compared to the other two groups. In a certain concentration range, FFAs <span class="hlt">induced</span> intracellular lipid accumulation and lipotoxicity of Hep<span class="hlt">G</span><span class="hlt">2</span> cells in a dose-dependent manner. PMID:21654881</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25721794','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25721794"><span>Cyanidin-3-glucoside inhibits glutamate-<span class="hlt">induced</span> Zn<span class="hlt">2</span>+ signaling and neuronal cell death in cultured rat hippocampal neurons by inhibiting Ca<span class="hlt">2</span>+-<span class="hlt">induced</span> mitochondrial depolarization and formation of reactive oxygen species.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yang, Ji Seon; Perveen, Shazia; Ha, Tae Joung; Kim, Seong Yun; Yoon, Shin Hee</p> <p>2015-05-05</p> <p>Cyanidin-3-glucoside (C3<span class="hlt">G</span>), a member of the anthocyanin family, is a potent natural antioxidant. However, effects of C3<span class="hlt">G</span> on glutamate-<span class="hlt">induced</span> [Zn(<span class="hlt">2</span>+)]i increase and neuronal cell death remain unknown. We studied the effects of C3<span class="hlt">G</span> on glutamate-<span class="hlt">induced</span> [Zn(<span class="hlt">2</span>+)]i increase and cell death in cultured rat hippocampal neurons from embryonic day 17 maternal Sprague-Dawley rats using digital imaging methods for Zn(<span class="hlt">2</span>+), Ca(<span class="hlt">2</span>+), reactive oxygen species (ROS), mitochondrial membrane potential and a MTT assay for cell survival. Treatment with glutamate (100 µM) for 7 min <span class="hlt">induces</span> reproducible [Zn(<span class="hlt">2</span>+)]i increase at 35 min interval in cultured rat hippocampal neurons. The intracellular Zn(<span class="hlt">2</span>+)-chelator TPEN markedly blocked glutamate-<span class="hlt">induced</span> [Zn(<span class="hlt">2</span>+)]i increase, but the extracellular Zn(<span class="hlt">2</span>+) chelator CaEDTA did not affect glutamate-<span class="hlt">induced</span> [Zn(<span class="hlt">2</span>+)]i increase. C3<span class="hlt">G</span> inhibited the glutamate-<span class="hlt">induced</span> [Zn(<span class="hlt">2</span>+)]i response in a concentration-dependent manner (IC50 of 14.1 ± 1.1 µg/ml). C3<span class="hlt">G</span> also significantly inhibited glutamate-<span class="hlt">induced</span> [Ca(<span class="hlt">2</span>+)]i increase. Two antioxidants such as Trolox and DTT significantly inhibited the glutamate-<span class="hlt">induced</span> [Zn(<span class="hlt">2</span>+)]i response, but they did not affect the [Ca(<span class="hlt">2</span>+)]i responses. C3<span class="hlt">G</span> blocked glutamate-<span class="hlt">induced</span> formation of ROS. Trolox and DTT also inhibited the formation of ROS. C3<span class="hlt">G</span> significantly inhibited glutamate-<span class="hlt">induced</span> mitochondrial depolarization. However, TPEN, Trolox and DTT did not affect the mitochondrial depolarization. C3<span class="hlt">G</span>, Trolox and DTT attenuated glutamate-<span class="hlt">induced</span> neuronal cell death in cultured rat hippocampal neurons, respectively. Taken together, all these results suggest that cyanidin-3-glucoside inhibits glutamate-<span class="hlt">induced</span> [Zn(<span class="hlt">2</span>+)]i increase through a release of Zn(<span class="hlt">2</span>+) from intracellular sources in cultured rat hippocampal neurons by inhibiting Ca(<span class="hlt">2</span>+)-<span class="hlt">induced</span> mitochondrial depolarization and formation of ROS, which is involved in neuroprotection against glutamate-<span class="hlt">induced</span> cell death. Copyright © 2015 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22465685-jaridonin-induced-g2-phase-arrest-human-esophageal-cancer-cells-caused-reactive-oxygen-species-dependent-cdc2-tyr15-phosphorylation-via-atmchk1-pathway','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22465685-jaridonin-induced-g2-phase-arrest-human-esophageal-cancer-cells-caused-reactive-oxygen-species-dependent-cdc2-tyr15-phosphorylation-via-atmchk1-pathway"><span>Jaridonin-<span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span>/M phase arrest in human esophageal cancer cells is caused by reactive oxygen species-dependent Cdc<span class="hlt">2</span>-tyr15 phosphorylation via ATM–Chk1/<span class="hlt">2</span>–Cdc25C pathway</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Ma, Yong-Cheng; Su, Nan; Shi, Xiao-Jing</p> <p>2015-01-15</p> <p>Jaridonin, a novel diterpenoid from Isodon rubescens, has been shown previously to inhibit proliferation of esophageal squamous cancer cells (ESCC) through <span class="hlt">G</span><span class="hlt">2</span>/M phase cell cycle arrest. However, the involved mechanism is not fully understood. In this study, we found that the cell cycle arrest by Jaridonin was associated with the increased expression of phosphorylation of ATM at Ser1981 and Cdc<span class="hlt">2</span> at Tyr15. Jaridonin also resulted in enhanced phosphorylation of Cdc25C via the activation of checkpoint kinases Chk1 and Chk<span class="hlt">2</span>, as well as in increased phospho-H<span class="hlt">2</span>A.X (Ser139), which is known to be phosphorylated by ATM in response to DNA damage. Furthermore,more » Jaridonin-mediated alterations in cell cycle arrest were significantly attenuated in the presence of NAC, implicating the involvement of ROS in Jaridonin's effects. On the other hand, addition of ATM inhibitors reversed Jaridonin-related activation of ATM and Chk1/<span class="hlt">2</span> as well as phosphorylation of Cdc25C, Cdc<span class="hlt">2</span> and H<span class="hlt">2</span>A.X and <span class="hlt">G</span><span class="hlt">2</span>/M phase arrest. In conclusion, these findings identified that Jaridonin-<span class="hlt">induced</span> cell cycle arrest in human esophageal cancer cells is associated with ROS-mediated activation of ATM–Chk1/<span class="hlt">2</span>–Cdc25C pathway. - Highlights: • Jaridonin <span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span>/M phase arrest through induction of redox imbalance. • Jaridonin increased the level of ROS through depleting glutathione in cell. • ATM–Chk1/<span class="hlt">2</span>–Cdc25C were involved in Jaridonin-<span class="hlt">induced</span> cell cycle arrest. • Jaridonin selectively inhibited cancer cell viability and cell cycle progression.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=145703','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=145703"><span>Acid-<span class="hlt">induced</span> exchange of the imino proton in <span class="hlt">G</span>.C pairs.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Nonin, S; Leroy, J L; Gueron, M</p> <p>1996-01-01</p> <p>Acid-<span class="hlt">induced</span> catalysis of imino proton exchange in <span class="hlt">G</span>.C pairs of DNA duplexes is surprisingly fast, being nearly as fast as for the isolated nucleoside, despite base-pair dissociation constants in the range of 10(-5) at neutral or basic pH. It is also observed in terminal <span class="hlt">G</span>.C pairs of duplexes and in base pairs of drug-DNA complexes. We have measured imino proton exchange in deoxyguanosine and in the duplex (ATATAGATCTATAT) as a function of pH. We show that acid-<span class="hlt">induced</span> exchange can be assigned to proton transfer from N7-protonated guanosine to cytidine in the open state of the pair. This is faster than transfer from neutral guanosine (the process of intrinsic catalysis previously characterized at neutral ph) due to the lower imino proton pK of the protonated form, 7.<span class="hlt">2</span> instead of 9.4. Other interpretations are excluded by a study of exchange catalysis by formiate and cytidine as exchange catalysts. The cross-over pH between the regimes of pH-independent and acid-<span class="hlt">induced</span> exchange rates is more basic in the case of base pairs than in the mononucleoside, suggestive of an increase by one to two decades in the dissociation constant of the base pair upon N7 protonation of <span class="hlt">G</span>. Acid-<span class="hlt">induced</span> catalysis is much weaker in A.T base pairs, as expected in view of the low pK for protonation of thymidine. PMID:8604298</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li class="active"><span>24</span></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_24 --> <div id="page_25" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li class="active"><span>25</span></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="481"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/8604298','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/8604298"><span>Acid-<span class="hlt">induced</span> exchange of the imino proton in <span class="hlt">G</span>.C pairs.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nonin, S; Leroy, J L; Gueron, M</p> <p>1996-02-15</p> <p>Acid-<span class="hlt">induced</span> catalysis of imino proton exchange in <span class="hlt">G</span>.C pairs of DNA duplexes is surprisingly fast, being nearly as fast as for the isolated nucleoside, despite base-pair dissociation constants in the range of 10(-5) at neutral or basic pH. It is also observed in terminal <span class="hlt">G</span>.C pairs of duplexes and in base pairs of drug-DNA complexes. We have measured imino proton exchange in deoxyguanosine and in the duplex (ATATAGATCTATAT) as a function of pH. We show that acid-<span class="hlt">induced</span> exchange can be assigned to proton transfer from N7-protonated guanosine to cytidine in the open state of the pair. This is faster than transfer from neutral guanosine (the process of intrinsic catalysis previously characterized at neutral ph) due to the lower imino proton pK of the protonated form, 7.<span class="hlt">2</span> instead of 9.4. Other interpretations are excluded by a study of exchange catalysis by formiate and cytidine as exchange catalysts. The cross-over pH between the regimes of pH-independent and acid-<span class="hlt">induced</span> exchange rates is more basic in the case of base pairs than in the mononucleoside, suggestive of an increase by one to two decades in the dissociation constant of the base pair upon N7 protonation of <span class="hlt">G</span>. Acid-<span class="hlt">induced</span> catalysis is much weaker in A.T base pairs, as expected in view of the low pK for protonation of thymidine.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27246850','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27246850"><span>The Clustered, Regularly Interspaced, Short Palindromic Repeats-associated Endonuclease 9 (CRISPR/Cas9)-created MDM<span class="hlt">2</span> T309<span class="hlt">G</span> Mutation Enhances Vitreous-<span class="hlt">induced</span> Expression of MDM<span class="hlt">2</span> and Proliferation and Survival of Cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Duan, Yajian; Ma, Gaoen; Huang, Xionggao; D'Amore, Patricia A; Zhang, Feng; Lei, Hetian</p> <p>2016-07-29</p> <p>The <span class="hlt">G</span>309 allele of SNPs in the mouse double minute (MDM<span class="hlt">2</span>) promoter locus is associated with a higher risk of cancer and proliferative vitreoretinopathy (PVR), but whether SNP <span class="hlt">G</span>309 contributes to the pathogenesis of PVR is to date unknown. The clustered regularly interspaced short palindromic repeats (CRISPR)-associated endonuclease (Cas) 9 from Streptococcus pyogenes (SpCas9) can be harnessed to manipulate a single or multiple nucleotides in mammalian cells. Here we delivered SpCas9 and guide RNAs using dual adeno-associated virus-derived vectors to target the MDM<span class="hlt">2</span> genomic locus together with a homologous repair template for creating the mutation of MDM<span class="hlt">2</span> T309<span class="hlt">G</span> in human primary retinal pigment epithelial (hPRPE) cells whose genotype is MDM<span class="hlt">2</span> T309T. The next-generation sequencing results indicated that there was 42.51% MDM<span class="hlt">2</span> <span class="hlt">G</span>309 in the edited hPRPE cells using adeno-associated viral CRISPR/Cas9. Our data showed that vitreous <span class="hlt">induced</span> an increase in MDM<span class="hlt">2</span> and subsequent attenuation of p53 expression in MDM<span class="hlt">2</span> T309<span class="hlt">G</span> hPRPE cells. Furthermore, our experimental results demonstrated that MDM<span class="hlt">2</span> T309<span class="hlt">G</span> in hPRPE cells enhanced vitreous-<span class="hlt">induced</span> cell proliferation and survival, suggesting that this SNP contributes to the pathogenesis of PVR. © 2016 by The American Society for Biochemistry and Molecular Biology, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27701048','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27701048"><span>Cobalt iron oxide nanoparticles <span class="hlt">induce</span> cytotoxicity and regulate the apoptotic genes through ROS in human liver cells (Hep<span class="hlt">G</span><span class="hlt">2</span>).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ahamed, Maqusood; Akhtar, Mohd Javed; Khan, M A Majeed; Alhadlaq, Hisham A; Alshamsan, Aws</p> <p>2016-12-01</p> <p>Cobalt iron oxide (CoFe <span class="hlt">2</span> O 4 ) nanoparticles (CIO NPs) have been one of the most widely explored magnetic NPs because of their excellent chemical stability, mechanical hardness and heat generating potential. However, there is limited information concerning the interaction of CIO NPs with biological systems. In this study, we investigated the reactive oxygen species (ROS) mediated cytotoxicity and apoptotic response of CIO NPs in human liver cells (Hep<span class="hlt">G</span><span class="hlt">2</span>). Diameter of crystalline CIO NPs was found to be 23nm with a band gap of 1.97eV. CIO NPs <span class="hlt">induced</span> cell viability reduction and membrane damage, and degree of induction was dose- and time-dependent. CIO NPs were also found to <span class="hlt">induce</span> oxidative stress revealed by induction of ROS, depletion of glutathione and lower activity of superoxide dismutase enzyme. Real-time PCR data has shown that mRNA level of tumor suppressor gene p53 and apoptotic genes (bax, CASP3 and CASP9) were higher, while the expression level of anti-apoptotic gene bcl-<span class="hlt">2</span> was lower in cells following exposure to CIO NPs. Activity of caspase-3 and caspase-9 enzymes was also higher in CIO NPs exposed cells. Furthermore, co-exposure of N-acetyl-cysteine (ROS scavenger) efficiently abrogated the modulation of apoptotic genes along with the prevention of cytotoxicity caused by CIO NPs. Overall, we observed that CIO NPs <span class="hlt">induced</span> cytotoxicity and apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells through ROS via p53 pathway. This study suggests that toxicity mechanisms of CIO NPs should be further investigated in animal models. Copyright © 2016 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29730480','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29730480"><span>Optimised detection of mitochondrial DNA strand <span class="hlt">breaks</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hanna, Rebecca; Crowther, Jonathan M; Bulsara, Pallav A; Wang, Xuying; Moore, David J; Birch-Machin, Mark A</p> <p>2018-05-04</p> <p>Intrinsic and extrinsic factors that <span class="hlt">induce</span> cellular oxidative stress damage tissue integrity and promote ageing, resulting in accumulative strand <span class="hlt">breaks</span> to the mitochondrial DNA (mtDNA) genome. Limited repair mechanisms and close proximity to superoxide generation make mtDNA a prominent biomarker of oxidative damage. Using human DNA we describe an optimised long-range qPCR methodology that sensitively detects mtDNA strand <span class="hlt">breaks</span> relative to a suite of short mitochondrial and nuclear DNA housekeeping amplicons, which control for any variation in mtDNA copy number. An application is demonstrated by detecting 16-36-fold mtDNA damage in human skin cells <span class="hlt">induced</span> by hydrogen peroxide and solar simulated radiation. Copyright © 2018 Elsevier B.V. and Mitochondria Research Society. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17274761','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17274761"><span>Alteration of mitochondrial membrane potential by Spirulina platensis C-phycocyanin <span class="hlt">induces</span> apoptosis in the doxorubicinresistant human hepatocellular-carcinoma cell line Hep<span class="hlt">G</span><span class="hlt">2</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Roy, Karnati R; Arunasree, Kalle M; Reddy, Nishant P; Dheeraj, Bhavanasi; Reddy, Gorla Venkateswara; Reddanna, Pallu</p> <p>2007-07-01</p> <p>C-PC (C-phycocyanin) is a water-soluble biliprotein from the filamentous cyanobacterium Spirulina platensis with potent antioxidant, anti-inflammatory and anticancerous properties. In the present study, the effect of C-PC was tested on the proliferation of doxorubicin-sensitive (S-Hep<span class="hlt">G</span><span class="hlt">2</span>) and -resistant (R-Hep<span class="hlt">G</span><span class="hlt">2</span>) HCC (hepatocellular carcinoma) cell lines. These studies indicate a 50% decrease in the proliferation of S- and R-Hep<span class="hlt">G</span><span class="hlt">2</span> cells treated with 40 and 50 microM C-PC for 24 h respectively. C-PC also enhanced the sensitivity of R-Hep<span class="hlt">G</span><span class="hlt">2</span> cells to doxorubicin. R-Hep<span class="hlt">G</span><span class="hlt">2</span> cells treated with C-PC showed typical apoptotic features such as membrane blebbing and DNA fragmentation. Flow-cytometric analysis of R-Hep<span class="hlt">G</span><span class="hlt">2</span> cells treated with 10, 25 and 50 microM C-PC for 24 h showed 18.8, 39.72 and 65.64% cells in sub-<span class="hlt">G</span>(0)/<span class="hlt">G</span>(1)-phase respectively. Cytochrome c release, decrease in membrane potential, caspase 3 activation and PARP [poly(ADP-ribose) polymerase] cleavage were observed in C-PC-treated R-Hep<span class="hlt">G</span><span class="hlt">2</span> cells. These studies also showed down-regulation of the anti-apoptotic protein Bcl-<span class="hlt">2</span> and up-regulation of the pro-apoptotic Bax (Bcl<span class="hlt">2</span>-associated X-protein) protein in the R-Hep<span class="hlt">G</span><span class="hlt">2</span> cells treated with C-PC. The present study thus demonstrates that C-PC <span class="hlt">induces</span> apoptosis in R-Hep<span class="hlt">G</span><span class="hlt">2</span> cells and its potential as an anti-HCC agent.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006AGUFMOS41C0628H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006AGUFMOS41C0628H"><span>Remote Sensing Characteristics of Wave <span class="hlt">Breaking</span> Rollers</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Haller, M. C.; Catalan, P.</p> <p>2006-12-01</p> <p>The wave roller has a primary influence on the balances of mass and momentum in the surf zone (e.<span class="hlt">g</span>. Svendsen, 1984; Dally and Brown, 1995; Ruessink et al., 2001). In addition, the roller area and its angle of inclination on the wave front are important quantities governing the dissipation rates in <span class="hlt">breaking</span> waves (e.<span class="hlt">g</span> Madsen et al., 1997). Yet, there have been very few measurements published of individual <span class="hlt">breaking</span> wave roller geometries in shallow water. A number of investigators have focused on observations of the initial jet-like motion at the onset of <span class="hlt">breaking</span> before the establishment of the wave roller (e.<span class="hlt">g</span>. Basco, 1985; Jansen, 1986), while Govender et al. (2002) provide observations of wave roller vertical cross-sections and angles of inclination for a pair of laboratory wave conditions. Nonetheless, presently very little is known about the growth, evolution, and decay of this aerated region of white water as it propagates through the surf zone; mostly due to the inherent difficulties in making the relevant observations. The present work is focused on analyzing observations of the time and space scales of individual shallow water <span class="hlt">breaking</span> wave rollers as derived from remote sensing systems. Using a high-resolution video system in a large-scale laboratory facility, we have obtained detailed measurements of the growth and evolution of the wave <span class="hlt">breaking</span> roller. In addition, by synchronizing the remote video with in-situ wave gages, we are able to directly relate the video intensity signal to the underlying wave shape. Results indicate that the horizontal length scale of <span class="hlt">breaking</span> wave rollers differs significantly from the previous observations of Duncan (1981), which has been a traditional basis for roller model parameterizations. The overall approach to the video analysis is new in the sense that we concentrate on individual <span class="hlt">breaking</span> waves, as opposed to the more commonly used time-exposure technique. In addition, a new parameter of interest, denoted Imax, is</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/6488477','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/6488477"><span>Increased incidence of chromosomal aberrations in peripheral lymphocytes of retired nickel workers.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Waksvik, H; Boysen, M; Høgetveit, A C</p> <p>1984-11-01</p> <p>Chromosomal aberrations and sister <span class="hlt">chromatid</span> exchanges were analysed in the peripheral lymphocytes of nine retired nickel refinery workers 4-15 years after the retirement and compared with 11 matched non-nickel exposed controls. None of the controls had previous occupations with known relation to induction of chromosomal aberrations nor sister <span class="hlt">chromatid</span> exchanges. The groups were equal as to socioeconomic status and environmental factors other than the occupational ones, which could influence the chromosome parameters, were to the largest possible extent excluded. The nickel workers' previous occupational employment involved exposure to inhalation of furnace dust of Ni3S<span class="hlt">2</span> and NiO or aerosols of NiCl<span class="hlt">2</span> and NiSO4. The concentration of nickel in the working atmospheres has been higher than 1.0 mg/m3 air and the exposure time more than 25 years. The retired nickel workers showed an increased incidence of <span class="hlt">breaks</span> (p less than 0.001) and gaps (p less than 0.05) but no difference in the incidence of sister <span class="hlt">chromatid</span> exchanges when compared with the controls.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18513048','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18513048"><span>Polymer-<span class="hlt">induced</span> phase separation and crystallization in immunoglobulin <span class="hlt">G</span> solutions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Li, Jianguo; Rajagopalan, Raj; Jiang, Jianwen</p> <p>2008-05-28</p> <p>We study the effects of the size of polymer additives and ionic strength on the phase behavior of a nonglobular protein-immunoglobulin <span class="hlt">G</span> (Ig<span class="hlt">G</span>)-by using a simple four-site model to mimic the shape of Ig<span class="hlt">G</span>. The interaction potential between the protein molecules consists of a Derjaguin-Landau-Verwey-Overbeek-type colloidal potential and an Asakura-Oosawa depletion potential arising from the addition of polymer. Liquid-liquid equilibria and fluid-solid equilibria are calculated by using the Gibbs ensemble Monte Carlo technique and the Gibbs-Duhem integration (GDI) method, respectively. Absolute Helmholtz energy is also calculated to get an initial coexisting point as required by GDI. The results reveal a nonmonotonic dependence of the critical polymer concentration rho(PEG) (*) (i.e., the minimum polymer concentration needed to <span class="hlt">induce</span> liquid-liquid phase separation) on the polymer-to-protein size ratio q (equivalently, the range of the polymer-<span class="hlt">induced</span> depletion interaction potential). We have developed a simple equation for estimating the minimum amount of polymer needed to <span class="hlt">induce</span> the liquid-liquid phase separation and show that rho(PEG) (*) approximately [q(1+q)(3)]. The results also show that the liquid-liquid phase separation is metastable for low-molecular weight polymers (q=0.<span class="hlt">2</span>) but stable at large molecular weights (q=1.0), thereby indicating that small sizes of polymer are required for protein crystallization. The simulation results provide practical guidelines for the selection of polymer size and ionic strength for protein phase separation and crystallization.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29434714','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29434714"><span>20(S)-Protopanaxadiol <span class="hlt">induces</span> apoptosis in human hepatoblastoma Hep<span class="hlt">G</span><span class="hlt">2</span> cells by downregulating the protein kinase B signaling pathway.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lu, Zeyuan; Xu, Huali; Yu, Xiaofeng; Wang, Yuchen; Huang, Long; Jin, Xin; Sui, Dayun</p> <p>2018-02-01</p> <p>Hepatoblastoma is the most common primary liver tumor for children aged <5 years old. 20(S)-Protopanaxadiol (PPD) is a ginsenoside extracted from Pananx quinquefolium L ., which inhibits tumor growth in several cancer cell lines. The purpose of the present study was to assess the anticancer activities of 20(S)-PPD in human hepatoblastoma Hep<span class="hlt">G</span><span class="hlt">2</span> cells. The cytotoxicity of 20(S)-PPD on Hep<span class="hlt">G</span><span class="hlt">2</span> cells was evaluated using an MTT assay. Apoptosis was detected using DAPI staining and flow cytometry. The expression of apoptosis-associated proteins was identified by western blotting. The results demonstrated that 20(S)-PPD inhibited the viability of Hep<span class="hlt">G</span><span class="hlt">2</span> cell in a dose and time-dependent manner. The IC 50 values were 81.35, 73.5, 48.79 µM at 24, 48 and 72 h, respectively. Topical morphological changes of apoptotic body formation following 20(S)-PPD treatment were detected by DAPI staining. The percentage of Annexin V-fluoroscein isothyiocyanate positive cells were 3.73, 17.61, 23.44 and 65.43% in Hep<span class="hlt">G</span><span class="hlt">2</span> cells treated with 0, 40, 50 and 60 µM of 20(S)-PPD, respectively. Furthermore, 20(S)-PPD upregulated the expression of Bax and downregulated the expression of Bcl-<span class="hlt">2</span> and also activated caspases-3 and -9, and Poly [ADP-ribose] polymerase cleavage. In addition, 20(S)-PPD inhibited the phosphorylation of protein kinase B (Akt; Ser473). The results indicate that 20(S)-PPD inhibits the viability of Hep<span class="hlt">G</span><span class="hlt">2</span> cells and <span class="hlt">induces</span> apoptosis in Hep<span class="hlt">G</span><span class="hlt">2</span> cells by inhibiting the phosphoinositide-3-kinase/Akt pathway.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4370595','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4370595"><span>Mitotic UV Irradiation <span class="hlt">Induces</span> a DNA Replication-Licensing Defect that Potentiates <span class="hlt">G</span>1 Arrest Response</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Morino, Masayuki; Nukina, Kohei; Sakaguchi, Hiroki; Maeda, Takeshi; Takahara, Michiyo; Shiomi, Yasushi; Nishitani, Hideo</p> <p>2015-01-01</p> <p>Cdt1 begins to accumulate in M phase and has a key role in establishing replication licensing at the end of mitosis or in early <span class="hlt">G</span>1 phase. Treatments that damage the DNA of cells, such as UV irradiation, <span class="hlt">induce</span> Cdt1 degradation through PCNA-dependent CRL4-Cdt<span class="hlt">2</span> ubiquitin ligase. How Cdt1 degradation is linked to cell cycle progression, however, remains unclear. In <span class="hlt">G</span>1 phase, when licensing is established, UV irradiation leads to Cdt1 degradation, but has little effect on the licensing state. In M phase, however, UV irradiation does not <span class="hlt">induce</span> Cdt1 degradation. When mitotic UV-irradiated cells were released into <span class="hlt">G</span>1 phase, Cdt1 was degraded before licensing was established. Thus, these cells exhibited both defective licensing and <span class="hlt">G</span>1 cell cycle arrest. The frequency of <span class="hlt">G</span>1 arrest increased in cells expressing extra copies of Cdt<span class="hlt">2</span>, and thus in cells in which Cdt1 degradation was enhanced, whereas the frequency of <span class="hlt">G</span>1 arrest was reduced in cell expressing an extra copy of Cdt1. The <span class="hlt">G</span>1 arrest response of cells irradiated in mitosis was important for cell survival by preventing the induction of apoptosis. Based on these observations, we propose that mammalian cells have a DNA replication-licensing checkpoint response to DNA damage <span class="hlt">induced</span> during mitosis. PMID:25798850</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27180624','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27180624"><span>A novel podocyte gene, semaphorin 3<span class="hlt">G</span>, protects glomerular podocyte from lipopolysaccharide-<span class="hlt">induced</span> inflammation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ishibashi, Ryoichi; Takemoto, Minoru; Akimoto, Yoshihiro; Ishikawa, Takahiro; He, Peng; Maezawa, Yoshiro; Sakamoto, Kenichi; Tsurutani, Yuya; Ide, Shintaro; Ide, Kana; Kawamura, Harukiyo; Kobayashi, Kazuki; Tokuyama, Hirotake; Tryggvason, Karl; Betsholtz, Christer; Yokote, Koutaro</p> <p>2016-05-16</p> <p>Kidney diseases including diabetic nephropathy have become huge medical problems, although its precise mechanisms are still far from understood. In order to increase our knowledge about the patho-physiology of kidney, we have previously identified >300 kidney glomerulus-enriched transcripts through large-scale sequencing and microarray profiling of the mouse glomerular transcriptome. One of the glomerulus-specific transcripts identified was semaphorin 3<span class="hlt">G</span> (Sema3<span class="hlt">G</span>) which belongs to the semaphorin family. The aim of this study was to analyze both the in vivo and in vitro functions of Sema3<span class="hlt">G</span> in the kidney. Sema3<span class="hlt">G</span> was expressed in glomerular podocytes. Although Sema3<span class="hlt">G</span> knockout mice did not show obvious glomerular defects, ultrastructural analyses revealed partially aberrant podocyte foot processes structures. When these mice were injected with lipopolysaccharide to <span class="hlt">induce</span> acute inflammation or streptozotocin to <span class="hlt">induce</span> diabetes, the lack of Sema3<span class="hlt">G</span> resulted in increased albuminuria. The lack of Sema3<span class="hlt">G</span> in podocytes also enhanced the expression of inflammatory cytokines including chemokine ligand <span class="hlt">2</span> and interleukin 6. On the other hand, the presence of Sema3<span class="hlt">G</span> attenuated their expression through the inhibition of lipopolysaccharide-<span class="hlt">induced</span> Toll like receptor 4 signaling. Taken together, our results surmise that the Sema3<span class="hlt">G</span> protein is secreted by podocytes and protects podocytes from inflammatory kidney diseases and diabetic nephropathy.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4867620','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4867620"><span>A novel podocyte gene, semaphorin 3<span class="hlt">G</span>, protects glomerular podocyte from lipopolysaccharide-<span class="hlt">induced</span> inflammation</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Ishibashi, Ryoichi; Takemoto, Minoru; Akimoto, Yoshihiro; Ishikawa, Takahiro; He, Peng; Maezawa, Yoshiro; Sakamoto, Kenichi; Tsurutani, Yuya; Ide, Shintaro; Ide, Kana; Kawamura, Harukiyo; Kobayashi, Kazuki; Tokuyama, Hirotake; Tryggvason, Karl; Betsholtz, Christer; Yokote, Koutaro</p> <p>2016-01-01</p> <p>Kidney diseases including diabetic nephropathy have become huge medical problems, although its precise mechanisms are still far from understood. In order to increase our knowledge about the patho-physiology of kidney, we have previously identified >300 kidney glomerulus-enriched transcripts through large-scale sequencing and microarray profiling of the mouse glomerular transcriptome. One of the glomerulus-specific transcripts identified was semaphorin 3<span class="hlt">G</span> (Sema3<span class="hlt">G</span>) which belongs to the semaphorin family. The aim of this study was to analyze both the in vivo and in vitro functions of Sema3<span class="hlt">G</span> in the kidney. Sema3<span class="hlt">G</span> was expressed in glomerular podocytes. Although Sema3<span class="hlt">G</span> knockout mice did not show obvious glomerular defects, ultrastructural analyses revealed partially aberrant podocyte foot processes structures. When these mice were injected with lipopolysaccharide to <span class="hlt">induce</span> acute inflammation or streptozotocin to <span class="hlt">induce</span> diabetes, the lack of Sema3<span class="hlt">G</span> resulted in increased albuminuria. The lack of Sema3<span class="hlt">G</span> in podocytes also enhanced the expression of inflammatory cytokines including chemokine ligand <span class="hlt">2</span> and interleukin 6. On the other hand, the presence of Sema3<span class="hlt">G</span> attenuated their expression through the inhibition of lipopolysaccharide-<span class="hlt">induced</span> Toll like receptor 4 signaling. Taken together, our results surmise that the Sema3<span class="hlt">G</span> protein is secreted by podocytes and protects podocytes from inflammatory kidney diseases and diabetic nephropathy. PMID:27180624</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5542668','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5542668"><span>Persistent damaged bases in DNA allow mutagenic <span class="hlt">break</span> repair in Escherichia coli</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Moore, Jessica M.; Correa, Raul; Rosenberg, Susan M.</p> <p>2017-01-01</p> <p>Bacteria, yeast and human cancer cells possess mechanisms of mutagenesis upregulated by stress responses. Stress-<span class="hlt">inducible</span> mutagenesis potentially accelerates adaptation, and may provide important models for mutagenesis that drives cancers, host pathogen interactions, antibiotic resistance and possibly much of evolution generally. In Escherichia coli repair of double-strand <span class="hlt">breaks</span> (DSBs) becomes mutagenic, using low-fidelity DNA polymerases under the control of the SOS DNA-damage response and RpoS general stress response, which upregulate and allow the action of error-prone DNA polymerases IV (DinB), II and V to make mutations during repair. Pol IV is implied to compete with and replace high-fidelity DNA polymerases at the DSB-repair replisome, causing mutagenesis. We report that up-regulated Pol IV is not sufficient for mutagenic <span class="hlt">break</span> repair (MBR); damaged bases in the DNA are also required, and that in starvation-stressed cells, these are caused by reactive-oxygen species (ROS). First, MBR is reduced by either ROS-scavenging agents or constitutive activation of oxidative-damage responses, both of which reduce cellular ROS levels. The ROS promote MBR other than by causing DSBs, saturating mismatch repair, oxidizing proteins, or <span class="hlt">inducing</span> the SOS response or the general stress response. We find that ROS drive MBR through oxidized guanines (8-oxo-d<span class="hlt">G</span>) in DNA, in that overproduction of a glycosylase that removes 8-oxo-d<span class="hlt">G</span> from DNA prevents MBR. Further, other damaged DNA bases can substitute for 8-oxo-d<span class="hlt">G</span> because ROS-scavenged cells resume MBR if either DNA pyrimidine dimers or alkylated bases are <span class="hlt">induced</span>. We hypothesize that damaged bases in DNA pause the replisome and allow the critical switch from high fidelity to error-prone DNA polymerases in the DSB-repair replisome, thus allowing MBR. The data imply that in addition to the indirect stress-response controlled switch to MBR, a direct cis-acting switch to MBR occurs independently of DNA breakage, caused by ROS</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28727736','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28727736"><span>Persistent damaged bases in DNA allow mutagenic <span class="hlt">break</span> repair in Escherichia coli.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Moore, Jessica M; Correa, Raul; Rosenberg, Susan M; Hastings, P J</p> <p>2017-07-01</p> <p>Bacteria, yeast and human cancer cells possess mechanisms of mutagenesis upregulated by stress responses. Stress-<span class="hlt">inducible</span> mutagenesis potentially accelerates adaptation, and may provide important models for mutagenesis that drives cancers, host pathogen interactions, antibiotic resistance and possibly much of evolution generally. In Escherichia coli repair of double-strand <span class="hlt">breaks</span> (DSBs) becomes mutagenic, using low-fidelity DNA polymerases under the control of the SOS DNA-damage response and RpoS general stress response, which upregulate and allow the action of error-prone DNA polymerases IV (DinB), II and V to make mutations during repair. Pol IV is implied to compete with and replace high-fidelity DNA polymerases at the DSB-repair replisome, causing mutagenesis. We report that up-regulated Pol IV is not sufficient for mutagenic <span class="hlt">break</span> repair (MBR); damaged bases in the DNA are also required, and that in starvation-stressed cells, these are caused by reactive-oxygen species (ROS). First, MBR is reduced by either ROS-scavenging agents or constitutive activation of oxidative-damage responses, both of which reduce cellular ROS levels. The ROS promote MBR other than by causing DSBs, saturating mismatch repair, oxidizing proteins, or <span class="hlt">inducing</span> the SOS response or the general stress response. We find that ROS drive MBR through oxidized guanines (8-oxo-d<span class="hlt">G</span>) in DNA, in that overproduction of a glycosylase that removes 8-oxo-d<span class="hlt">G</span> from DNA prevents MBR. Further, other damaged DNA bases can substitute for 8-oxo-d<span class="hlt">G</span> because ROS-scavenged cells resume MBR if either DNA pyrimidine dimers or alkylated bases are <span class="hlt">induced</span>. We hypothesize that damaged bases in DNA pause the replisome and allow the critical switch from high fidelity to error-prone DNA polymerases in the DSB-repair replisome, thus allowing MBR. The data imply that in addition to the indirect stress-response controlled switch to MBR, a direct cis-acting switch to MBR occurs independently of DNA breakage, caused by ROS</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26219504','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26219504"><span>Hexachlorobenzene <span class="hlt">induces</span> cell proliferation, and aryl hydrocarbon receptor expression (AhR) in rat liver preneoplastic foci, and in the human hepatoma cell line Hep<span class="hlt">G</span><span class="hlt">2</span>. AhR is a mediator of ERK1/<span class="hlt">2</span> signaling, and cell cycle regulation in HCB-treated Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>de Tomaso Portaz, Ana Clara; Caimi, Giselle Romero; Sánchez, Marcela; Chiappini, Florencia; Randi, Andrea S; Kleiman de Pisarev, Diana L; Alvarez, Laura</p> <p>2015-10-02</p> <p>Hexachlorobenzene (HCB) is a widespread environmental pollutant, and a liver tumor promoter in rodents. Depending on the particular cell lines studied, exposure to these compounds may lead to cell proliferation, terminal differentiation, or apoptosis. The aryl hydrocarbon receptor (AhR) is a ligand-activated transcription factor that is involved in drug and xenobiotic metabolism. AhR can also modulate a variety of cellular and physiological processes that can affect cell proliferation and cell fate determination. The mechanisms by which AhR ligands, both exogenous and endogenous, affect these processes involve multiple interactions between AhR and other signaling pathways. In the present study, we examined the effect of HCB on cell proliferation and AhR expression, using an initiation-promotion hepatocarcinogenesis protocol in rat liver and in the human-derived hepatoma cell line, Hep<span class="hlt">G</span><span class="hlt">2</span>. Female Wistar rats were initiated with a single dose of 100 mg/kg of diethylnitrosamine (DEN) at the start of the experiment. Two weeks later, daily dosing of 100 mg/kg HCB was maintained for 10 weeks. Partial hepatectomy was performed 3 weeks after initiation. The number and area of glutathione S-transferase-P (GST-P)-positive foci, in the rat liver were used as biomarkers of liver precancerous lesions. Immunohistochemical staining showed an increase in proliferating cell nuclear antigen (PCNA)-positive cells, along with enhanced AhR protein expression in hepatocytes within GST-P-positive foci of (DEN HCB) group, when compared to DEN. In a similar manner, Western blot analysis demonstrated that HCB <span class="hlt">induced</span> PCNA and AhR protein expression in Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Flow cytometry assay indicated that the cells were accumulated at S and <span class="hlt">G</span><span class="hlt">2</span>/M phases of the cell cycle. HCB increased cyclin D1 protein levels and ERK1/<span class="hlt">2</span> phosphorylation in a dose-dependent manner. Treatment of cells with a selective MEK1 inhibitor, prevented HCB-stimulatory effect on PCNA and cyclinD1, indicating that these effects</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1852522','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1852522"><span>Curcumin <span class="hlt">Induces</span> <span class="hlt">G</span><span class="hlt">2</span>/M Arrest and Apoptosis in Cisplatin-Resistant Human Ovarian Cancer Cells by Modulating Akt and p38 MAPK</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Weir, Nathan M.; Selvendiran, Karuppaiyah; Kutala, Vijay Kumar; Tong, Liyue; Vishwanath, Shilpa; Rajaram, Murugesan; Tridandapani, Susheela; Anant, Shrikant; Kuppusamy, Periannan</p> <p>2007-01-01</p> <p>Curcumin, a major active component of turmeric, is known to <span class="hlt">induce</span> apoptosis in several types of cancer cells, but little is known about its activity in chemoresistant cells. Hence, the aim of the present study was to investigate the anticancer properties of curcumin in cisplatin-resistant human ovarian cancer cells in vitro. The results indicated that curcumin inhibited the proliferation of both cisplatin-resistant (CR) and sensitive (CS) human ovarian cancer cells almost equally. Enhanced superoxide generation was observed in both CR and CS cells treated with curcumin. Curcumin <span class="hlt">induced</span> <span class="hlt">G</span><span class="hlt">2</span>/M phase cell-cycle arrest in CR cells by enhancing the p53 phosphorylation and apoptosis through the activation of caspase-3 followed by PARP degradation. Curcumin also inhibited the phosphorylation of Akt while the phosphorylation of p38 MAPK was enhanced. In summary, our results showed that curcumin inhibits the proliferation of cisplatin-resistant ovarian cancer cells through the induction of superoxide generation, <span class="hlt">G</span><span class="hlt">2</span>/M arrest, and apoptosis. PMID:17218783</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=chip&pg=6&id=EJ1036774','ERIC'); return false;" href="https://eric.ed.gov/?q=chip&pg=6&id=EJ1036774"><span>Model <span class="hlt">Breaking</span> Points Conceptualized</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Vig, Rozy; Murray, Eileen; Star, Jon R.</p> <p>2014-01-01</p> <p>Current curriculum initiatives (e.<span class="hlt">g</span>., National Governors Association Center for Best Practices and Council of Chief State School Officers 2010) advocate that models be used in the mathematics classroom. However, despite their apparent promise, there comes a point when models <span class="hlt">break</span>, a point in the mathematical problem space where the model cannot,…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27209165','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27209165"><span>MicroRNA expression in the vildagliptin-treated two- and three-dimensional Hep<span class="hlt">G</span><span class="hlt">2</span> cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yamashita, Yasunari; Asakura, Mitsutoshi; Mitsugi, Ryo; Fujii, Hideaki; Nagai, Kenichiro; Atsuda, Koichiro; Itoh, Tomoo; Fujiwara, Ryoichi</p> <p>2016-06-01</p> <p>Vildagliptin is an inhibitor of dipeptidyl peptidase-4 that is used for the treatment of type <span class="hlt">2</span> diabetes mellitus. While vildagliptin can <span class="hlt">induce</span> hepatic dysfunction in humans, the molecular mechanism has not been determined yet. Recent studies indicated that certain types of microRNA (miRNA) were linking to the development of drug-<span class="hlt">induced</span> hepatotoxicity. In the present study, therefore, we identified hepatic miRNAs that were highly <span class="hlt">induced</span> or reduced by the vildagliptin treatment in mice. MiR-222 and miR-877, toxicity-associated miRNAs, were <span class="hlt">induced</span> 31- and 53-fold, respectively, by vildagliptin in the liver. While a number of miRNAs were significantly regulated by the orally treated vildagliptin in vivo, such regulation was not observed in the vildagliptin-treated Hep<span class="hlt">G</span><span class="hlt">2</span> cells. In addition to the regular two-dimensional (<span class="hlt">2</span>D) culture, we carried out the three-dimensional (3D) culturing of Hep<span class="hlt">G</span><span class="hlt">2</span> cells. In the 3D-Hep<span class="hlt">G</span><span class="hlt">2</span> cells, a significant reduction of miR-222 was observed compared to the expression level in <span class="hlt">2</span>D-Hep<span class="hlt">G</span><span class="hlt">2</span> cells. A slight induction of miR-222 by vildagliptin was observed in the 3D-Hep<span class="hlt">G</span><span class="hlt">2</span> cells, although miR-877 was not <span class="hlt">induced</span> by vildagliptin even in the 3D-Hep<span class="hlt">G</span><span class="hlt">2</span> cells. Further investigations are needed to overcome the discrepancy in the responsiveness of the miRNA expressions to vildagliptin between in vivo and in vitro. Copyright © 2016 The Japanese Society for the Study of Xenobiotics. Published by Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25111860','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25111860"><span>Alkali-soluble polysaccharide, isolated from Lentinus edodes, <span class="hlt">induces</span> apoptosis and <span class="hlt">G</span><span class="hlt">2</span>/M cell cycle arrest in H22 cells through microtubule depolymerization.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>You, Ru-Xu; Liu, Jin-Yu; Li, Shi-Jun; Wang, Liu; Wang, Kai-Ping; Zhang, Yu</p> <p>2014-12-01</p> <p>The aim of the study was to evaluate the pro-apoptotic effects of polysaccharides derived from Lentinus edodes and further elucidated the mechanisms of this action. Our results demonstrated that marked morphological changes of apoptosis were observed after treatment of L. edodes polysaccharides [Lentinan (LTN)]. Moreover, LTN-<span class="hlt">induced</span> cell apoptosis was characterized by a rapid stimulation of reactive oxygen species production, the loss of mitochondrial membrane potential and an increase in intracellular concentration of Ca(<span class="hlt">2</span>+) . In addition, the results of the haematoxylin and eosin and TUNEL assay further confirmed that LTN-<span class="hlt">induced</span> apoptosis in vivo. Furthermore, flow cytometry analysis showed that LTN could arrest the cell cycle at <span class="hlt">G</span><span class="hlt">2</span>/M phase, and immunofluorescence showed LTN caused disruption of microtubule. These results suggest that disruption of cellular microtubule network, arrest of the cell cycle at <span class="hlt">G</span><span class="hlt">2</span>/M phase and induction of apoptosis may be one of the possible mechanisms of anti-tumour effect of LTN. Copyright © 2014 John Wiley & Sons, Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/5851472-single-apg-cis-diamminedichloroplatinum-ii-adduct-induced-mutagenesis-escherichia-coli','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/5851472-single-apg-cis-diamminedichloroplatinum-ii-adduct-induced-mutagenesis-escherichia-coli"><span>Single d(Ap<span class="hlt">G</span>)/cis-diamminedichloroplatinum(II) adduct-<span class="hlt">induced</span> mutagenesis in Escherichia coli</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Burnouf, D.; Fuchs, R.P.P.; Gauthier, C.</p> <p>1990-08-01</p> <p>The mutation spectrum <span class="hlt">induced</span> by the widely used antitumor drug cis-diamminedichloroplatinum(II) (cis-DDP) showed that cisDDP(d(Ap<span class="hlt">G</span>)) adducts, although they account for only 25% of the lesions formed are {approx}5 times more mutagenic than the major GG adduct. The authors report the construction of vectors bearing a single cisDDP(d(Ap<span class="hlt">G</span>)) lesion and their use in mutagenesis experiments in Escherichia coli. The mutagenic processing of the lesion is found to depend strictly on induction of the SOS system of the bacterial host cells. In SOS-<span class="hlt">induced</span> cells, mutation frequencies of 1-<span class="hlt">2</span>% were detected. All these mutations are targeted to the 5{prime} base of the adduct.more » Single A {yields} T transversions are mainly observed (80%), whereas A {yields} <span class="hlt">G</span> transitions account for 10% of the total mutations. Tandem base-pair substitutions involving the adenine residue and the thymine residue immediately 5{prime} to the adduct occur at a comparable frequency (10%). No selective loss of the strand bearing the platinum adduct was seen, suggesting that, in vivo, cisDDP(d(Ap<span class="hlt">G</span>)) adducts are not blocking lesions. The high mutation specificity of cisDDP-(d(Ap<span class="hlt">G))-induced</span> mutagenesis is discussed in relation to structural data.« less</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li class="active"><span>25</span></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_25 --> <div class="footer-extlink text-muted" style="margin-bottom:1rem; text-align:center;">Some links on this page may take you to non-federal websites. Their policies may differ from this site.</div> </div><!-- container --> <footer><a id="backToTop" href="#top"> </a><nav><a id="backToTop" href="#top"> </a><ul class="links"><a id="backToTop" href="#top"> </a><li><a id="backToTop" href="#top"></a><a href="/sitemap.html">Site Map</a></li> <li><a href="/members/index.html">Members Only</a></li> <li><a href="/website-policies.html">Website Policies</a></li> <li><a href="https://doe.responsibledisclosure.com/hc/en-us" target="_blank">Vulnerability Disclosure Program</a></li> <li><a href="/contact.html">Contact Us</a></li> </ul> <div class="small">Science.gov is maintained by the U.S. Department of Energy's <a href="https://www.osti.gov/" target="_blank">Office of Scientific and Technical Information</a>, in partnership with <a href="https://www.cendi.gov/" target="_blank">CENDI</a>.</div> </nav> </footer> <script type="text/javascript"><!-- // var lastDiv = ""; function showDiv(divName) { // hide last div if (lastDiv) { document.getElementById(lastDiv).className = "hiddenDiv"; } //if value of the box is not nothing and an object with that name exists, then change the class if (divName && document.getElementById(divName)) { document.getElementById(divName).className = "visibleDiv"; lastDiv = divName; } } //--> </script> <script> /** * Function that tracks a click on an outbound link in Google Analytics. * This function takes a valid URL string as an argument, and uses that URL string * as the event label. */ var trackOutboundLink = function(url,collectionCode) { try { h = window.open(url); setTimeout(function() { ga('send', 'event', 'topic-page-click-through', collectionCode, url); }, 1000); } catch(err){} }; </script> <!-- Google Analytics --> <script> (function(i,s,o,g,r,a,m){i['GoogleAnalyticsObject']=r;i[r]=i[r]||function(){ (i[r].q=i[r].q||[]).push(arguments)},i[r].l=1*new Date();a=s.createElement(o), m=s.getElementsByTagName(o)[0];a.async=1;a.src=g;m.parentNode.insertBefore(a,m) })(window,document,'script','//www.google-analytics.com/analytics.js','ga'); ga('create', 'UA-1122789-34', 'auto'); ga('send', 'pageview'); </script> <!-- End Google Analytics --> <script> showDiv('page_1') </script> </body> </html>