308
Views
4
CrossRef citations to date
0
Altmetric
Review

Targeting gap junctional intercellular communication by hepatocarcinogenic compounds

ORCID Icon, ORCID Icon, ORCID Icon, ORCID Icon, ORCID Icon, ORCID Icon & ORCID Icon show all

References

  • Aasen, T., E. Leithe, S. V. Graham, P. Kameritsch, M. D. Mayan, M. Mesnil, K. Pogoda, and A. Tabernero. 2019. Connexins in cancer: Bridging the gap to the clinic. Oncogene 38:4429–51.
  • Abou Hashieh, I., S. Mathieu, F. Besson, and A. Gerolami. 1996. Inhibition of gap junction intercellular communications of cultured rat hepatocytes by ethanol: Role of ethanol metabolism. J. Hepatol. 24:360–67.
  • Anderson, C., H. Catoe, and R. Werner. 2006. MIR-206 regulates connexin43 expression during skeletal muscle development. Nucleic Acids Res. 34:5863–71.
  • Andrysik, Z., J. Prochazkova, M. Kabatkova, L. Umannova, P. Simeckova, J. Kohoutek, A. Kozubik, M. Machala, and J. Vondracek. 2013. Aryl hydrocarbon receptor-mediated disruption of contact inhibition is associated with connexin43 downregulation and inhibition of gap junctional intercellular communication. Arch. Toxicol. 87:491–503.
  • Bager, Y., K. Kenne, V. Krutovskikh, M. Mesnil, O. Traub, and L. Warngard. 1994. Alteration in expression of gap junction proteins in rat liver after treatment with the tumour promoter 3,4,5,3ʹ,4ʹ-pentachlorobiphenyl. Carcinogenesis 15:2439–43.
  • Bager, Y., M. C. Lindebro, P. Martel, C. Chaumontet, and L. Warngard. 1997. Altered function, localization and phosphorylation of gap junctions in rat liver epithelial, IAR 20, cells after treatment with PCBs or TCDD. Environ. Toxicol. Pharmacol. 3:257–66.
  • Baker, T. K., A. P. Kwiatkowski, B. V. Madhukar, and J. E. Klaunig. 1995. Inhibition of gap junctional intercellular communication by 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in rat hepatocytes. Carcinogenesis 16:2321–26.
  • Balasubramaniyan, V., D. K. Dhar, A. E. Warner, W. Y. V. Li, A. F. Amiri, B. Bright, R. P. Mookerjee, N. A. Davies, D. L. Becker, and R. Jalan. 2013. Importance of Connexin-43 based gap junction in cirrhosis and acute-on-chronic liver failure. J. Hepatol. 58:1194–200.
  • Bargiello, T. A., S. Oh, Q. Tang, N. K. Bargiello, T. L. Dowd, and T. Kwon. 2018. Gating of Connexin Channels by transjunctional-voltage: Conformations and models of open and closed states. Biochim. Biophys. Acta Biomembr. 1860:22–39.
  • Barhoumi, R., and R. C. Burghardt. 1996. Kinetic analysis of the chronology of patulin- and gossypol-induced cytotoxicity in vitro. Fundam. Appl. Toxicol. 30:290–97.
  • Berthoud, V. M., V. Iwanij, A. M. Garcia, and J. C. Saez. 1992. Connexins and glucagon receptors during development of rat hepatic acinus. Am. J. Physiol. 263:650–58.
  • Bokkala, S., H. M. Reis, E. Rubin, and S. K. Joseph. 2001. Effect of angiotensin II and ethanol on the expression of connexin 43 in WB rat liver epithelial cells. Biochem. J. 357:769–77.
  • Braeuning, A., A. Oberemm, T. Heise, U. Gundert-Remy, J. G. Hengstler, and A. Lampen. 2019. In vitro proteomic analysis of methapyrilene toxicity in rat hepatocytes reveals effects on intermediary metabolism. Arch. Toxicol. 93:369–83.
  • Chipman, J. K., A. Mally, and G. O. Edwards. 2003. Disruption of gap junctions in toxicity and carcinogenicity. Toxicol. Sci. 71:146–53.
  • Cogliati, B., S. Crespo Yanguas, T. C. da Silva, T. P. A. Aloia, M. S. Nogueira, M. A. Real-Lima, L. M. Chaible, D. S. Sanches, J. Willebrords, M. Maes, et al. 2016. Connexin32 deficiency exacerbates carbon tetrachloride-induced hepatocellular injury and liver fibrosis in mice. Toxicol. Mech. Meth. 26:362–70.
  • Cogliati, B., T. C. Da Silva, T. P. Aloia, L. M. Chaible, M. A. Real-Lima, D. S. Sanches, P. Matsuzaki, F. J. Hernandez-Blazquez, and M. L. Dagli. 2011. Morphological and molecular pathology of CCL4-induced hepatic fibrosis in connexin43-deficient mice. Microsc. Res. Tech. 74:421–29.
  • Cooper, R. G., and T. Magwere. 2008. Nitric oxide-mediated pathogenesis during nicotine and alcohol consumption. Indian J. Physiol. Pharmacol. 52:11–18.
  • Cooreman, A., R. Van Campenhout, S. Ballet, P. Annaert, B. Van Den Bossche, I. Colle, B. Cogliati, and M. Vinken. 2019. Connexin and pannexin (hemi)channels: Emerging targets in the treatment of liver disease. Hepatology 69:1317–23.
  • Correa, P. R., M. T. Guerra, M. F. Leite, D. C. Spray, and M. H. Nathanson. 2004. Endotoxin unmasks the role of gap junctions in the liver. Biochem. Biophys. Res. Commun. 322:718–26.
  • Corton, J. C., M. L. Cunningham, B. T. Hummer, C. Lau, B. Meek, J. M. Peters, J. A. Popp, L. Rhomberg, J. Seed, and J. E. Klaunig. 2014. Mode of action framework analysis for receptor-mediated toxicity: The peroxisome proliferator-activated receptor alpha (PPARalpha) as a case study. Crit. Rev. Toxicol. 44:1–49.
  • Cowles, C., A. Mally, and J. K. Chipman. 2007. Different mechanisms of modulation of gap junction communication by non-genotoxic carcinogens in rat liver in vivo. Toxicology 238:49–59.
  • Crespo Yanguas, S., T. C. da Silva, I. V. A. Pereira, J. Willebrords, M. Maes, M. Sayuri Nogueira, I. Alves de Castro, I. Leclercq, G. R. Romualdo, L. F. Barbisan, et al. 2018. TAT-Gap19 and carbenoxolone alleviate liver fibrosis in mice. Int. J. Mol. Sci 19:817.
  • De Maio, A., C. Gingalewski, N. G. Theodorakis, and M. G. Clemens. 2000. Interruption of hepatic gap junctional communication in the rat during inflammation induced by bacterial lipopolysaccharide. Shock 14:53–59.
  • Decrock, E., M. Vinken, M. Bol, K. D’Herde, V. Rogiers, P. Vandenabeele, D. V. Krysko, G. Bultynck, and L. Leybaert. 2011. Calcium and connexin-based intercellular communication, a deadly catch? Cell Calcium 50:310–21.
  • Decrock, E., M. Vinken, E. De Vuyst, D. V. Krysko, K. D’Herde, T. Vanhaecke, P. Vandenabeele, V. Rogiers, and L. Leybaert. 2009. Connexin-related signaling in cell death: To live or let die? Cell Death Differ. 16:524–36.
  • Doktorova, T. Y., G. Ates, M. Vinken, T. Vanhaecke, and V. Rogiers. 2014a. Way forward in case of a false positive in vitro genotoxicity result for a cosmetic substance? Toxicol. In Vitro 28:54–59.
  • Doktorova, T. Y., M. Pauwels, M. Vinken, T. Vanhaecke, and V. Rogiers. 2012. Opportunities for an alternative integrating testing strategy for carcinogen hazard assessment? Crit. Rev. Toxicol. 42:91–106.
  • Doktorova, T. Y., R. Yildirimman, L. Ceelen, M. Vilardell, T. Vanhaecke, M. Vinken, G. Ates, A. Heymans, H. Gmuender, R. Bort, et al. 2014b. Testing chemical carcinogenicity by using a transcriptomics HepaRG-based model? Excli J 13:623–37.
  • Eastmond, D. A. 2008. Evaluating genotoxicity data to identify a mode of action and its application in estimating cancer risk at low doses: A case study involving carbon tetrachloride. Environ. Mol. Mutagen. 49:132–41.
  • Elaut, G., T. Henkens, P. Papeleu, S. Snykers, M. Vinken, T. Vanhaecke, and V. Rogiers. 2006. Molecular mechanisms underlying the dedifferentiation process of isolated hepatocytes and their cultures. Curr. Drug Metab. 7:629–60.
  • Elcock, F. J., J. K. Chipman, and R. A. Roberts. 1998. The rodent nongenotoxic hepatocarcinogen and peroxisome proliferator nafenopin inhibits intercellular communication in rat but not guinea-pig hepatocytes, perturbing S-phase but not apoptosis. Arch. Toxicol. 72:439–44.
  • Elcock, F. J., E. Deag, R. A. Roberts, and J. K. Chipman. 2000. Nafenopin causes protein kinase C-mediated serine phosphorylation and loss of function of connexin 32 protein in rat hepatocytes without aberrant expression or localization. Toxicol. Sci. 56:86–94.
  • Eugenin, E. A., H. E. Gonzalez, H. A. Sanchez, M. C. Branes, and J. C. Saez. 2007. Inflammatory conditions induce gap junctional communication between rat Kupffer cells both in vivo and in vitro. Cell. Immunol. 247:103–10.
  • Fallon, M. B., M. H. Nathanson, A. Mennone, J. C. Saez, A. D. Burgstahler, and J. M. Anderson. 1995. Altered expression and function of hepatocyte gap junctions after common bile duct ligation in the rat. Am. J. Physiol. 268:1186–94.
  • Fischer, R., R. Reinehr, T. P. Lu, A. Schonicke, U. Warskulat, H. P. Dienes, and D. Haussinger. 2005. Intercellular communication via gap junctions in activated rat hepatic stellate cells. Gastroenterology 128:433–48.
  • Gagliano, N., I. D. Donne, C. Torri, M. Migliori, F. Grizzi, A. Milzani, C. Filippi, G. Annoni, P. Colombo, F. Costa, et al. 2006. Early cytotoxic effects of ochratoxin A in rat liver: A morphological, biochemical and molecular study. Toxicology 225:214–24.
  • Garcia, I. E., H. A. Sanchez, A. D. Martinez, and M. A. Retamal. 2018. Redox-mediated regulation of connexin proteins; focus on nitric oxide. Biochim. Biophys. Acta Biomembr. 1860:91–95.
  • Garciarena, C. D., A. Malik, P. Swietach, A. P. Moreno, and R. D. Vaughan-Jones. 2018. Distinct moieties underlie biphasic H(+) gating of connexin43 channels, producing a pH optimum for intercellular communication. Faseb J. 32:1969–81.
  • Gingalewski, C., K. Wang, M. G. Clemens, and A. De Maio. 1996. Posttranscriptional regulation of connexin 32 expression in liver during acute inflammation. J. Cell Physiol. 166:461–67.
  • Goel, G., H. P. Makkar, G. Francis, and K. Becker. 2007. Phorbol esters: Structure, biological activity, and toxicity in animals. Int. J. Toxicol. 26:279–88.
  • Gonzalez, H. E., E. A. Eugenin, G. Garces, N. Solis, M. Pizarro, L. Accatino, and J. C. Saez. 2002. Regulation of hepatic connexins in cholestasis: Possible involvement of Kupffer cells and inflammatory mediators. Am. J. Physiol. Gastrointest. Liver Physiol. 282:G991–G1001.
  • Greenwel, P., J. Rubin, M. Schwartz, E. L. Hertzberg, and M. Rojkind. 1993. Liver fat-storing cell clones obtained from a CCl4-cirrhotic rat are heterogeneous with regard to proliferation, expression of extracellular-matrix components, interleukin-6, and connexin-43. Lab. Invest. 69:210–16.
  • Grimm, F. A., D. Hu, I. Kania-Korwel, H. J. Lehmler, G. Ludewig, K. C. Hornbuckle, M. W. Duffel, A. Bergman, and L. W. Robertson. 2015. Metabolism and metabolites of polychlorinated biphenyls. Crit. Rev. Toxicol. 45:245–72.
  • Guan, X., W. J. Bonney, and R. J. Ruch. 1995. Changes in gap junction permeability, gap junction number, and connexin43 expression in lindane-treated rat liver epithelial cells. Toxicol. Appl. Pharmacol. 130:79–86.
  • Guo, X., J. E. Seo, X. Li, and N. Mei. 2020. Genetic toxicity assessment using liver cell models: Past, present, and future. J. Toxicol. Environ. Health B. 23:27–50.
  • Hernandez, L. G., J. van Benthem, and G. E. Johnson. 2013. A mode-of-action approach for the identification of genotoxic carcinogens. PLoS ONE 8:e64532.
  • Hernandez, L. G., H. van Steeg, M. Luijten, and J. van Benthem. 2009. Mechanisms of non-genotoxic carcinogens and importance of a weight of evidence approach. Mutat. Res. 682:94–109.
  • Hernandez-Guerra, M., Y. Gonzalez-Mendez, Z. A. de Ganzo, E. Salido, J. C. Garcia-Pagan, B. Abrante, A. M. Malagon, J. Bosch, and E. Quintero. 2014. Role of gap junctions modulating hepatic vascular tone in cirrhosis. Liver Int. 34:859–68.
  • Hernandez-Guerra, M., A. Hadjihambi, and R. Jalan. 2019. Gap junctions in liver disease: Implications for pathogenesis and therapy. J. Hepatol. 70:759–72.
  • Herrmann, S., M. Seidelin, H. C. Bisgaard, and O. Vang. 2002. Indolo[3,2-b]carbazole inhibits gap junctional intercellular communication in rat primary hepatocytes and acts as a potential tumor promoter. Carcinogenesis 23:1861–68.
  • Hokaiwado, N., M. Asamoto, M. Futakuchi, K. Ogawa, S. Takahashi, and T. Shirai. 2007. Both early and late stages of hepatocarcinogenesis are enhanced in cx32 dominant negative mutant transgenic rats with disrupted gap junctional intercellular communication. J. Membr. Biol. 218:101–06.
  • Hong, X., Q. Wang, Y. Yang, S. Zheng, X. Tong, S. Zhang, L. Tao, and A. L. Harris. 2012. Gap junctions propagate opposite effects in normal and tumor testicular cells in response to cisplatin. Cancer Lett. 317:165–71.
  • Hu, D., H. Zou, T. Han, J. Xie, N. Dai, L. Zhuo, J. Gu, J. Bian, Y. Yuan, X. Liu, et al. 2016. Gap junction blockage promotes cadmium-induced apoptosis in BRL 3A derived from Buffalo rat liver cells. J. Vet. Sci. 17:63–70.
  • Huong, B. T. M., L. D. Tuyen, D. H. Tuan, L. Brimer, and A. Dalsgaard. 2016. Dietary exposure to aflatoxin B1, ochratoxin A and fuminisins of adults in Lao Cai province, Viet Nam: A total dietary study approach. Food Chem. Toxicol. 98:127–33.
  • Hutchinson, R. W., R. Barhoumi, J. M. Miles, and R. C. Burghardt. 1998. Attenuation of gossypol cytotoxicity by cyclic AMP in a rat liver cell line. Toxicol. Appl. Pharmacol. 151:311–18.
  • IARC. 2018. In DDT, Lindane, and 2,4-D. Lyon (FR).
  • Imlay, J. A. 2008. Cellular defenses against superoxide and hydrogen peroxide. Annu. Rev. Biochem. 77:755–76.
  • Isenberg, J. S., L. M. Kamendulis, D. C. Ackley, J. H. Smith, G. Pugh Jr., A. W. Lington, R. H. McKee, and J. E. Klaunig. 2001. Reversibility and persistence of di-2-ethylhexyl phthalate (DEHP)- and phenobarbital-induced hepatocellular changes in rodents. Toxicol. Sci. 64:192–99.
  • Isenberg, J. S., L. M. Kamendulis, J. H. Smith, D. C. Ackley, G. Pugh Jr., A. W. Lington, and J. E. Klaunig. 2000. Effects of Di-2-ethylhexyl phthalate (DEHP) on gap-junctional intercellular communication (GJIC), DNA synthesis, and peroxisomal beta oxidation (PBOX) in rat, mouse, and hamster liver. Toxicol. Sci. 56:73–85.
  • Ito, Y., K. Nakajima, Y. Masubuchi, S. Kikuchi, F. Saito, Y. Akahori, M. Jin, T. Yoshida, and M. Shibutani. 2019. Differential responses on energy metabolic pathway reprogramming between genotoxic and non-genotoxic hepatocarcinogens in rat liver cells. J. Toxicol. Pathol. 32:261–74.
  • Jeon, S. H., M. H. Cho, and J. H. Cho. 2001. Effects of cadmium on gap junctional intercellular communication in WB-F344 rat liver epithelial cells. Human Exp. Toxicol. 20:577–83.
  • Jeong, S. H., S. S. Habeebu, and C. D. Klaassen. 2000. Cadmium decreases gap junctional intercellular communication in mouse liver. Toxicol. Sci. 57:156–66.
  • Jiang, Y., J. Chen, J. Tong, and T. Chen. 2014. Trichloroethylene-induced gene expression and DNA methylation changes in B6C3F1 mouse liver. PLoS ONE 9:e116179.
  • Jiang, Y., J. Chen, C. Yue, H. Zhang, J. Tong, J. Li, and T. Chen. 2017. The role of miR-182-5p in hepatocarcinogenesis of trichloroethylene in mice. Toxicol. Sci. 156:208–16.
  • Jung, J. W., S. D. Cho, N. S. Ahn, S. R. Yang, J. S. Park, E. H. Jo, J. W. Hwang, O. I. Aruoma, Y. S. Lee, and K. S. Kang. 2006. Effects of the histone deacetylases inhibitors sodium butyrate and trichostatin A on the inhibition of gap junctional intercellular communication by H2O2- and 12-O-tetradecanoylphorbol-13-acetate in rat liver epithelial cells. Cancer Lett. 241:301–08.
  • Kabak, B., A. D. Dobson, and I. Var. 2006. Strategies to prevent mycotoxin contamination of food and animal feed: A review. Crit Rev Food Sci. Nutr. 46:593–619.
  • Kabatkova, M., J. Svobodova, K. Pencikova, D. S. Mohatad, L. Smerdova, A. Kozubik, M. Machala, and J. Vondracek. 2015. Interactive effects of inflammatory cytokine and abundant low-molecular-weight PAHs on inhibition of gap junctional intercellular communication, disruption of cell proliferation control, and the AhR-dependent transcription. Toxicol. Lett. 232:113–21.
  • Kamendulis, L. M., J. S. Isenberg, J. H. Smith, G. Pugh Jr., A. W. Lington, and J. E. Klaunig. 2002. Comparative effects of phthalate monoesters on gap junctional intercellular communication and peroxisome proliferation in rodent and primate hepatocytes. J. Toxicol. Environ. Health A 65:569–88.
  • Karin, M., and D. Dhar. 2016. Liver carcinogenesis: From naughty chemicals to soothing fat and the surprising role of NRF2. Carcinogenesis 37:541–46.
  • Kataoka, C., M. Nihei, M. Nimata, and K. Sawadaishi. 2016. Development of a model immunoassay utilizing monoclonal antibodies of different specificities for quantitative determination of dieldrin and heptachlors in their mixtures. J. Agric. Food Chem. 64:8950–57.
  • Kato, H., A. Naiki-Ito, T. Naiki, S. Suzuki, Y. Yamashita, S. Sato, H. Sagawa, A. Kato, T. Kuno, and S. Takahashi. 2016. Connexin 32 dysfunction promotes ethanol-related hepatocarcinogenesis via activation of Dusp1-Erk axis. Oncotarget 7:2009–21.
  • Kawasaki, Y., A. Kubomoto, and H. Yamasaki. 2007. Control of intracellular localization and function of Cx43 by SEMA3F. J. Membr. Biol. 217:53–61.
  • Kawasaki, Y., Y. Omori, Q. Li, Y. Nishikawa, T. Yoshioka, M. Yoshida, K. Ishikawa, and K. Enomoto. 2011. Cytoplasmic accumulation of connexin32 expands cancer stem cell population in human HuH7 hepatoma cells by enhancing its self-renewal. Int. J. Cancer 128:51–62.
  • Kenne, K., R. Fransson-Steen, S. Honkasalo, and L. Warngard. 1994. Two inhibitors of gap junctional intercellular communication, TPA and endosulfan: Different effects on phosphorylation of connexin 43 in the rat liver epithelial cell line, IAR 20. Carcinogenesis 15:1161–65.
  • Kersten, S., and R. Stienstra. 2017. The role and regulation of the peroxisome proliferator activated receptor alpha in human liver. Biochimie 136:75–84.
  • Klaunig, J. E., R. J. Ruch, and E. L. Lin. 1989. Effects of trichloroethylene and its metabolites on rodent hepatocyte intercellular communication. Toxicol. Appl. Pharmacol. 99:454–65.
  • Klaunig, J. E., R. J. Ruch, and C. M. Weghorst. 1990. Comparative effects of phenobarbital, DDT, and lindane on mouse hepatocyte gap junctional intercellular communication. Toxicol. Appl. Pharmacol. 102:553–63.
  • Koffler, L. D., M. J. Fernstrom, T. E. Akiyama, F. J. Gonzalez, and R. J. Ruch. 2002. Positive regulation of connexin32 transcription by hepatocyte nuclear factor-1alpha. Arch. Biochem. Biophys. 407:160–67.
  • Krutovskikh, V. A., M. Mesnil, G. Mazzoleni, and H. Yamasaki. 1995. Inhibition of rat liver gap junction intercellular communication by tumor-promoting agents in vivo. Association with aberrant localization of connexin proteins. Lab. Invest. 72:571–77.
  • Kushida, M., T. Sukata, S. Uwagawa, K. Ozaki, A. Kinoshita, H. Wanibuchi, K. Morimura, Y. Okuno, and S. Fukushima. 2005. Low dose DDT inhibition of hepatocarcinogenesis initiated by diethylnitrosamine in male rats: Possible mechanisms. Toxicol. Appl. Pharmacol. 208:285–94.
  • Leithe, E., M. Mesnil, and T. Aasen. 2018. The connexin 43 C-terminus: A tail of many tales. Biochim. Biophys. Acta Biomembr. 1860:48–64.
  • Leithe, E., and E. Rivedal. 2004. Ubiquitination and down-regulation of gap junction protein connexin-43 in response to 12-O-tetradecanoylphorbol 13-acetate treatment. J. Biol. Chem. 279:50089–96.
  • Li, Q., Y. Omori, Y. Nishikawa, T. Yoshioka, Y. Yamamoto, and K. Enomoto. 2007. Cytoplasmic accumulation of connexin32 protein enhances motility and metastatic ability of human hepatoma cells in vitro and in vivo. Int. J. Cancer. 121:536–46.
  • Li, S., J. Zhao, R. Huang, M. F. Santillo, K. A. Houck, and M. Xia. 2019. Use of high-throughput enzyme-based assay with xenobiotic metabolic capability to evaluate the inhibition of acetylcholinesterase activity by organophosphorous pesticides. Toxicol. In Vitro 56:93–100.
  • Li, X., L. Yu, J. Gao, X. Bi, J. Zhang, S. Xu, M. Wang, M. Chen, F. Qiu, and G. Fu. 2018. Apelin ameliorates high glucose-induced downregulation of connexin 43 via AMPK-dependent pathway in neonatal rat cardiomyocytes. Aging Dis 9:66–76.
  • Liang, Z., T. Li, S. Jiang, J. Xu, W. Di, Z. Yang, W. Hu, and Y. Yang. 2017. AMPK: A novel target for treating hepatic fibrosis. Oncotarget 8:62780–92.
  • Lin, Y. S., J. L. Caffrey, P. C. Hsu, M. H. Chang, M. F. Faramawi, and J. W. Lin. 2012. Environmental exposure to dioxin-like compounds and the mortality risk in the U.S. population. Int. J. Hyg. Environ. Health 215:541–46.
  • Lin, Z. J., J. Ming, L. Yang, J. Z. Du, N. Wang, and H. J. Luo. 2016. Mechanism of regulatory effect of microRNA-206 on connexin 43 in distant metastasis of breast cancer. Chin. Med. J. (Engl) 129:424–34.
  • Liu, J. 2014. Ethanol and liver: Recent insights into the mechanisms of ethanol-induced fatty liver. World J. Gastroenterol. 20:14672–85.
  • Loch-Caruso, R., M. M. Galvez, K. Brant, and D. Chung. 2004. Cell and toxicant specific phosphorylation of conexin43: Effects of lindane and TPA on rat myometrial and WB-F344 liver cell gap junctions. Cell Biol. Toxicol. 20:147–69.
  • Lu, Y., and A. I. Cederbaum. 2008. CYP2E1 and oxidative liver injury by alcohol. Free Radic. Biol. Med. 44:723–38.
  • Luther, J., M. K. Gala, N. Borren, R. Masia, R. P. Goodman, I. H. Moeller, E. DiGiacomo, A. Ehrlich, A. Warren, M. L. Yarmush, et al. 2018. Hepatic connexin 32 associates with nonalcoholic fatty liver disease severity. Hepatol. Commun. 2:786–97.
  • Machala, M., L. Blaha, J. Vondracek, J. E. Trosko, J. Scott, and B. L. Upham. 2003. Inhibition of gap junctional intercellular communication by noncoplanar polychlorinated biphenyls: Inhibitory potencies and screening for potential mode(s) of action. Toxicol. Sci. 76:102–11.
  • Maes, M., S. Crespo Yanguas, J. Willebrords, B. Cogliati, and M. Vinken. 2015a. Connexin and pannexin signaling in gastrointestinal and liver disease. Transl. Res. 166:332–43.
  • Maes, M., S. Crespo Yanguas, J. Willebrords, J. L. Weemhoff, T. C. da Silva, E. Decrock, M. Lebofsky, I. V. A. Pereira, L. Leybaert, A. Farhood, et al. 2017. Connexin hemichannel inhibition reduces acetaminophen-induced liver injury in mice. Toxicol. Lett. 278:30–37.
  • Maes, M., M. R. McGill, T. C. da Silva, C. Abels, M. Lebofsky, C. Maria Monteiro de Araujo, T. Tiburcio, I. Veloso Alves Pereira, J. Willebrords, S. Crespo Yanguas, et al. 2016. Involvement of connexin43 in acetaminophen-induced liver injury. Biochim. Biophys. Acta 1862:1111–21.
  • Maes, M., S. C. Yanguas, J. Willebrords, and M. Vinken. 2015b. Models and methods for in vitro testing of hepatic gap junctional communication. Toxicol. In Vitro 30:569–77.
  • Mally, A., and J. K. Chipman. 2002. Non-genotoxic carcinogens: Early effects on gap junctions, cell proliferation and apoptosis in the rat. Toxicology 180:233–48.
  • Mally, A., M. Decker, M. Bekteshi, and W. Dekant. 2006. Ochratoxin A alters cell adhesion and gap junction intercellular communication in MDCK cells. Toxicology 223:15–25.
  • Manclus, J. J., A. Abad, M. Y. Lebedev, F. Mojarrad, B. Mickova, J. V. Mercader, J. Primo, M. A. Miranda, and A. Montoya. 2004. Development of a monoclonal immunoassay selective for chlorinated cyclodiene insecticides. J. Agric. Food Chem. 52:2776–84.
  • Maqbool, F., S. Mostafalou, H. Bahadar, and M. Abdollahi. 2016. Review of endocrine disorders associated with environmental toxicants and possible involved mechanisms. Life Sci. 145:265–73.
  • Maroni, M., C. Colosio, A. Ferioli, and A. Fait. 2000. Biological monitoring of pesticide exposure: A review. Toxicology 143:1–118.
  • Marx-Stoelting, P., J. Mahr, T. Knorpp, S. Schreiber, M. F. Templin, T. Ott, A. Buchmann, and M. Schwarz. 2008. Tumor promotion in liver of mice with a conditional Cx26 knockout. Toxicol. Sci. 103:260–67.
  • Matesic, D. F., H. L. Rupp, W. J. Bonney, R. J. Ruch, and J. E. Trosko. 1994. Changes in gap-junction permeability, phosphorylation, and number mediated by phorbol ester and non-phorbol-ester tumor promoters in rat liver epithelial cells. Mol. Carcinogen. 10:226–36.
  • McVicker, B. L., D. J. Tuma, and C. A. Casey. 2007. Effect of ethanol on pro-apoptotic mechanisms in polarized hepatic cells. World J. Gastroenterol. 13:4960–66.
  • Menezes, A., G. H. Dos Reis, M. C. Oliveira-Nunes, F. Mariath, M. Cabanel, B. Pontes, N. G. Castro, J. M. de Brito, and K. Carneiro. 2019. Live cell imaging supports a key role for histone deacetylase as a molecular target during glioblastoma malignancy downgrade through tumor competence modulation. J. Oncol.
  • Miyashita, T., A. Takeda, M. Iwai, and T. Shimazu. 1991. Single administration of hepatotoxic chemicals transiently decreases the gap-junction-protein levels of connexin 32 in rat liver. Eur. J. Biochem. 196:37–42.
  • Moennikes, O., A. Buchmann, A. Romualdi, T. Ott, J. Werringloer, K. Willecke, and M. Schwarz. 2000. Lack of phenobarbital-mediated promotion of hepatocarcinogenesis in connexin32-null mice. Cancer Res. 60:5087–91.
  • Mograbi, B., E. Corcelle, N. Defamie, M. Samson, M. Nebout, D. Segretain, P. Fenichel, and G. Pointis. 2003. Aberrant Connexin 43 endocytosis by the carcinogen lindane involves activation of the ERK/mitogen-activated protein kinase pathway. Carcinogenesis 24:1415–23.
  • Mrema, E. J., F. M. Rubino, G. Brambilla, A. Moretto, A. M. Tsatsakis, and C. Colosio. 2013. Persistent organochlorinated pesticides and mechanisms of their toxicity. Toxicology 307:74–88.
  • Na, M. R., S. K. Koo, D. Y. Kim, S. D. Park, S. K. Rhee, K. W. Kang, and C. O. Joe. 1995. In vitro inhibition of gap junctional intercellular communication by chemical carcinogens. Toxicology 98:199–206.
  • Nakata, Y., M. Iwai, S. Kimura, and T. Shimazu. 1996. Prolonged decrease in hepatic connexin32 in chronic liver injury induced by carbon tetrachloride in rats. J. Hepatol. 25:529–37.
  • Neveu, M. J., J. R. Hully, D. L. Paul, and H. C. Pitot. 1990. Reversible alteration in the expression of the gap junctional protein connexin 32 during tumor promotion in rat liver and its role during cell proliferation. Cancer Commun. 2:21–31.
  • Nielsen, M. S., L. N. Axelsen, P. L. Sorgen, V. Verma, M. Delmar, and N. H. Holstein-Rathlou. 2012. Gap junctions. Compr. Physiol. 2:1981–2035.
  • Nohmi, T. 2018. Thresholds of genotoxic and non-genotoxic carcinogens. Toxicol. Res. 34:281–90.
  • Ogawa, K., P. Pitchakarn, S. Suzuki, T. Chewonarin, M. Tang, S. Takahashi, A. Naiki-Ito, S. Sato, S. Takahashi, M. Asamoto, et al. 2012. Silencing of connexin 43 suppresses invasion, migration and lung metastasis of rat hepatocellular carcinoma cells. Cancer Sci. 103:860–67.
  • Okamura, A., M. Kamijima, E. Shibata, K. Ohtani, K. Takagi, J. Ueyama, Y. Watanabe, M. Omura, H. Wang, G. Ichihara, et al. 2005. A comprehensive evaluation of the testicular toxicity of dichlorvos in Wistar rats. Toxicology 213:129–37.
  • Oyamada, M., Y. Oyamada, and T. Takamatsu. 2005. Regulation of connexin expression. Biochim. Biophys. Acta 1719:6–23.
  • Oyedele, O. A., C. N. Ezekiel, M. Sulyok, M. C. Adetunji, B. Warth, O. O. Atanda, and R. Krska. 2017. Mycotoxin risk assessment for consumers of groundnut in domestic markets in Nigeria. Int. J. Food Microbiol. 251:24–32.
  • Page, J. L., M. C. Johnson, K. M. Olsavsky, S. C. Strom, H. Zarbl, and C. J. Omiecinski. 2007. Gene expression profiling of extracellular matrix as an effector of human hepatocyte phenotype in primary cell culture. Toxicol. Sci. 97:384–97.
  • Patrizi, B., and M. Siciliani de Cumis. 2018. TCDD toxicity mediated by epigenetic mechanisms. Int. J. Mol. Sci 19:4101.
  • Peracchia, C. 2020. Calmodulin-mediated regulation of gap junction channels. Int. J. Mol. Sci. 21:485.
  • Piechocki, M. P., R. D. Burk, and R. J. Ruch. 1999. Regulation of connexin32 and connexin43 gene expression by DNA methylation in rat liver cells. Carcinogenesis 20:401–06.
  • Pierucci, F., A. Frati, R. Squecco, E. Lenci, C. Vicenti, J. Slavik, F. Francini, M. Machala, and E. Meacci. 2017. Non-dioxin-like organic toxicant PCB153 modulates sphingolipid metabolism in liver progenitor cells: Its role in Cx43-formed gap junction impairment. Arch. Toxicol. 91:749–60.
  • Plante, I., M. Charbonneau, and D. G. Cyr. 2002. Decreased gap junctional intercellular communication in hexachlorobenzene-induced gender-specific hepatic tumor formation in the rat. Carcinogenesis 23:1243–49.
  • Plante, I., M. Charbonneau, and D. G. Cyr. 2006. Activation of the integrin-linked kinase pathway downregulates hepatic connexin32 via nuclear Akt. Carcinogenesis 27:1923–29.
  • Plante, I., D. G. Cyr, and M. Charbonneau. 2007. Sexual dimorphism in the regulation of liver connexin32 transcription in hexachlorobenzene-treated rats. Toxicol. Sci. 96:47–57.
  • Pohanish, R. P. 2015. Sittig’s Handbook of Pesticides and Agricultural Chemicals. 2nd.
  • Pugh, G., Jr., J. S. Isenberg, L. M. Kamendulis, D. C. Ackley, L. J. Clare, R. Brown, A. W. Lington, J. H. Smith, and J. E. Klaunig. 2000. Effects of di-isononyl phthalate, di-2-ethylhexyl phthalate, and clofibrate in cynomolgus monkeys. Toxicol. Sci. 56:181–88.
  • Rahman, M. F., M. Mahboob, K. Danadevi, B. Saleha Banu, and P. Grover. 2002. Assessment of genotoxic effects of chloropyriphos and acephate by the comet assay in mice leucocytes. Mutat. Res. 516:139–47.
  • Rasic, D., V. Micek, M. S. Klaric, and M. Peraica. 2019. Oxidative stress as a mechanism of combined OTA and CTN toxicity in rat plasma, liver and kidney. Human Exp. Toxicol. 38:434–45.
  • Rasic, D., M. Mladinic, D. Zeljezic, A. Pizent, S. Stefanovic, D. Milicevic, P. Konjevoda, and M. Peraica. 2018. Effects of combined treatment with ochratoxin A and citrinin on oxidative damage in kidneys and liver of rats. Toxicon 146:99–105.
  • Riquelme, M. A., R. Kar, S. Gu, and J. X. Jiang. 2013. Antibodies targeting extracellular domain of connexins for studies of hemichannels. Neuropharmacology 75:525–32.
  • Rivedal, E., and E. Leithe. 2005. Connexin43 synthesis, phosphorylation, and degradation in regulation of transient inhibition of gap junction intercellular communication by the phorbol ester TPA in rat liver epithelial cells. Exp. Cell Res. 302:143–52.
  • Rivedal, E., and H. Opsahl. 2001. Role of PKC and MAP kinase in EGF- and TPA-induced connexin43 phosphorylation and inhibition of gap junction intercellular communication in rat liver epithelial cells. Carcinogenesis 22:1543–50.
  • Roemer, E., H. P. Lammerich, L. L. Conroy, and D. Weisensee. 2013. Characterization of a gap-junctional intercellular communication (GJIC) assay using cigarette smoke. Toxicol. Lett. 219:248–53.
  • Rosenberg, E., D. C. Spray, and L. M. Reid. 1992. Transcriptional and posttranscriptional control of connexin mRNAs in periportal and pericentral rat hepatocytes. Eur. J. Cell Biol. 59:21–26.
  • Ruch, R. J., R. Fransson, S. Flodstrom, L. Warngard, and J. E. Klaunig. 1990. Inhibition of hepatocyte gap junctional intercellular communication by endosulfan, chlordane and heptachlor. Carcinogenesis 11:1097–101.
  • Rusyn, I., W. A. Chiu, L. H. Lash, H. Kromhout, J. Hansen, and K. Z. Guyton. 2014. Trichloroethylene: Mechanistic, epidemiologic and other supporting evidence of carcinogenic hazard. Pharmacol. Ther. 141:55–68.
  • Saez, J. C. 1997. Intercellular gap junctional communication is required for an optimal metabolic response of the functional units of liver. Hepatology 25:775–76.
  • Sai, K., K. S. Kang, A. Hirose, R. Hasegawa, J. E. Trosko, and T. Inoue. 2001. Inhibition of apoptosis by pentachlorophenol in v-myc-transfected rat liver epithelial cells: Relation to down-regulation of gap junctional intercellular communication. Cancer Lett. 173:163–74.
  • Sai, K., J. Kanno, R. Hasegawa, J. E. Trosko, and T. Inoue. 2000. Prevention of the down-regulation of gap junctional intercellular communication by green tea in the liver of mice fed pentachlorophenol. Carcinogenesis 21:1671–76.
  • Sai, K., B. L. Upham, K. S. Kang, R. Hasegawa, T. Inoue, and J. E. Trosko. 1998. Inhibitory effect of pentachlorophenol on gap junctional intercellular communication in rat liver epithelial cells in vitro. Cancer Lett. 130:9–17.
  • Sanchez, D. I., B. Gonzalez-Fernandez, I. Crespo, B. San-Miguel, M. Alvarez, J. Gonzalez-Gallego, and M. J. Tunon. 2018. Melatonin modulates dysregulated circadian clocks in mice with diethylnitrosamine-induced hepatocellular carcinoma. J. Pineal. Res. 65:e12506.
  • Schalper, K. A., D. Carvajal-Hausdorf, and M. P. Oyarzo. 2014. Possible role of hemichannels in cancer. Front. Physiol 5:237.
  • Seifert, J. 2014. Changes in mouse liver and chicken embryo yolk sac membrane soluble proteins due to an organophosphorous insecticide (OPI) diazinon linked to several noncholinergic OPI effects in mice and chicken embryos. Pest. Biochem. Physiol. 116:74–82.
  • Setshedi, M., J. R. Wands, and S. M. Monte. 2010. Acetaldehyde adducts in alcoholic liver disease. Oxid Med. Cell Longev. 3:178–85.
  • Shimizu, K., M. Onishi, E. Sugata, Y. Sokuza, C. Mori, T. Nishikawa, K. Honoki, and T. Tsujiuchi. 2007. Disturbance of DNA methylation patterns in the early phase of hepatocarcinogenesis induced by a choline-deficient L-amino acid-defined diet in rats. Cancer Sci. 98:1318–22.
  • Shin, H. S., H. J. Lee, M. C. Pyo, D. Ryu, and K. W. Lee. 2019. Ochratoxin A-induced hepatotoxicity through phase I and phase II reactions regulated by AhR in liver cells. Toxins (Basel) 11:377.
  • Sidhu, J. S., F. Liu, and C. J. Omiecinski. 2004. Phenobarbital responsiveness as a uniquely sensitive indicator of hepatocyte differentiation status: Requirement of dexamethasone and extracellular matrix in establishing the functional integrity of cultured primary rat hepatocytes. Exp. Cell Res. 292:252–64.
  • Simeckova, P., J. Vondracek, Z. Andrysik, J. Zatloukalova, P. Krcmar, A. Kozubik, and M. Machala. 2009. The 2,2ʹ,4,4ʹ,5,5ʹ-hexachlorobiphenyl-enhanced degradation of connexin 43 involves both proteasomal and lysosomal activities. Toxicol. Sci. 107:9–18.
  • Sinyuk, M., E. E. Mulkearns-Hubert, O. Reizes, and J. Lathia. 2018. Cancer connectors: Connexins, gap junctions, and communication. Front. Oncol. 8:646.
  • Smith, J. H., J. S. Isenberg, G. Pugh Jr., L. M. Kamendulis, D. Ackley, A. W. Lington, and J. E. Klaunig. 2000. Comparative in vivo hepatic effects of Di-isononyl phthalate (DINP) and related C7-C11 dialkyl phthalates on gap junctional intercellular communication (GJIC), peroxisomal beta-oxidation (PBOX), and DNA synthesis in rat and mouse liver. Toxicol. Sci. 54:312–21.
  • Solan, J. L., and P. D. Lampe. 2018. Spatio-temporal regulation of connexin43 phosphorylation and gap junction dynamics. Biochim. Biophys. Acta Biomembr. 1860:83–90.
  • Sorgen, P. L., A. J. Trease, G. Spagnol, M. Delmar, and M. S. Nielsen. 2018. Protein-protein interactions with connexin 43: Regulation and function. Int. J. Mol. Sci. 19:1428.
  • Sovadinova, I., P. Babica, H. Boke, E. Kumar, A. Wilke, J. S. Park, J. E. Trosko, and B. L. Upham. 2015. Phosphatidylcholine specific PLC-induced dysregulation of gap junctions, a robust cellular response to environmental toxicants, and prevention by resveratrol in a rat liver cell model. PLoS ONE 10:e0124454.
  • Spannbrucker, T., N. Ale-Agha, C. Goy, N. Dyballa-Rukes, P. Jakobs, K. Jander, J. Altschmied, K. Unfried, and J. Haendeler. 2019. Induction of a senescent like phenotype and loss of gap junctional intercellular communication by carbon nanoparticle exposure of lung epithelial cells. Exp. Gerontol. 117:106–12.
  • Spray, D. C., Z. C. Ye, and B. R. Ransom. 2006. Functional connexin “hemichannels”: A critical appraisal. Glia 54:758–73.
  • Stahl, S., C. Ittrich, P. Marx-Stoelting, C. Kohle, T. Ott, A. Buchmann, and M. Schwarz. 2005. Effect of the tumor promoter phenobarbital on the pattern of global gene expression in liver of connexin32-wild-type and connexin32-deficient mice. Int. J. Cancer 115:861–69.
  • Tachikawa, M., Y. Kaneko, S. Ohtsuki, Y. Uchida, M. Watanabe, H. Ohtsuka, and T. Terasaki. 2020. Targeted proteomics-based quantitative protein atlas of pannexin and connexin subtypes in mouse and human tissues and cancer cell lnes. J. Pharm. Sci. 109:1161–68.
  • Temme, A., T. Ott, T. Haberberger, O. Traub, and K. Willecke. 2000. Acute-phase response and circadian expression of connexin26 are not altered in connexin32-deficient mouse liver. Cell Tissue Res. 300:111–17.
  • Theodorakis, N. G., and A. De Maio. 1999. Cx32 mRNA in rat liver: Effects of inflammation on poly(A) tail distribution and mRNA degradation. Am. J. Physiol. 276:1249–57.
  • Tryndyak, V., B. Borowa-Mazgaj, F. A. Beland, and I. P. Pogribny. 2019. Gene expression and cytosine DNA methylation alterations in induced pluripotent stem-cell-derived human hepatocytes treated with low doses of chemical carcinogens. Arch. Toxicol. 93:3335–44.
  • Tryndyak, V., I. Kindrat, K. Dreval, M. I. Churchwell, F. A. Beland, and I. P. Pogribny. 2018. Effect of aflatoxin B1, benzo[a]pyrene, and methapyrilene on transcriptomic and epigenetic alterations in human liver HepaRG cells. Food Chem. Toxicol. 121:214–23.
  • Upham, B. L., L. Blaha, P. Babica, J. S. Park, I. Sovadinova, C. Pudrith, A. M. Rummel, L. M. Weis, K. Sai, P. K. Tithof, et al. 2008. Tumor promoting properties of a cigarette smoke prevalent polycyclic aromatic hydrocarbon as indicated by the inhibition of gap junctional intercellular communication via phosphatidylcholine-specific phospholipase C. Cancer Sci. 99:696–705.
  • Upham, B. L., M. Guzvic, J. Scott, J. M. Carbone, L. Blaha, C. Coe, L. L. Li, A. M. Rummel, and J. E. Trosko. 2007. Inhibition of gap junctional intercellular communication and activation of mitogen-activated protein kinase by tumor-promoting organic peroxides and protection by resveratrol. Nutr. Cancer 57:38–47.
  • Vinken, M. 2016. Regulation of connexin signaling by the epigenetic machinery. Biochim. Biophys. Acta 1859:262–68.
  • Vinken, M., J. De Kock, A. G. Oliveira, G. B. Menezes, B. Cogliati, M. L. Dagli, T. Vanhaecke, and V. Rogiers. 2012. Modifications in connexin expression in liver development and cancer. Cell Commun. Adhes. 19:55–62.
  • Vinken, M., T. Doktorova, E. Decrock, L. Leybaert, T. Vanhaecke, and V. Rogiers. 2009. Gap junctional intercellular communication as a target for liver toxicity and carcinogenicity. Crit. Rev. Biochem. Mol. Biol. 44:201–22.
  • Vinken, M., T. Doktorova, H. Ellinger-Ziegelbauer, H. J. Ahr, E. Lock, P. Carmichael, E. Roggen, J. van Delft, J. Kleinjans, J. Castell, et al. 2008a. The carcinoGENOMICS project: Critical selection of model compounds for the development of omics-based in vitro carcinogenicity screening assays. Mutat. Res. 659:202–10.
  • Vinken, M., T. Henkens, E. De Rop, J. Fraczek, T. Vanhaecke, and V. Rogiers. 2008b. Biology and pathobiology of gap junctional channels in hepatocytes. Hepatology 47:1077–88.
  • Vinken, M., T. Vanhaecke, P. Papeleu, S. Snykers, T. Henkens, and V. Rogiers. 2006. Connexins and their channels in cell growth and cell death. Cell. Signal. 18:592–600.
  • VoPham, T., K. A. Bertrand, J. E. Hart, F. Laden, M. M. Brooks, J. M. Yuan, E. O. Talbott, D. Ruddell, C. H. Chang, and J. L. Weissfeld. 2017. Pesticide exposure and liver cancer: A review. Cancer Causes Control 28:177–90.
  • Wagner, E. D., S. M. McMillan, and M. J. Plewa. 2005. Cytotoxicity of organophosphorus ester (OP) insecticides and cytotoxic synergism of 2-acetoxyacetylaminofluorene (2AAAF) in Chinese hamster ovary (CHO) cells. Bull. Environ. Contam. Toxicol. 75:329–34.
  • Wang, J., X. F. Yu, O. U. Yang, N., . Q. Luo, J. Tong, T. Chen, and J. Li. 2019a. Role of DNA methylation regulation of miR-130b expression in human lung cancer using bioinformatics analysis. J. Toxicol. Environ. Health A 82:935–43.
  • Wang, W., Y. K. Shim, J. E. Michalek, E. Barber, L. M. Saleh, B. Y. Choi, C. P. Wang, N. Ketchum, R. Costello, G. E. Marti, et al. 2020. Serum microRNA profiles among dioxin exposed veterans with monoclonal gammopathy of undetermined significance. J. Toxicol. Environ. Health A 83:269–78.
  • Wang, Y., H. Zhu, and K. Kannan. 2019b. A review of biomonitoring of phthalate exposures. Toxics 7:E21.
  • Warner, K. A., M. J. Fernstrom, and R. J. Ruch. 2003. Inhibition of mouse hepatocyte gap junctional intercellular communication by phenobarbital correlates with strain-specific hepatocarcinogenesis. Toxicol. Sci. 71:190–97.
  • Warngard, L., Y. Bager, Y. Kato, K. Kenne, and U. G. Ahlborg. 1996. Mechanistical studies of the inhibition of intercellular communication by organochlorine compounds. Arch. Toxicol. Suppl. 18:149–59.
  • Wei, S., C. Cassara, X. Lin, and R. D. Veenstra. 2019. Calcium-calmodulin gating of a pH-insensitive isoform of connexin43 gap junctions. Biochem. J. 476:1137–48.
  • Wilde, E. C., K. E. Chapman, L. M. Stannard, A. L. Seager, K. Brusehafer, U. K. Shah, J. A. Tonkin, M. R. Brown, J. R. Verma, A. T. Doherty, et al. 2018. A novel, integrated in vitro carcinogenicity test to identify genotoxic and non-genotoxic carcinogens using human lymphoblastoid cells. Arch. Toxicol. 92:935–51.
  • Willebrords, J., B. Cogliati, I. V. A. Pereira, T. C. da Silva, S. Crespo Yanguas, M. Maes, V. M. Govoni, A. Lima, D. A. Felisbino, E. Decrock, et al. 2017a. Inhibition of connexin hemichannels alleviates non-alcoholic steatohepatitis in mice. Sci. Rep 7:8268.
  • Willebrords, J., M. Maes, S. C. Yanguas, and M. Vinken. 2017b. Inhibitors of connexin and pannexin channels as potential therapeutics. Pharmacol. Ther. 180:144–60.
  • Wu, J. G., L. Lin, T. G. Luan, Y. S. C. Gilbert, and C. Lan. 2007. Effects of organophosphorus pesticides and their ozonation byproducts on gap junctional intercellular communication in rat liver cell line. Food Chem. Toxicol. 45:2057–63.
  • Xiang, Y., Q. Wang, Y. Guo, H. Ge, Y. Fu, X. Wang, and L. Tao. 2019. Cx32 exerts anti-apoptotic and pro-tumor effects via the epidermal growth factor receptor pathway in hepatocellular carcinoma. J. Exp. Clin. Cancer Res. 38:145.
  • Yamasaki, H. 1995. Non-genotoxic mechanisms of carcinogenesis: Studies of cell transformation and gap junctional intercellular communication. Toxicol. Lett. 77:55–61.
  • Yamasaki, H., J. Ashby, M. Bignami, W. Jongen, K. Linnainmaa, R. F. Newbold, G. Nguyen-Ba, S. Parodi, E. Rivedal, D. Schiffmann, et al. 1996. Nongenotoxic carcinogens: Development of detection methods based on mechanisms: A European project. Mutat. Res. 353:47–63.
  • Yamasaki, H., V. Krutovskikh, M. Mesnil, T. Tanaka, M. L. Zaidan-Dagli, and Y. Omori. 1999. Role of connexin (gap junction) genes in cell growth control and carcinogenesis. C. R. Acad. Sci. III 322:151–59.
  • Yamashita, Y., M. Shimada, N. Harimoto, S. Tanaka, K. Shirabe, H. Ijima, K. Nakazawa, J. Fukuda, K. Funatsu, and Y. Maehara. 2004. cDNA microarray analysis in hepatocyte differentiation in Huh 7 cells. Cell Transplant 13:793–99.
  • Yang, L., C. Dong, L. Tian, X. Ji, L. Yang, and L. Li. 2019. Gadolinium chloride restores the function of the gap junctional intercellular communication between hepatocytes in a liver injury. Int. J. Mol. Sci. 20:3748.
  • Yang, Y., J. Zhu, N. Zhang, Y. Zhao, W. Y. Li, F. Y. Zhao, Y. R. Ou, S. K. Qin, and Q. Wu. 2016. Impaired gap junctions in human hepatocellular carcinoma limit intrinsic oxaliplatin chemosensitivity: A key role of connexin 26. Int. J. Oncol. 48:703–13.
  • Ye, Y. X., D. Bombick, K. Hirst, G. X. Zhang, C. C. Chang, J. E. Trosko, and T. Akera. 1990. The modulation of gap junctional communication by gossypol in various mammalian cell lines in vitro. Fundam. Appl. Toxicol. 14:817–32.
  • Yeh, S. P., T. G. Sung, C. C. Chang, W. Cheng, and C. M. Kuo. 2005. Effects of an organophosphorus insecticide, trichlorfon, on hematological parameters of the giant freshwater prawn, Macrobrachium rosenbergii (de Man). Aquaculture 243:383–92.
  • Yu, M., Q. Zou, X. Wu, G. Han, and X. Tong. 2017. Connexin 32 affects doxorubicin resistance in hepatocellular carcinoma cells mediated by Src/FAK signaling pathway. Biomed. Pharmacother. 95:1844–52.
  • Zhang, D., M. Kaneda, K. Nakahama, S. Arii, and I. Morita. 2007. Connexin 43 expression promotes malignancy of HuH7 hepatocellular carcinoma cells via the inhibition of cell-cell communication. Cancer Lett. 252:208–15.
  • Zhao, B., W. Zhao, Y. Wang, Y. Xu, J. Xu, K. Tang, S. Zhang, Z. Yin, Q. Wu, and X. Wang. 2015. Connexin32 regulates hepatoma cell metastasis and proliferation via the p53 and Akt pathways. Oncotarget 6:10116–33.
  • Zou, H., X. Liu, T. Han, D. Hu, Y. Wang, Y. Yuan, J. Gu, J. Bian, J. Zhu, and Z. P. Liu. 2015b. Salidroside protects against cadmium-induced hepatotoxicity in rats via GJIC and MAPK pathways. PLoS ONE 10:e0129788.
  • Zou, H., L. Zhuo, T. Han, D. Hu, X. Yang, Y. Wang, Y. Yuan, J. Gu, J. Bian, X. Liu, et al. 2015a. Autophagy and gap junctional intercellular communication inhibition are involved in cadmium-induced apoptosis in rat liver cells. Biochem. Biophys. Res. Commun. 459:713–19.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.