Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 1744 word(s) 1744 2022-01-03 08:21:40 |
2 References are added. + 2121 word(s) 3865 2022-01-11 11:51:50 | |
3 format correct -2035 word(s) 1830 2022-01-13 07:04:54 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Li, X. Inorganic SOD1 Inhibitors with Anti-Cancer Prospects. Encyclopedia. Available online: https://encyclopedia.pub/entry/18058 (accessed on 05 July 2024).
Li X. Inorganic SOD1 Inhibitors with Anti-Cancer Prospects. Encyclopedia. Available at: https://encyclopedia.pub/entry/18058. Accessed July 05, 2024.
Li, Xiang. "Inorganic SOD1 Inhibitors with Anti-Cancer Prospects" Encyclopedia, https://encyclopedia.pub/entry/18058 (accessed July 05, 2024).
Li, X. (2022, January 11). Inorganic SOD1 Inhibitors with Anti-Cancer Prospects. In Encyclopedia. https://encyclopedia.pub/entry/18058
Li, Xiang. "Inorganic SOD1 Inhibitors with Anti-Cancer Prospects." Encyclopedia. Web. 11 January, 2022.
Inorganic SOD1 Inhibitors with Anti-Cancer Prospects
Edit

For eukaryotic cells, reactive oxygen species (ROS) encompass a group of molecules derived from oxygen. Due to the well-established role of ROS in cell signaling, cancer cells always have higher levels of endogenous ROS to enhance rapid cell growth and proliferation through the mitogen-activated protein kinase (MAPK)/extracellular-regulated kinase 1/2 (ERK1/2), phosphoinositide-3-kinase (PI3K)/Akt, nuclear factor-κB (NF-κB), and hypoxia-sensitive α (HIF1α) pathways. Evaluated ROS have frequently been observed in various cancers, which activate multiple pro-tumourigenic signaling, and induce survival and proliferation of cancer cells. Hydrogen peroxide and superoxide anion are the most important redox signaling agents in cancer cells, whose homeostasis is maintained by dozens of growth factors, cytokines and antioxidant enzymes. Therefore, antioxidant enzymes, especially Cu/Zn superoxide dismutase (SOD1), tend to have higher activities to maintain the homeostasis of ROS in cancer cells. We can inhibit the activity of SOD1 using copper chelators to kill cancer cells. 

anti-cancer reactive oxygen species chelators antioxidant enzymes SOD1 mitochondria inhibitors

1. Introduction

For eukaryotic cells, reactive oxygen species (ROS) encompass a group of molecules derived from oxygen, such as hydrogen peroxide (H2O2), superoxide anion (O2•−), organic hydroperoxides (ROOH), singlet molecular oxygen (1O2), hydroxyl radical (•OH), alkoxyl radical (•OR), and peroxyl radical (•OOR) [1][2][3]. ROS are mainly formed by reduction–oxidation reactions or by electronic excitation (Figure 1A) [1][2][3] and have evolved as regulators of multiple signaling pathways [4][5][6][7][8]. Two species, H2O2 and O2•−, are the most important redox signaling agents in the cells [4][5][6][7][8]. H2O2 is the major ROS in organisms, with its concentration always maintained within 1~100 nM under normal conditions [8]. For O2•−, the concentration is also maintained at about 0.01 nM, much lower than that of H2O2 [8].
Figure 1. ROS in cancer cells. (A) Generation and chemical structures of ROS. (B) Brief metabolic process and signal regulation of intracellular ROS. (C) Balancing ROS generation and scavenging in cancer cells to remain in the tumorigenic range.
Dozens of growth factors, cytokines, and antioxidant enzymes control the homeostasis of intracellular H2O2 and O2•− (Figure 1B) [9]. O2•− is prominently generated by the mitochondrial electron transport chain (Mito-ETC), NADPH oxidase (NOX) complex, and endoplasmic reticulum (ER) system, and is rapidly converted to H2O2 by superoxide dismutases (SODs) [10][11][12]. Subsequently, H2O2 is mainly detoxified to H2O by catalase (CAT), glutathione peroxidase (GPX), and peroxiredoxin (Prx) [12]. It is worth mentioning that •OH formed by metal-catalyzed Fenton reaction is the most reactive ROS; it can oxidize biological macromolecules indiscriminately, such as DNA, proteins, and lipids [12]. Therefore, maintaining the homeostasis of intracellular ROS is essential for cell growth, proliferation, and survival [6][9][10].
Due to the well-established role of ROS in cell signaling, cancer cells always have higher levels of endogenous ROS to enhance rapid cell growth and proliferation through the mitogen-activated protein kinase (MAPK)/extracellular-regulated kinase 1/2 (ERK1/2), phosphoinositide-3-kinase (PI3K)/Akt, nuclear factor-κB (NF-κB), and hypoxia-sensitive α (HIF1α) pathways [13][14][15][16][17][18]. Indeed, higher levels of ROS have already been observed in various cancer cells [11][19]. If the intracellular ROS levels increase dramatically to toxic concentrations, oxidative stress will cause irreversible damage and may eventually lead to the death of cancer cells [11][20]. To maintain the elevated mitogenic signaling without incurring substantial oxidative damage by a proper balance of ROS, the antioxidant enzymes in cancer cells, such as Cu/Zn superoxide dismutase (SOD1), GPX, and Prx, should harbor higher levels of activity (Figure 1C) [21][22][23][24].
Elevated levels of ROS are always involved in the initiation and progression of cancer. Hence, intervening in the homeostasis of ROS in cancer cells is an effective anti-cancer strategy [25][26]. So far, a variety of chelators or metal complexes based on the regulation of ROS have been reported as anti-cancer agents [27][28][29][30][31][32][33][34]. For those antioxidant enzymes where the active center is a metal ion, chelators can be used to competitively bind to the metal ion and thus inhibit the enzymatic activity to achieve the regulation of intracellular ROS, including SOD1 inhibitors tetrathiomolybdate (ATN-224) and LD100 [32][33]. On the other hand, several metal complexes can regulate the ROS levels in cancer cells through other mechanisms to achieve anti-cancer purposes, such as TxrR inhibition and mitochondrial dysfunction [34][35][36]. Regulating the relative level of ROS in cancer cells through the mechanism of metal coordination has become an important branch with broad prospects in the field of cancer therapy.

2. Inorganic SOD1 Inhibitors with Anti-Cancer Prospects

In mammals, the main biological function of SODs is to catalyze the dismutation of O2•− into H2O2 and O2 [37][38]. Cu/Zn superoxide dismutase (SOD1), the major SOD, mainly exists in the formation of homodimers in cells and is widely distributed in the nucleus, the cytoplasm, and the intermembrane space (IMS) of mitochondria [39]. Next, Mn superoxide dismutase (SOD2) exclusively exists in the mitochondrial matrix [40]. An extracellular form of SOD (EC-SOD), also a Cu/Zn-containing SOD, is tetrameric and exists in most mammals [40]. Besides this, SOD1 also regulates multiple redox signals to control growth and metabolic pathways, such as glucose metabolism and transcription [5][41][42][43]. Therefore, SODs, especially SOD1, are the first firewall to resist oxidative stress.
Recently, emerging evidence from researchers has indicated that SOD1 is usually overexpressed in cancer cells; its activity is essential to maintain higher ROS levels under the critical threshold during aberrant energy metabolism of cancer progression [39]. For example, SOD1 accumulations were observed not only in the cytoplasm but also in the nucleus of human primary breast and mammary cancers [44]. Besides this, prostate cancer cells (DU145) also have higher levels of activity and expression of SOD1, compared with normal prostate cells (RWPE-1) [5]. In vitro studies also showed that the fast growth of non-small cell lung cancer (NSCLC) and leukemia depends on the high activity of SOD1, which controls the oncogenic KRAS and EGFR pathways [45][46], as well as other cancer cells and xenograft tumors [47]. In general, SOD1 is recognized as a promising anti-cancer target, and several small-molecule targeting drugs for SOD1 have already entered the preclinical and clinical development stages [48].
Since the activity of SOD1 mainly comes from the copper ion in the active center, a vast majority of SOD1 inhibitors are competitive chelators of copper ions. In 1975, Heikkila et al. found that diethyldithiocarbamate (DDC) can competitively bind to copper ions (Figure 2), thereby inhibiting SOD1 activity at a millimolar level [49]. After being inhibited by DDC, SOD1 cannot restore enzyme activity through dialysis, but adding CuSO4 during dialysis restores SOD1 activity [49]. In 1979, Misra systematically explored the mechanism by which DDC inhibits SOD1 activity [50]. In Phase I, one DDC molecule first coordinates with the copper(II) center in SOD1, with retention of activity. In Phase II, a second DDC displaces the copper(II) center, with a loss of activity. The shortcomings of DDC as a SOD1 inhibitor are mainly reflected in its high working concentration and poor specificity, such as its interference with the activity of cytochrome c oxidase [51][52]. Nevertheless, DDC still has a wide range of anti-cancer applications, and DDC effectively inhibits SOD1 activity to kill cancer cells [53][54][55].
Figure 2. Summary of SOD1 inhibitors based on copper chelation. The specific activity inhibition of SOD1 selectively kills cancer cells by regulating the intracellular ROS signaling network.
In 2005, Ding et al. found that clioquinol (5-chloro-7-iodo-8-hydroxyquinoline, CQ), a metal chelator of copper/zinc/iron, is another SOD1 inhibitor (Figure 2), because the copper and zinc ions in the active sites of SOD1 are coordinated by CQ [56]. Structural characterization of the zinc(II) and copper(II) complexes with CQ indicated that the stoichiometry of ligand to metal is 2:1 [57]. Therefore, CQ can effectively inhibit SOD1 at micromolar concentrations (IC50: 6.7~43.1 μM) and induce the death of a variety of cancer cells through the caspase-3-mediated apoptosis pathway [56]. It cannot be ignored that CQ has the risk of destroying copper homeostasis during its inhibition of SOD1 activity in cells [58].
Tetrathiomolybdate is an orally available copper chelator that has been shown to have efficacy as an anti-angiogenic and anti-tumor agent in multiple cancers [59][60]ATN-224 is the second-generation choline salt of tetrathiomolybdate with improved performance (Figure 2) and is being evaluated in several phase II trials in cancer patients [61]. Doñate et al. found that ATN-224 can selectively bind copper with high affinity, and SOD1 is the main target for the anti-angiogenic activity of this chelator [32][62]ATN-224 also inhibits intracellular SOD1 activity at micromolar concentrations (IC50: 1.4~185 μM), but has specificity for copper binding, all of which makes it one of the most popular SOD1 inhibitors [60][62]. Every sulfur atom in tetrathiomolybdate can coordinate with copper and may then form metal clusters with copper enzymes, thereby inhibiting the activity of copper proteins, such as SOD1, cytochrome c oxidase, and ceruloplasmin [63][64]. Therefore, ATN-224 may also interfere with intracellular copper homeostasis or inhibit other copper enzymes. In cancer treatment, ATN-224-mediated SOD1 inhibition led to the downregulation of PDGF and increase of O2•−, prevented the formation of high levels of H2O2, and protected protein tyrosine phosphatases from oxidation by H2O2 [60]. Therefore, SOD1 inhibition by ATN-224 results in the down-regulation of multiple signaling pathways for cancer cell function, such as ERK1/2 and anti-apoptotic factor Mcl1 [48][60].
Considering that the known SOD1 inhibitors have various defects, such as low efficiency, weak specificity, and interference with the homeostasis of metal ions, we designed a next-generation SOD1 inhibitor (LD100) based on copper coordination chemistry and the catalytic cycle in the active site (Figure 2) [33]LD100 was designed through the combination of thiosemicarbazone and phenol derivatives, because thiosemicarbazone contains a copper chelating moiety, -C(SH)-NH-, and the phenolic hydroxyl can further facilitate the copper coordination. Besides this, LD100 also contains a fluorescent group chromone, which not only can be used to track the entry of LD100 into cells, but also enables LD100 to better occupy the substrate channel of SOD1. Therefore, LD100 has a strong binding ability to copper ions in solution and can effectively inhibit the activity of SOD1 in vitro and in vivo (IC50 of LD100 to SOD1 in HeLa cells: 0.18 μM) [33]. Through specific inhibition of SOD1 activity, LD100 can efficiently up-regulate the intracellular concentration of O2•−, down-regulate the concentration of H2O2, down-regulate the phosphorylation of ERK1/2, and finally induce the apoptosis of cancer cells [33]. In summary, LD100 may be the most effective SOD1 inhibitor so far and has application prospects for cancer treatment. Using this inhibitor, we also systematically explored the mechanism of how SOD1 activity inhibition selectively kills cancer cells [5]. The rapid growth and proliferation of cancer cells always depend on higher SOD1 activity, so cancer cells are more sensitive to SOD1 inhibition. During SOD1 inhibition in cancer cells, LD100 could repress the ERK, PI3K-Akt, and NF-κB pathways; arrest the cell cycle; and induce mitochondria-dependent apoptosis [5].

3. Prespective

SOD1 is indeed a recognized target for cancer treatment. At present, a variety of chelators have been used for SOD1 inhibition. LD100 may be the most effective inhibitor designed through coordination chemistry. However, the use of inorganic strategies to develop anti-cancer drugs based on SOD1 inhibition still requires further efforts. First, we need to solve the problem of compatibility between targeted and clinical deliveries. On the other hand, we also should reduce the side effects of chelating agents while ensuring the efficiency of SOD1 inhibition. The summary of SOD1 metal-chelating inhibition can provide a reference for the design of SOD1 inhibitors with anti-cancer effects in the future.

References

  1. Waszczak, C.; Carmody, M.; Kangasjärvi, J. Reactive oxygen species in plant signaling. Annu. Rev. Plant Biol. 2018, 69, 209–236.
  2. Weinberg, F.; Ramnath, N.; Nagrath, D. Reactive oxygen species in the tumor microenvironment: An overview. Cancers 2019, 11, 1191.
  3. Yang, B.; Chen, Y.; Shi, J. Reactive oxygen species (ROS)-based nanomedicine. Chem. Rev. 2019, 119, 4881–4985.
  4. Marcec, M.J.; Gilroy, S.; Poovaiah, B.W.; Tanaka, K. Mutual interplay of Ca2+ and ROS signaling in plant immune response. Plant Sci. 2019, 283, 343–354.
  5. Li, X.; Chen, Y.; Zhao, J.; Shi, J.; Wang, M.; Qiu, S.; Hu, Y.; Xu, Y.; Cui, Y.; Liu, C.; et al. The specific inhibition of SOD1 selectively promotes apoptosis of cancer cells via regulation of the ROS signaling network. Oxidative Med. Cell. Longev. 2019, 2019, 9706792.
  6. Schieber, M.; Chandel, N.S. ROS function in redox signaling and oxidative stress. Curr. Biol. 2014, 24, R453–R462.
  7. Milkovic, L.; Cipak Gasparovic, A.; Cindric, M.; Mouthuy, P.A.; Zarkovic, N. Short overview of ROS as cell function regulators and their implications in therapy concepts. Cells 2019, 8, 793.
  8. Sies, H.; Jones, D.P. Reactive oxygen species (ROS) as pleiotropic physiological signalling agents. Nat. Rev. Mol. Cell Biol. 2020, 21, 363–383.
  9. D’Autréaux, B.; Toledano, M.B. ROS as signalling molecules: Mechanisms that generate specificity in ROS homeostasis. Nat. Rev. Mol. Cell Biol. 2007, 8, 813–824.
  10. Reczek, C.R.; Chandel, N.S. ROS-dependent signal transduction. Curr. Opin. Cell Biol. 2015, 33, 8–13.
  11. Moloney, J.N.; Cotter, T.G. ROS signalling in the biology of cancer. Semin. Cell Dev. Biol. 2018, 80, 50–64.
  12. Wang, Y.; Branicky, R.; Noë, A.; Hekimi, S. Superoxide dismutases: Dual roles in controlling ROS damage and regulating ROS signaling. J. Cell Biol. 2018, 217, 1915–1928.
  13. Su, X.; Shen, Z.; Yang, Q.; Sui, F.; Pu, J.; Ma, J.; Ma, S.; Yao, D.; Ji, M.; Hou, P. Vitamin C kills thyroid cancer cells through ROS-dependent inhibition of MAPK/ERK and PI3K/AKT pathways via distinct mechanisms. Theranostics 2019, 9, 4461.
  14. Li, Y.; Liang, R.; Zhang, X.; Wang, J.; Shan, C.; Liu, S.; Li, L.; Zhang, S. Copper chaperone for superoxide dismutase promotes breast cancer cell proliferation and migration via ROS-mediated MAPK/ERK signaling. Front. Pharmacol. 2019, 10, 356–367.
  15. Steelman, L.S.; Abrams, S.L.; Whelan, J.; Bertrand, F.E.; Ludwig, D.E.; Bäsecke, J.; Libra, M.; Stivala, F.; Milella, M.; Tafuri, A.; et al. Contributions of the Raf/MEK/ERK, PI3K/PTEN/Akt/mTOR and Jak/STAT pathways to leukemia. Leukemia 2008, 22, 686–707.
  16. Yeo, D.; Hwang, S.J.; Kim, W.J.; Youn, H.J.; Lee, H.J. The aqueous extract from Artemisia capillaris inhibits acute gastric mucosal injury by inhibition of ROS and NF-κB. Biomed. Pharmacother. 2018, 99, 681–687.
  17. Park, S.A.; Na, H.K.; Kim, E.H.; Cha, Y.N.; Surh, Y.J. 4-Hydroxyestradiol induces anchorage-independent growth of human mammary epithelial cells via activation of IκB kinase: Potential role of reactive oxygen species. Cancer Res. 2009, 69, 2416–2424.
  18. Castelli, S.; Ciccarone, F.; Tavian, D.; Ciriolo, M.R. ROS-dependent HIF1α activation under forced lipid catabolism entails glycolysis and mitophagy as mediators of higher proliferation rate in cervical cancer cells. J. Exp. Clin. Cancer Res. 2021, 40, 1–18.
  19. Szatrowski, T.P.; Nathan, C.F. Production of large amounts of hydrogen peroxide by human tumor cells. Cancer Res. 1991, 51, 794–798.
  20. Perillo, B.; Di Donato, M.; Pezone, A.; Di Zazzo, E.; Giovannelli, P.; Galasso, G.; Castoria, G.; Migliaccio, A. ROS in cancer therapy: The bright side of the moon. Exp. Mol. Med. 2020, 52, 192–203.
  21. Hu, Y.; Rosen, D.G.; Zhou, Y.; Feng, L.; Yang, G.; Liu, J.; Huang, P. Mitochondrial manganese-superoxide dismutase expression in ovarian cancer: Role in cell proliferation and response to oxidative stress. J. Biol. Chem. 2005, 280, 39485–39492.
  22. Saydam, N.; Kirb, A.; Demir, Ö.; Hazan, E.; Oto, Ö.; Saydam, O.; Güner, G. Determination of glutathione, glutathione reductase, glutathione peroxidase and glutathione S-transferase levels in human lung cancer tissues. Cancer Lett. 1997, 119, 13–19.
  23. Murawaki, Y.; Tsuchiya, H.; Kanbe, T.; Harada, K.; Yashima, K.; Nozaka, K.; Tanida, O.; Kohno, M.; Mukoyama, T.; Nishimuki, E.; et al. Aberrant expression of selenoproteins in the progression of colorectal cancer. Cancer Lett. 2008, 259, 218–230.
  24. Oberley, T.D.; Oberley, L.W. Antioxidant enzyme levels in cancer. Histol. Histopathol. 1997, 12, 525–535.
  25. De Sá Junior, P.L.; Câmara, D.A.D.; Porcacchia, A.S.; Fonseca, P.M.M.; Jorge, S.D.; Araldi, R.P.; Ferreira, A.K. The roles of ROS in cancer heterogeneity and therapy. Oxidative Med. Cell. Longev. 2017, 2017, 2467940.
  26. Chio, I.I.C.; Tuveson, D.A. ROS in cancer: The burning question. Trends Mol. Med. 2017, 23, 411–429.
  27. Zehra, S.; Cirilli, I.; Silvestri, S.; Gómez-Ruiz, S.; Tabassum, S.; Arjmand, F. Structure elucidation, in vitro binding studies and ROS-dependent anti-cancer activity of Cu (II) and Zn (II) phthaloylglycinate (phen) complexes against MDA-MB-231 cells. Metallomics 2021, 13, mfab064.
  28. Guo, W.; Ye, S.; Cao, N.; Huang, J.; Gao, J.; Chen, Q. ROS-mediated autophagy was involved in cancer cell death induced by novel copper (II) complex. Exp. Toxicol. Pathol. 2010, 62, 577–582.
  29. Liu, J.; Guo, W.; Li, J.; Li, X.; Geng, J.; Chen, Q.; Gao, J. Tumor-targeting novel manganese complex induces ROS-mediated apoptotic and autophagic cancer cell death. Int. J. Mol. Med. 2015, 35, 607–616.
  30. Marloye, M.; Berger, G.; Gelbcke, M.; Dufrasne, F. A survey of the mechanisms of action of anticancer transition metal complexes. Future Med. Chem. 2016, 8, 2263–2286.
  31. Sîrbu, A.; Palamarciuc, O.; Babak, M.V.; Lim, J.; Ohui, K.; Enyedy, E.A.; Shova, S.; Darvasiová, D.; Rapta, P.; Ang, W.H.; et al. Copper (II) thiosemicarbazone complexes induce marked ROS accumulation and promote nrf2-mediated antioxidant response in highly resistant breast cancer cells. Dalton Trans. 2017, 46, 3833–3847.
  32. Donate, F.; Juarez, J.C.; Burnett, M.E.; Manuia, M.M.; Guan, X.; Shaw, D.E.; Smith, E.L.P.; Timucin, C.; Braunstein, M.J.; Batuman, O.A.; et al. Identification of biomarkers for the antiangiogenic and antitumour activity of the superoxide dismutase 1 (SOD1) inhibitor tetrathiomolybdate (ATN-224). Br. J. Cancer 2008, 98, 776–783.
  33. Dong, X.; Zhang, Z.; Zhao, J.; Lei, J.; Chen, Y.; Li, X.; Chen, H.; Tian, J.; Zhang, D.; Liu, C.; et al. The rational design of specific SOD1 inhibitors via copper coordination and their application in ROS signaling research. Chem. Sci. 2016, 7, 6251–6262.
  34. Kalaivani, P.; Saranya, S.; Poornima, P.; Prabhakaran, R.; Dallemer, F.; Padma, V.V.; Natarajan, K. Biological evaluation of new nickel (II) metallates: Synthesis, DNA/protein binding and mitochondrial mediated apoptosis in human lung cancer cells (A549) via ROS hypergeneration and depletion of cellular antioxidant pool. Eur. J. Med. Chem. 2014, 82, 584–599.
  35. Zhang, P.; Sadler, P.J. Redox-active metal complexes for anticancer therapy. Eur. J. Inorg. Chem. 2017, 2017, 1541–1548.
  36. Imberti, C.; Zhang, P.; Huang, H.; Sadler, P.J. New designs for phototherapeutic transition metal complexes. Angew. Chem. Int. Ed. 2020, 59, 61–73.
  37. Borgstahl, G.E.O.; Oberley-Deegan, R.E. Superoxide dismutases (SODs) and SOD mimetics. Antioxidants 2018, 7, 156.
  38. Robinett, N.G.; Peterson, R.L.; Culotta, V.C. Eukaryotic copper-only superoxide dismutases (SODs): A new class of SOD enzymes and SOD-like protein domains. J. Biol. Chem. 2018, 293, 4636–4643.
  39. Papa, L.; Manfredi, G.; Germain, D. SOD1, an unexpected novel target for cancer therapy. Genes Cancer 2014, 5, 15–21.
  40. Wang, X.; Zhang, H.; Sapio, R.; Yang, J.; Wong, J.; Zhang, X.; Guo, J.Y.; Pine, S.; Remmen, H.V.; Li, H.; et al. SOD1 regulates ribosome biogenesis in KRAS mutant non-small cell lung cancer. Nat. Commun. 2021, 12, 1–15.
  41. Tsang, C.K.; Liu, Y.; Thomas, J.; Zhang, Y.; Zheng, X.F.S. Superoxide dismutase 1 acts as a nuclear transcription factor to regulate oxidative stress resistance. Nat. Commun. 2014, 5, 1–11.
  42. Li, X.; Qiu, S.; Shi, J.; Wang, S.; Wang, M.; Xu, Y.; Nie, Z.; Liu, C.; Liu, C. A new function of copper zinc superoxide dismutase: As a regulatory DNA-binding protein in gene expression in response to intracellular hydrogen peroxide. Nucleic Acids Res. 2019, 10, 5074–5085.
  43. Reddi, A.R.; Culotta, V.C. SOD1 integrates signals from oxygen and glucose to repress respiration. Cell 2013, 152, 224–235.
  44. Papa, L.; Hahn, M.; Marsh, E.L.; Evans, B.S.; Germain, D. SOD2 to SOD1 switch in breast cancer. J. Biol. Chem. 2014, 289, 5412–5416.
  45. Glasauer, A.; Sena, L.A.; Diebold, L.P.; Mazar, A.P.; Chandel, N.S. Targeting SOD1 reduces experimental non–small-cell lung cancer. J. Clin. Investig. 2014, 124, 117–128.
  46. Somwar, R.; Erdjument-Bromage, H.; Larsson, E.; Shum, D.; Lockwood, W.W. Superoxide dismutase 1 (SOD1) is a target for a small molecule identified in a screen for inhibitors of the growth of lung adenocarcinoma cell lines. Proc. Natl. Acad. Sci. USA 2011, 108, 16375–16380.
  47. Gomez, M.L.; Shah, N.; Kenny, T.C.; Jenkins Jr, E.C.; Germain, D. SOD1 is essential for oncogene-driven mammary tumor formation but dispensable for normal development and proliferation. Oncogene 2019, 38, 5751–5765.
  48. Che, M.; Wang, R.; Li, X.; Wang, H.Y.; Zheng, X.F.S. Expanding roles of superoxide dismutases in cell regulation and cancer. Drug Discov. Today 2016, 21, 143–149.
  49. Heikkila, R.E.; Cabbat, F.S.; Cohen, G. In vivo inhibition of superoxide dismutase in mice by diethyldithiocarbamate. J. Biol. Chem. 1976, 251, 2182–2185.
  50. Misra, H.P. Reaction of copper-zinc superoxide dismutase with diethyldithiocarbamate. J. Biol. Chem. 1979, 254, 11623–11628.
  51. Singh, N.; Savanur, M.A.; Srivastava, S.; D’Silva, P.; Mugesh, G. A manganese oxide nanozyme prevents the oxidative damage of biomolecules without affecting the endogenous antioxidant system. Nanoscale 2019, 11, 3855–3863.
  52. Griffiths, D.E.; Wharton, D.C. Studies of the electron transport system XXXV. Purification and properties of cytochrome oxidase. J. Biol. Chem. 1961, 236, 1850–1856.
  53. Skrott, Z.; Cvek, B. Diethyldithiocarbamate complex with copper: The mechanism of action in cancer cells. Mini Rev. Med. Chem. 2012, 12, 1184–1192.
  54. Feuser, P.E.; Cordeiro, A.P.; Silveira, G.B.; Borges Corrêa, M.E.A.; Silveira, P.C.L.; Sayer, C.; Hermes de Araújo, P.H.; Machado-de-Ávila, R.A.; Dal Bó, A.G. Co-encapsulation of sodium diethyldithiocarbamate (DETC) and zinc phthalocyanine (ZnPc) in liposomes promotes increases phototoxic activity against (MDA-MB 231) human breast cancer cells. Colloids Surf. B Biointerfaces 2021, 197, 111434.
  55. Cho, H.Y.; Mavi, A.; Chueng, S.T.D.; Borges Corrêa, M.E.A.; Silveira, P.C.L.; Sayer, C.; Araújo, P.H.H.; Machado-de-Ávila, R.A.; Dal Bó, A.G. Tumor homing reactive oxygen species nanoparticle for enhanced cancer therapy. ACS Appl. Mater. Interfaces 2019, 11, 23909–23918.
  56. Ding, W.Q.; Liu, B.; Vaught, J.L.; Yamauchi, H.; Lind, S.E. Anticancer activity of the antibiotic clioquinol. Cancer Res. 2005, 65, 3389–3395.
  57. Di Vaira, M.; Bazzicalupi, C.; Orioli, P.; Messori, L.; Bruni, B.; Zatta, P. Clioquinol, a drug for Alzheimer’s disease specifically interfering with brain metal metabolism: Structural characterization of its zinc (II) and copper (II) complexes. Inorg. Chem. 2004, 43, 3795–3797.
  58. Katsuyama, M.; Kimura, E.; Ibi, M.; Iwata, K.; Matsumoto, M.; Asaoka, N.; Yabe-Nishimura, C. Clioquinol inhibits dopamine-β-hydroxylase secretion and noradrenaline synthesis by affecting the redox status of ATOX1 and copper transport in human neuroblastoma SH-SY5Y cells. Arch. Toxicol. 2021, 95, 135–148.
  59. Brewer, G.J.; Dick, R.D.; Grover, D.K.; LeClaire, V.; Tseng, M.; Wicha, M.; Pienta, K.; Redman, B.G.; Jahan, T.; Sondak, V.K.; et al. Treatment of metastatic cancer with tetrathiomolybdate, an anticopper, antiangiogenic agent: Phase I study. Clin. Cancer Res. 2000, 6, 1–10.
  60. Juarez, J.C.; Manuia, M.; Burnett, M.E.; Betancourt, O.; Boivin, B.; Shaw, D.E.; Tonks, N.K.; Mazar, A.P.; Donate, F. Superoxide dismutase 1 (SOD1) is essential for H2O2-mediated oxidation and inactivation of phosphatases in growth factor signaling. Proc. Natl. Acad. Sci. USA 2008, 105, 7147–7152.
  61. Lin, J.; Zahurak, M.; Beer, T.M.; Ryan, C.J.; Wilding, G.; Mathew, P.; Morris, M.; Callahan, J.A.; Gordon, G.; Reich, S.D.; et al. A non-comparative randomized phase II study of 2 doses of ATN-224, a copper/zinc superoxide dismutase inhibitor, in patients with biochemically recurrent hormone-naïve prostate cancer. In Urologic Oncology: Seminars and Original Investigations; Elsevier: Amsterdam, The Netherlands, 2013; Volume 5, pp. 581–588.
  62. Juarez, J.C.; Betancourt, O.; Pirie-Shepherd, S.R.; Guan, X.; Price, M.L.; Shaw, D.E.; Mazar, A.P.; Doñate, F. Copper binding by tetrathiomolybdate attenuates angiogenesis and tumor cell proliferation through the inhibition of superoxide dismutase. Clin. Cancer Res. 2006, 12, 4974–4982.
  63. Maiti, B.K.; Moura, J.J. Diverse biological roles of the tetrathiomolybdate anion. Coord. Chem. Rev. 2021, 429, 213635.
  64. Alvarez, H.M.; Xue, Y.; Robinson, C.D.; Canalizo-Hernandez, M.A.; Marvin, R.G.; Kelly, R.A.; Mondragon, A.; Penner-Hahn, J.E.; O’Halloran, T.V. Tetrathiomolybdate inhibits copper trafficking proteins through metal cluster formation. Science 2010, 327, 331–334.
More
Information
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 395
Entry Collection: Biopharmaceuticals Technology
Revisions: 3 times (View History)
Update Date: 13 Jan 2022
1000/1000
Video Production Service