Stem cells and the impact of ROS signaling

Carolina L Bigarella, Raymond Liang, Saghi Ghaffari, Carolina L Bigarella, Raymond Liang, Saghi Ghaffari

Abstract

An appropriate balance between self-renewal and differentiation is crucial for stem cell function during both early development and tissue homeostasis throughout life. Recent evidence from both pluripotent embryonic and adult stem cell studies suggests that this balance is partly regulated by reactive oxygen species (ROS), which, in synchrony with metabolism, mediate the cellular redox state. In this Primer, we summarize what ROS are and how they are generated in the cell, as well as their downstream molecular targets. We then review recent findings that provide molecular insights into how ROS signaling can influence stem cell homeostasis and lineage commitment, and discuss the implications of this for reprogramming and stem cell ageing. We conclude that ROS signaling is an emerging key regulator of multiple stem cell populations.

Keywords: Embryonic stem cells; Hematopoietic stem cells; Metabolism; Mitochondria; ROS.

© 2014. Published by The Company of Biologists Ltd.

Figures

Fig. 1.
Fig. 1.
ROS generation and scavenging. (A) Reactive oxygen species (ROS) include superoxide (O2.−), hydrogen peroxide (H2O2) and the highly reactive hydroxyl radical (OH.) (shown in red). O2.− can be generated from complexes I and III (shown in B) or through the oxidation of NADPH by NADPH oxidases. Subsequent reduction to H2O2 is catalyzed by superoxide dismutase (SOD). H2O2 can be further reduced to water (H2O) by catalase or can spontaneously oxidize iron (Fe2+) to form the highly reactive OH.. Under conditions of oxidative stress, when ROS generation outpaces the ROS scavenging system, accumulating levels of ROS oxidize and damage various cellular components. (B) The electron transport chain complexes I-IV harness electrons from NADH in a series of redox reactions, which are coupled to pumping protons (H+) into the mitochondrial intermembrane space. The proton motive force, a combination of the membrane potential (charge) and the concentration gradient (pH), powers ATP synthase (complex V). Normally, O2 acts as the final electron acceptor at complex IV, but aberrant reduction of O2 can occur at complexes I and III (red arrows), leading to the generation of O2.− (red).
Fig. 2.
Fig. 2.
Redox sensor molecules. Intricate control of reactive oxygen species (ROS) can be either directly or indirectly mediated by several transcription factors (blue), as well as by kinases (yellow) and phosphatases (green). Other regulators, such as the cytokine signaling inhibitor LNK, the modulator KEAP1, the E3 ubiquitin ligase MDM2, the cell cycle inhibitors p16INK4A and p19ARF (which are negatively modulated by the polycomb group member BMI1), the complex mTORC1, TXNIP, and the antioxidant enzyme GPX3 (all shown in orange) can also control ROS levels. Dashed arrows and lines indicate regulations that have not been explicitly shown to occur in stem cells; unbroken lines represent interactions that have been shown in stem cells. p53 has both antioxidant and pro-oxidant functions (shown in somatic cells). AKT, protein kinase B; FOXO, forkhead box O protein; GPX3, glutathione peroxidase 3; HIF, hypoxia-inducible factor; KEAP1, kelch-like ECH-associated protein 1; MDM2, transformed mouse 3T3 cell double minute 2; MEIS1, Meis homeobox 1; mTORC1, mammalian target of rapamycin complex 1; NRF2, nuclear factor erythroid 2-related factor 2; PTEN, phosphate and tensin homolog; TXNIP, thioredoxin-interacting protein;
Fig. 3.
Fig. 3.
Metabolic crosstalk between key signaling pathways in stem cells via ROS and other metabolic co-factors. Glycolysis (depicted by light-blue arrows) is a catabolic process, creating energy via the conversion of glucose to pyruvate. The glycolytic intermediate glucose 6-phosphate (G6P) can also be shunted into the pentose phosphate pathway (dark-blue arrows) to produce precursors for nucleotide biosynthesis and also to regenerate NADPH, a co-factor that replenishes the reduced glutathione pools. In turn, the antioxidant glutathione (GSH) mitigates oxidative stress. The pentose phosphate pathway is especially important for the anabolic (energy consuming) demands of stem cells and for the regeneration of glutathione. Pyruvate can be catalyzed to lactate to regenerate NAD+, a required co-factor for continued flux through glycolysis and is the preferred path for stem cells. Pyruvate can also be further metabolized inside the mitochondria, beginning with the conversion to acetyl CoA, which feeds into the tricarboxylic acid cycle (TCA) cycle. The TCA cycle generates reducing co-factors that power the electron transport chain (ETC) and subsequent production of ATP, a process known as oxidative phosphorylation. Some key metabolic enzymes that are described in the text are outlined in black. The metabolic processes described can affect concentrations of mitochondrial membrane potential (ΔΨ), metabolic intermediates (acetyl CoA, α-ketoglutarate), co-factors (NADPH, NAD+, AMP/ADP) and ROS, which in turn affect the function of nutrient-sensing (depicted by green arrows) and redox-sensitive proteins (depicted by gold stars). Many of these proteins can then alter cellular processes and ultimately regulate stem cell fate. Additionally, some proteins, such as the transcription factors FOXO3 and HIF1α, can modulate the expression of metabolic genes to fit the needs of stem cells (pink background depicts the metabolic reactions and blue background indicates signaling pathways. AKT, protein kinase B; AMPK, 5′ adenosine monophosphate-activated protein kinase; BMI1, Bmi1 polycomb ring finger oncogene; F6P, fructose 6-phosphate; F1,6P, fructose 1,6-bisphosphate; FOXO, forkhead box O protein; G3P, glyceraldehyde 3-phosphate; G6P, glucose 6-phosphate; GSSH, oxidized glutathione; HIF1α, hypoxia-inducible factor 1α; LDH, lactate dehydrogenase; LKB1 (STK11), serine/threonine kinase 11; LNK, SH2B adaptor protein 3 (SH2B3; mTORC1, mammalian target of rapamycin complex 1; p16INK4A, cyclin-dependent kinase inhibitor 2A (CDKN2A); p19ARF, cyclin-dependent kinase inhibitor 2A (CDKN2A); p38 MAPK, p38 mitogen-activated protein kinase; p53, transformation related protein 53 (TRP53); PDH, pyruvate dehydrogenase complex; PDK, pyruvate dehydrogenase kinase; PEP, phosphoenolpyruvate; PFK, phosphofructokinase; PI3K, phosphatidylinositol 3-kinase; PK, pyruvate kinase; PRDM16, PR domain containing 16; PTEN, phosphate and tensin homolog; R5P, ribulose 5-phosphate; RTK, receptor tyrosine kinase; SIRT1, sirtuin 1; TSC1/2, tuberous sclerosis 1/2; UCP2, uncoupling protein 2.
Fig. 4.
Fig. 4.
Effects of mitochondrial function on HSC maintenance and ROS production. Mitochondria in hematopoietic stem cells (HSCs) are characterized as having an immature morphology with lower metabolic activity, as determined by lower ATP output, O2 consumption, total mass and membrane potential (ΔΨ) when compared with more differentiated cells. These attributes appear to be required for stem cell and especially for HSC maintenance. The balance between stem cell maintenance (green) and the loss of quiescence and self-renewal capacity (red) can be influenced by the abundance and activity of certain proteins (blue boxes). Knock out of the corresponding genes leads to disruption of normal mitochondrial status in stem cells and changes in ROS either by decreasing the function and health of the mitochondria (top), increasing mitochondria biogenesis (middle) or activating the mitochondria oxidative phosphorylation pathway (bottom). BIM, BCL2-like 11; DJ-1, Parkinson disease (autosomal recessive, early onset) 7 (PARK7); HIF, hypoxia-inducible factor; LKB1 (STK11), serine/threonine kinase 11; PDK, pyruvate dehydrogenase kinase; TSC1, tuberous sclerosis 1; UCP2, uncoupling protein 2.
Fig. 5.
Fig. 5.
ROS effects on stem cells. Quiescent and/or self-renewing stem cells display low reactive oxygen species (ROS) levels due to their strong antioxidant machinery, which is maintained by proteins such as FOXO3 (forkhead box O3), p53 [transformation related protein 53 (TRP53)], HIF1 (hypoxia-inducible factor 1) and ATM (ataxia telangiectasia mutated kinase). Intermediate ROS levels prime stem cells for differentiation and under this context some proteins (such as p53) might act as pro-oxidant factors (dashed arrows). High ROS levels cause stem cell senescence and death. Question marks indicate unidentified proteins responsible for senescence and cell death under high ROS conditions in stem cells.

References

    1. Abbas H. A., MacCio D. R., Coskun S., Jackson J. G., Hazen A. L., Sills T. M., You M. J., Hirschi K. K. and Lozano G. (2010). Mdm2 is required for survival of hematopoietic stem cells/progenitors via dampening of ROS-induced p53 activity. Cell Stem Cell 7, 606-617 10.1016/j.stem.2010.09.013
    1. Anastasiou D., Poulogiannis G., Asara J. M., Boxer M. B., Jiang J.-k., Shen M., Bellinger G., Sasaki A. T., Locasale J. W., Auld D. S. et al. (2011). Inhibition of pyruvate kinase M2 by reactive oxygen species contributes to cellular antioxidant responses. Science 334, 1278-1283 10.1126/science.1211485
    1. Angkeow P., Deshpande S. S., Qi B., Liu Y.-X., Park Y. C., Jeon B. H., Ozaki M. and Irani K. (2002). Redox factor-1: an extra-nuclear role in the regulation of endothelial oxidative stress and apoptosis. Cell Death Differ. 9, 717-725 10.1038/sj.cdd.4401025
    1. Armstrong L., Tilgner K., Saretzki G., Atkinson S. P., Stojkovic M., Moreno R., Przyborski S. and Lako M. (2010). Human induced pluripotent stem cell lines show stress defense mechanisms and mitochondrial regulation similar to those of human embryonic stem cells. Stem Cells 28, 661-673 10.1002/stem.307
    1. Baccelli I. and Trumpp A. (2012). The evolving concept of cancer and metastasis stem cells. J. Cell Biol. 198, 281-293 10.1083/jcb.201202014
    1. Beckman K. B. and Ames B. N. (1998). The free radical theory of aging matures. Physiol. Rev. 78, 547-581.
    1. Bhandari D. R., Seo K.-W., Roh K.-H., Jung J.-W., Kang S.-K. and Kang K.-S. (2010). REX-1 expression and p38 MAPK activation status can determine proliferation/differentiation fates in human mesenchymal stem cells. PLoS ONE 5, e10493 10.1371/journal.pone.0010493
    1. Bonello S., Zähringer C., BelAiba R. S., Djordjevic T., Hess J., Michiels C., Kietzmann T. and Görlach A. (2007). Reactive oxygen species activate the HIF-1alpha promoter via a functional NFkappaB site. Arterioscler. Thromb. Vasc. Biol. 27, 755-761 10.1161/01.ATV.0000258979.92828.bc
    1. Brown K., Xie S., Qiu X., Mohrin M., Shin J., Liu Y., Zhang D., Scadden D. T. and Chen D. (2013). SIRT3 reverses aging-associated degeneration. Cell Rep. 3, 319-327 10.1016/j.celrep.2013.01.005
    1. Brunelle J. K., Bell E. L., Quesada N. M., Vercauteren K., Tiranti V., Zeviani M., Scarpulla R. C. and Chandel N. S. (2005). Oxygen sensing requires mitochondrial ROS but not oxidative phosphorylation. Cell Metabol. 1, 409-414 10.1016/j.cmet.2005.05.002
    1. Brunet A., Sweeney L. B., Sturgill J. F., Chua K. F., Greer P. L., Lin Y., Tran H., Ross S. E., Mostoslavsky R., Cohen H. Y. et al. (2004). Stress-dependent regulation of FOXO transcription factors by the SIRT1 deacetylase. Science 303, 2011-2015 10.1126/science.1094637
    1. Challen G. A., Sun D., Jeong M., Luo M., Jelinek J., Berg J. S., Bock C., Vasanthakumar A., Gu H., Xi Y. et al. (2012). Dnmt3a is essential for hematopoietic stem cell differentiation. Nat. Genet. 44, 23-31 10.1038/ng.1009
    1. Chandel N. S., Maltepe E., Goldwasser E., Mathieu C. E., Simon M. C. and Schumacker P. T. (1998). Mitochondrial reactive oxygen species trigger hypoxia-induced transcription. Proc. Natl. Acad. Sci. USA 95, 11715-11720 10.1073/pnas.95.20.11715
    1. Chen C., Liu Y., Liu R., Ikenoue T., Guan K.-L., Liu Y. and Zheng P. (2008). TSC-mTOR maintains quiescence and function of hematopoietic stem cells by repressing mitochondrial biogenesis and reactive oxygen species. J. Exp. Med. 205, 2397-2408 10.1084/jem.20081297
    1. Chuikov S., Levi B. P., Smith M. L. and Morrison S. J. (2010). Prdm16 promotes stem cell maintenance in multiple tissues, partly by regulating oxidative stress. Nat. Cell Biol. 12, 999-1006 10.1038/ncb2101
    1. Chung S., Dzeja P. P., Faustino R. S., Perez-Terzic C., Behfar A. and Terzic A (2007). Mitochondrial oxidative metabolism is required for the cardiac differentiation of stem cells. Nat. Clin. Pract. Cardiovasc. Med. 4Suppl. 1, S60-S67 10.1038/ncpcardio0766
    1. Daitoku H., Hatta M., Matsuzaki H., Aratani S., Ohshima T., Miyagishi M., Nakajima T. and Fukamizu A. (2004). Silent information regulator 2 potentiates Foxo1-mediated transcription through its deacetylase activity. Proc. Natl. Acad. Sci. U. S. A. 101, 10042-10047 10.1073/pnas.0400593101
    1. Dansen T. B., Smits L. M. M., van Triest M. H., de Keizer P. L. J., van Leenen D., Koerkamp M. G., Szypowska A., Meppelink A., Brenkman A. B., Yodoi J. et al. (2009). Redox-sensitive cysteines bridge p300/CBP-mediated acetylation and FoxO4 activity. Nat. Chem. Biol. 5, 664-672 10.1038/nchembio.194
    1. De Bock K., Georgiadou M., Schoors S., Kuchnio A., Wong B. W., Cantelmo A. R., Quaegebeur A., Ghesquière B., Cauwenberghs S., Eelen G. et al. (2013). Role of PFKFB3-driven glycolysis in vessel sprouting. Cell 154, 651-663 10.1016/j.cell.2013.06.037
    1. Diehn M., Cho R. W., Lobo N. A., Kalisky T., Dorie M. J., Kulp A. N., Qian D., Lam J. S., Ailles L. E., Wong M., et al. . (2009). Association of reactive oxygen species levels and radioresistance in cancer stem cells. Nature 458, 780-783 10.1038/nature07733
    1. Ding L., Liang X.-G., Hu Y., Zhu D.-Y. and Lou Y.-J. (2008). Involvement of p38MAPK and reactive oxygen species in icariin-induced cardiomyocyte differentiation of murine embryonic stem cells in vitro. Stem Cells Dev. 17, 751-760 10.1089/scd.2007.0206
    1. Dröge W. (2002). Aging-related changes in the thiol/disulfide redox state: implications for the use of thiol antioxidants. Exp. Gerontol. 37, 1333-1345 10.1016/S0531-5565(02)00175-4
    1. Du J., Li Q., Tang F., Puchowitz M. A., Fujioka H., Dunwoodie S. L., Danielpour D. and Yang Y.-C. (2014). Cited2 is required for the maintenance of glycolytic metabolism in adult hematopoietic stem cells. Stem Cells Dev. 23, 83-94 10.1089/scd.2013.0370
    1. Epstein A. C. R., Gleadle J. M., McNeill L. A., Hewitson K. S., O'Rourke J., Mole D. R., Mukherji M., Metzen E., Wilson M. I., Dhanda A. et al. (2001). C. elegans EGL-9 and mammalian homologs define a family of dioxygenases that regulate HIF by prolyl hydroxylation. Cell 107, 43-54 10.1016/S0092-8674(01)00507-4
    1. Ezashi T., Das P. and Roberts R. M. (2005). Low O2 tensions and the prevention of differentiation of hES cells. Proc. Natl. Acad. Sci. USA 102, 4783-4788 10.1073/pnas.0501283102
    1. Finkel T. (2003). Oxidant signals and oxidative stress. Curr. Opin. Cell Biol. 15, 247-254 10.1016/S0955-0674(03)00002-4
    1. Finkel T. (2011). Signal transduction by reactive oxygen species. J. Cell Biol. 194, 7-15 10.1083/jcb.201102095
    1. Folmes C. D. L., Nelson T. J., Martinez-Fernandez A., Arrell D. K., Lindor J. Z., Dzeja P. P., Ikeda Y., Perez-Terzic C. and Terzic A. (2011). Somatic oxidative bioenergetics transitions into pluripotency-dependent glycolysis to facilitate nuclear reprogramming. Cell Metabol. 14, 264-271 10.1016/j.cmet.2011.06.011
    1. Folmes C. D. L., Dzeja P. P., Nelson T. J. and Terzic A. (2012). Metabolic plasticity in stem cell homeostasis and differentiation. Cell Stem Cell 11, 596-606 10.1016/j.stem.2012.10.002
    1. Forsyth N. R., Musio A., Vezzoni P., Simpson A. H. R. W., Noble B. S. and McWhir J. (2006). Physiologic oxygen enhances human embryonic stem cell clonal recovery and reduces chromosomal abnormalities. Cloning Stem Cells 8, 16-23 10.1089/clo.2006.8.16
    1. Foudi A., Hochedlinger K., Van Buren D., Schindler J. W., Jaenisch R., Carey V. and Hock H. (2009). Analysis of histone 2B-GFP retention reveals slowly cycling hematopoietic stem cells. Nat. Biotechnol. 27, 84-90 10.1038/nbt.1517
    1. Friedman J. S., Lopez M. F., Fleming M. D., Rivera A., Martin F. M., Welsh M. L., Boyd A., Doctrow S. R. and Burakoff S. J. (2004). SOD2-deficiency anemia: protein oxidation and altered protein expression reveal targets of damage, stress response, and antioxidant responsiveness. Blood 104, 2565-2573 10.1182/blood-2003-11-3858
    1. Funes J. M., Quintero M., Henderson S., Martinez D., Qureshi U., Westwood C., Clements M. O., Bourboulia D., Pedley R. B., Moncada S. et al. (2007). Transformation of human mesenchymal stem cells increases their dependency on oxidative phosphorylation for energy production. Proc. Natl. Acad. Sci. USA 104, 6223-6228 10.1073/pnas.0700690104
    1. Gan B., Hu J., Jiang S., Liu Y., Sahin E., Zhuang L., Fletcher-Sananikone E., Colla S., Wang Y. A., Chin L. et al. (2010). Lkb1 regulates quiescence and metabolic homeostasis of haematopoietic stem cells. Nature 468, 701-704 10.1038/nature09595
    1. Gomes A. P., Price N. L., Ling A. J. Y., Moslehi J. J., Montgomery M. K., Rajman L., White J. P., Teodoro J. S., Wrann C. D., Hubbard B. P. et al. (2013). Declining NAD(+) induces a pseudohypoxic state disrupting nuclear-mitochondrial communication during aging. Cell 155, 1624-1638 10.1016/j.cell.2013.11.037
    1. Guo Y., Einhorn L., Kelley M., Hirota K., Yodoi J., Reinbold R., Scholer H., Ramsey H. and Hromas R. (2004). Redox regulation of the embryonic stem cell transcription factor oct-4 by thioredoxin. Stem Cells 22, 259-264 10.1634/stemcells.22-3-259
    1. Guo Y.-L., Chakraborty S., Rajan S. S., Wang R. and Huang F. (2010a). Effects of oxidative stress on mouse embryonic stem cell proliferation, apoptosis, senescence, and self-renewal. Stem Cells Dev. 19, 1321-1331 10.1089/scd.2009.0313
    1. Guo Z., Kozlov S., Lavin M. F., Person M. D. and Paull T. T. (2010b). ATM activation by oxidative stress. Science 330, 517-521 10.1126/science.1192912
    1. Gurumurthy S., Xie S. Z., Alagesan B., Kim J., Yusuf R. Z., Saez B., Tzatsos A., Ozsolak F., Milos P., Ferrari F. et al. (2010). The Lkb1 metabolic sensor maintains haematopoietic stem cell survival. Nature 468, 659-663 10.1038/nature09572
    1. Gut P. and Verdin E. (2013). The nexus of chromatin regulation and intermediary metabolism. Nature 502, 489-498 10.1038/nature12752
    1. Guzman J. N., Sanchez-Padilla J., Wokosin D., Kondapalli J., Ilijic E., Schumacker P. T. and Surmeier D. J. (2010). Oxidant stress evoked by pacemaking in dopaminergic neurons is attenuated by DJ-1. Nature 468, 696-700 10.1038/nature09536
    1. Haigis M. C. and Sinclair D. A. (2010). Mammalian sirtuins: biological insights and disease relevance. Annu. Rev. Pathol. 5, 253-295 10.1146/annurev.pathol.4.110807.092250
    1. Han M.-K., Song E.-K., Guo Y., Ou X., Mantel C. and Broxmeyer H. E. (2008). SIRT1 regulates apoptosis and Nanog expression in mouse embryonic stem cells by controlling p53 subcellular localization. Cell Stem Cell 2, 241-251 10.1016/j.stem.2008.01.002
    1. Harman D. (1972). Free radical theory of aging: dietary implications. Am. J. Clin. Nutr. 25, 839-843.
    1. Harris J. M., Esain V., Frechette G. M., Harris L. J., Cox A. G., Cortes M., Garnaas M. K., Carroll K. J., Cutting C. C., Khan T. et al. (2013). Glucose metabolism impacts the spatiotemporal onset and magnitude of HSC induction in vivo. Blood 121, 2483-2493 10.1182/blood-2012-12-471201
    1. Herault O., Hope K. J., Deneault E., Mayotte N., Chagraoui J., Wilhelm B. T., Cellot S., Sauvageau M., Andrade-Navarro M. A., Hébert J. et al. (2012). A role for GPx3 in activity of normal and leukemia stem cells. J. Exp. Med. 209, 895-901 10.1084/jem.20102386
    1. Hernández-García D., Wood C. D., Castro-Obregón S. and Covarrubias L. (2010). Reactive oxygen species: a radical role in development? Free Radic. Biol. Med. 49, 130-143 10.1016/j.freeradbiomed.2010.03.020
    1. Hjalgrim L. L., Rostgaard K., Hjalgrim H., Westergaard T., Thomassen H., Forestier E., Gustafsson G., Kristinsson J., Melbye M. and Schmiegelow K. (2004). Birth weight and risk for childhood leukemia in Denmark, Sweden, Norway, and Iceland. J. Natl. Cancer Inst. 96, 1549-1556 10.1093/jnci/djh287
    1. Hochmuth C. E., Biteau B., Bohmann D. and Jasper H. (2011). Redox regulation by Keap1 and Nrf2 controls intestinal stem cell proliferation in Drosophila. Cell Stem Cell 8, 188-199 10.1016/j.stem.2010.12.006
    1. Holmström K. M. and Finkel T. (2014). Cellular mechanisms and physiological consequences of redox-dependent signalling. Nat. Rev. Mol. Cell Biol. 15, 411-421 10.1038/nrm3801
    1. Hom J. R., Quintanilla R. A., Hoffman D. L., de Mesy Bentley K. L., Molkentin J. D., Sheu S.-S. and Porter G. A. (2011). The permeability transition pore controls cardiac mitochondrial maturation and myocyte differentiation. Dev. Cell 21, 469-478 10.1016/j.devcel.2011.08.008
    1. Irani K., Xia Y., Zweier J. L., Sollott S. J., Der C. J., Fearon E. R., Sundaresan M., Finkel T. and Goldschmidt-Clermont P. J. (1997). Mitogenic signaling mediated by oxidants in Ras-transformed fibroblasts. Science 275, 1649-1652 10.1126/science.275.5306.1649
    1. Ito K., Hirao A., Arai F., Matsuoka S., Takubo K., Hamaguchi I., Nomiyama K., Hosokawa K., Sakurada K., Nakagata N. et al. (2004). Regulation of oxidative stress by ATM is required for self-renewal of haematopoietic stem cells. Nature 431, 997-1002 10.1038/nature02989
    1. Ito K., Hirao A., Arai F., Takubo K., Matsuoka S., Miyamoto K., Ohmura M., Naka K., Hosokawa K., Ikeda Y. et al. (2006). Reactive oxygen species act through p38 MAPK to limit the lifespan of hematopoietic stem cells. Nat. Med. 12, 446-451 10.1038/nm1388
    1. Ito K., Carracedo A., Weiss D., Arai F., Ala U., Avigan D. E., Schafer Z. T., Evans R. M., Suda T., Lee C.-H. et al. (2012). A PML–PPAR-δ pathway for fatty acid oxidation regulates hematopoietic stem cell maintenance. Nat. Med. 18, 1350-1358 10.1038/nm.2882
    1. Itoh K., Wakabayashi N., Katoh Y., Ishii T., Igarashi K., Engel J. D. and Yamamoto M. (1999). Keap1 represses nuclear activation of antioxidant responsive elements by Nrf2 through binding to the amino-terminal Neh2 domain. Genes Dev. 13, 76-86 10.1101/gad.13.1.76
    1. Jang Y.-Y. and Sharkis S. J. (2007). A low level of reactive oxygen species selects for primitive hematopoietic stem cells that may reside in the low-oxygenic niche. Blood 110, 3056-3063 10.1182/blood-2007-05-087759
    1. Janssen-Heininger Y. M. W., Mossman B. T., Heintz N. H., Forman H. J., Kalyanaraman B., Finkel T., Stamler J. S., Rhee S. G. and van der Vliet A. (2008). Redox-based regulation of signal transduction: principles, pitfalls, and promises. Free Radic. Biol. Med. 45, 1-17 10.1016/j.freeradbiomed.2008.03.011
    1. Ji J., Sharma V., Qi S., Guarch M. E., Zhao P., Luo Z., Fan W., Wang Y., Mbabaali F., Neculai D. et al. (2014). Antioxidant supplementation reduces genomic aberrations in human induced pluripotent stem cells. Stem Cell Rep. 2, 44-51 10.1016/j.stemcr.2013.11.004
    1. Joshi A. and Kundu M. (2013). Mitophagy in hematopoietic stem cells: the case for exploration. Autophagy 9, 1737-1749 10.4161/auto.26681
    1. Jung H., Kim M. J., Kim D. O., Kim W. S., Yoon S.-J., Park Y.-J., Yoon S. R., Kim T.-D., Suh H.-W., Yun S. et al. (2013). TXNIP maintains the hematopoietic cell pool by switching the function of p53 under oxidative stress. Cell Metabol. 18, 75-85 10.1016/j.cmet.2013.06.002
    1. Juntilla M. M., Patil V. D., Calamito M., Joshi R. P., Birnbaum M. J. and Koretzky G. A. (2010). AKT1 and AKT2 maintain hematopoietic stem cell function by regulating reactive oxygen species. Blood 115, 4030-4038 10.1182/blood-2009-09-241000
    1. Kanda Y., Hinata T., Kang S. W. and Watanabe Y. (2011). Reactive oxygen species mediate adipocyte differentiation in mesenchymal stem cells. Life Sci. 89, 250-258 10.1016/j.lfs.2011.06.007
    1. Kang J., Gemberling M., Nakamura M., Whitby F. G., Handa H., Fairbrother W. G. and Tantin D. (2009). A general mechanism for transcription regulation by Oct1 and Oct4 in response to genotoxic and oxidative stress. Genes Dev. 23, 208-222 10.1101/gad.1750709
    1. Kim J. and Wong P. K. Y. (2009). Loss of ATM impairs proliferation of neural stem cells through oxidative stress-mediated p38 MAPK signaling. Stem Cells 27, 1987-1998 10.1002/stem.125
    1. Kim Y. S., Kang M. J. and Cho Y. M. (2013). Low production of reactive oxygen species and high DNA repair: mechanism of radioresistance of prostate cancer stem cells. Anticancer Res. 33, 4469-4474.
    1. Kobayashi Y., Furukawa-Hibi Y., Chen C., Horio Y., Isobe K., Ikeda K. and Motoyama N. (2005). SIRT1 is critical regulator of FOXO-mediated transcription in response to oxidative stress. Int. J. Mol. Med. 16, 237-243.
    1. Kocabas F., Zheng J., Thet S., Copeland N. G., Jenkins N. A., DeBerardinis R. J., Zhang C. and Sadek A. H. (2012). Meis1 regulates the metabolic phenotype and oxidant defense of hematopoietic stem cells. Blood 120, 4963-4972 10.1182/blood-2012-05-432260
    1. Kohli R. M. and Zhang Y. (2013). TET enzymes, TDG and the dynamics of DNA demethylation. Nature 502, 472-479 10.1038/nature12750
    1. Kunisaki Y., Bruns I., Scheiermann C., Ahmed J., Pinho S., Zhang D., Mizoguchi T., Wei Q., Lucas D., Ito K. et al. (2013). Arteriolar niches maintain haematopoietic stem cell quiescence. Nature 502, 637-643 10.1038/nature12612
    1. Lagadinou E. D., Sach A., Callahan K., Rossi R. M., Neering S. J., Minhajuddin M., Ashton J. M., Pei S., Grose V., O'Dwyer K. M. et al. (2013). BCL-2 inhibition targets oxidative phosphorylation and selectively eradicates quiescent human leukemia stem cells. Cell Stem Cell 12, 329-341 10.1016/j.stem.2012.12.013
    1. Lapointe J. and Hekimi S. (2010). When a theory of aging ages badly. Cell. Mol. Life Sci. 67, 1-8 10.1007/s00018-009-0138-8
    1. Le Belle J. E., Orozco N. M., Paucar A. A., Saxe J. P., Mottahedeh J., Pyle A. D., Wu H. and Kornblum H. I. (2011). Proliferative neural stem cells have high endogenous ROS levels that regulate self-renewal and neurogenesis in a PI3K/Akt-dependant manner. Cell Stem Cell 8, 59-71 10.1016/j.stem.2010.11.028
    1. Lee J. M. (1998). Inhibition of p53-dependent apoptosis by the KIT tyrosine kinase: regulation of mitochondrial permeability transition and reactive oxygen species generation. Oncogene 17, 1653-1662 10.1038/sj.onc.1202102
    1. Lee S.-R., Yang K.-S., Kwon J., Lee C., Jeong W. and Rhee S. G. (2002). Reversible inactivation of the tumor suppressor PTEN by H2O2. J. Biol. Chem. 277, 20336-20342 10.1074/jbc.M111899200
    1. Lessard J. and Sauvageau G. (2003). Bmi-1 determines the proliferative capacity of normal and leukaemic stem cells. Nature 423, 255-260 10.1038/nature01572
    1. Liang R. and Ghaffari S. (2014). Stem cells, redox signaling, and stem cell aging. Antioxid. Redox Signal. 20, 1902-1916 10.1089/ars.2013.5300
    1. Lim J.-H., Lee Y.-M., Chun Y.-S., Chen J., Kim J.-E. and Park J.-W. (2010). Sirtuin 1 modulates cellular responses to hypoxia by deacetylating hypoxia-inducible factor 1alpha. Mol. Cell 38, 864-878 10.1016/j.molcel.2010.05.023
    1. Liu H., Colavitti R., Rovira I. I. and Finkel T. (2005). Redox-dependent transcriptional regulation. Circ. Res. 97, 967-974 10.1161/01.RES.0000188210.72062.10
    1. Liu Y., Elf S. E., Miyata Y., Sashida G., Liu Y., Huang G., Di Giandomenico S., Lee J. M., Deblasio A., Menendez S. et al. (2009a). p53 regulates hematopoietic stem cell quiescence. Cell Stem Cell 4, 37-48 10.1016/j.stem.2008.11.006
    1. Liu J., Cao L., Chen J., Song S., Lee I. H., Quijano C., Liu H., Keyvanfar K., Chen H., Cao L.-Y. et al. (2009b). Bmi1 regulates mitochondrial function and the DNA damage response pathway. Nature 459, 387-392 10.1038/nature08040
    1. Liu J., Cao L. and Finkel T. (2011). Oxidants, metabolism, and stem cell biology. Free Radic. Biol. Med. 51, 2158-2162 10.1016/j.freeradbiomed.2011.10.434
    1. Mandal S., Lindgren A. G., Srivastava A. S., Clark A. T. and Banerjee U. (2011). Mitochondrial function controls proliferation and early differentiation potential of embryonic stem cells. Stem Cells 29, 486-495 10.1002/stem.590
    1. Marty C., Lacout C., Droin N., Le Couédic J.-P., Ribrag V., Solary E., Vainchenker W., Villeval J.-L. and Plo I. (2013). A role for reactive oxygen species in JAK2V617F myeloproliferative neoplasm progression. Leukemia 27, 2187-2195 10.1038/leu.2013.102
    1. Maryanovich M., Oberkovitz G., Niv H., Vorobiyov L., Zaltsman Y., Brenner O., Lapidot T., Jung S. and Gross A. (2012). The ATM-BID pathway regulates quiescence and survival of haematopoietic stem cells. Nat. Cell Biol. 14, 535-541 10.1038/ncb2468
    1. Mishra B. P., Zaffuto K. M., Artinger E. L., Org T., Mikkola H. K. A., Cheng C., Djabali M. and Ernst P. (2014). The histone methyltransferase activity of MLL1 is dispensable for hematopoiesis and leukemogenesis. Cell Rep. 7, 1239-1247 10.1016/j.celrep.2014.04.015
    1. Miyamoto K., Araki K. Y., Naka K., Arai F., Takubo K., Yamazaki S., Matsuoka S., Miyamoto T., Ito K., Ohmura M. et al. (2007). Foxo3a is essential for maintenance of the hematopoietic stem cell pool. Cell Stem Cell 1, 101-112 10.1016/j.stem.2007.02.001
    1. Molofsky A. V., Pardal R., Iwashita T., Park I.-K., Clarke M. F. and Morrison S. J. (2003). Bmi-1 dependence distinguishes neural stem cell self-renewal from progenitor proliferation. Nature 425, 962-967 10.1038/nature02060
    1. Moran-Crusio K., Reavie L., Shih A., Abdel-Wahab O., Ndiaye-Lobry D., Lobry C., Figueroa M. E., Vasanthakumar A., Patel J., Zhao X. et al. (2011). Tet2 loss leads to increased hematopoietic stem cell self-renewal and myeloid transformation. Cancer Cell 20, 11-24 10.1016/j.ccr.2011.06.001
    1. Morimoto H., Iwata K., Ogonuki N., Inoue K., Atsuo O., Kanatsu-Shinohara M., Morimoto T., Yabe-Nishimura C. and Shinohara T. (2013). ROS are required for mouse spermatogonial stem cell self-renewal. Cell Stem Cell 12, 774-786 10.1016/j.stem.2013.04.001
    1. Motohashi H. and Yamamoto M. (2004). Nrf2-Keap1 defines a physiologically important stress response mechanism. Trends Mol. Med. 10, 549-557 10.1016/j.molmed.2004.09.003
    1. Motta M. C., Divecha N., Lemieux M., Kamel C., Chen D., Gu W., Bultsma Y., McBurney M. and Guarente L. (2004). Mammalian SIRT1 represses forkhead transcription factors. Cell 116, 551-563 10.1016/S0092-8674(04)00126-6
    1. Mouchiroud L., Houtkooper R. H., Moullan N., Katsyuba E., Ryu D., Cantó C., Mottis A., Jo Y.-S., Viswanathan M., Schoonjans K. et al. (2013). The NAD(+)/sirtuin pathway modulates longevity through activation of mitochondrial UPR and FOXO signaling. Cell 154, 430-441 10.1016/j.cell.2013.06.016
    1. Murata H., Ihara Y., Nakamura H., Yodoi J., Sumikawa K. and Kondo T. (2003). Glutaredoxin exerts an antiapoptotic effect by regulating the redox state of Akt. J. Biol. Chem. 278, 50226-50233 10.1074/jbc.M310171200
    1. Murphy M. P. (2009). How mitochondria produce reactive oxygen species. Biochem. J. 417, 1-13 10.1042/BJ20081386
    1. Murphy M. P., Holmgren A., Larsson N.-G., Halliwell B., Chang C. J., Kalyanaraman B., Rhee S. G., Thornalley P. J., Partridge L., Gems D. et al. (2011). Unraveling the biological roles of reactive oxygen species. Cell Metabol. 13, 361-366 10.1016/j.cmet.2011.03.010
    1. Murry C. E. and Keller G. (2008). Differentiation of embryonic stem cells to clinically relevant populations: lessons from embryonic development. Cell 132, 661-680 10.1016/j.cell.2008.02.008
    1. Nakada D., Saunders T. L. and Morrison S. J. (2010). Lkb1 regulates cell cycle and energy metabolism in haematopoietic stem cells. Nature 468, 653-658 10.1038/nature09571
    1. Nathan C. and Cunningham-Bussel A. (2013). Beyond oxidative stress: an immunologist's guide to reactive oxygen species. Nat. Rev. Immunol. 13, 349-361 10.1038/nri3423
    1. Norddahl G. L., Pronk C. J., Wahlestedt M., Sten G., Nygren J. M., Ugale A., Sigvardsson M. and Bryder D. (2011). Accumulating mitochondrial DNA mutations drive premature hematopoietic aging phenotypes distinct from physiological stem cell aging. Cell Stem Cell 8, 499-510 10.1016/j.stem.2011.03.009
    1. Oburoglu L., Tardito S., Fritz V., de Barros S. C., Merida P., Craveiro M., Mamede J., Cretenet G., Mongellaz C., An X. et al. (2014). Glucose and glutamine metabolism regulate human hematopoietic stem cell lineage specification. Cell Stem Cell p15, 169-184. 10.1016/j.stem.2014.06.002
    1. Ou X., Lee M. R., Huang X., Messina-Graham S. and Broxmeyer H. E. (2014). SIRT1 positively regulates autophagy and mitochondria function in embryonic stem cells under oxidative stress. Stem Cells 32, 1183-1194 10.1002/stem.1641
    1. Owusu-Ansah E. and Banerjee U. (2009). Reactive oxygen species prime Drosophila haematopoietic progenitors for differentiation. Nature 461, 537-541 10.1038/nature08313
    1. Paik J.-h., Ding Z., Narurkar R., Ramkissoon S., Muller F., Kamoun W. S., Chae S.-S., Zheng H., Ying H., Mahoney J. et al. (2009). FoxOs cooperatively regulate diverse pathways governing neural stem cell homeostasis. Cell Stem Cell 5, 540-553 10.1016/j.stem.2009.09.013
    1. Parks D., Bolinger R. and Mann K. (1997). Redox state regulates binding of p53 to sequence-specific DNA, but not to non-specific or mismatched DNA. Nucleic Acids Res. 25, 1289-1295 10.1093/nar/25.6.1289
    1. Pei S., Minhajuddin M., Callahan K. P., Balys M., Ashton J. M., Neering S. J., Lagadinou E. D., Corbett C., Ye H., Liesveld J. L. et al. (2013). Targeting aberrant glutathione metabolism to eradicate human acute myelogenous leukemia cells. J. Biol. Chem. 288, 33542-33558 10.1074/jbc.M113.511170
    1. Polyak K., Xia Y., Zweier J. L., Kinzler K. W. and Vogelstein B. (1997). A model for p53-induced apoptosis. Nature 389, 300-305 10.1038/38525
    1. Prigione A., Fauler B., Lurz R., Lehrach H. and Adjaye J. (2010). The senescence-related mitochondrial/oxidative stress pathway is repressed in human induced pluripotent stem cells. Stem Cells 28, 721-733 10.1002/stem.404
    1. Renault V. M., Rafalski V. A., Morgan A. A., Salih D. A. M., Brett J. O., Webb A. E., Villeda S. A., Thekkat P. U., Guillerey C., Denko N. C. et al. (2009). FoxO3 regulates neural stem cell homeostasis. Cell Stem Cell 5, 527-539 10.1016/j.stem.2009.09.014
    1. Rimmelé P., Bigarella C. L., Liang R., Izac B., Dieguez-Gonzalez R., Barbet G., Donovan M., Brugnara C., Blander J. M., Sinclair D. A. et al. . 2014). Aging-like phenotype and defective lineage specification in SIRT1-deleted hematopoietic stem and progenitor cells. Stem Cell Rep. 3, 44-59 10.1016/j.stemcr.2014.04.015
    1. Ritterson Lew C. and Tolan D. R. (2013). Aldolase sequesters WASP and affects WASP/Arp2/3-stimulated actin dynamics. J. Cell. Biochem. 114, 1928-1939 10.1002/jcb.24538
    1. Rodgers J. T., Lerin C., Haas W., Gygi S. P., Spiegelman B. M. and Puigserver P. (2005). Nutrient control of glucose homeostasis through a complex of PGC-1alpha and SIRT1. Nature 434, 113-118 10.1038/nature03354
    1. Rodrigues M., Turner O., Stolz D., Griffith L. G. and Wells A. (2012). Production of reactive oxygen species by multipotent stromal cells/mesenchymal stem cells upon exposure to fas ligand. Cell Transplant. 21, 2171-2187 10.3727/096368912X639035
    1. Rouault-Pierre K., Lopez-Onieva L., Foster K., Anjos-Afonso F., Lamrissi-Garcia I., Serrano-Sanchez M., Mitter R., Ivanovic Z., de Verneuil H., Gribben J. et al. (2013). HIF-2α protects human hematopoietic stem/progenitors and acute myeloid leukemic cells from apoptosis induced by endoplasmic reticulum stress. Cell Stem Cell 13, 549-563 10.1016/j.stem.2013.08.011
    1. Sablina A. A., Budanov A. V., Ilyinskaya G. V., Agapova L. S., Kravchenko J. E. and Chumakov P. M. (2005). The antioxidant function of the p53 tumor suppressor. Nat. Med. 11, 1306-1313 10.1038/nm1320
    1. Saito Y., Uchida N., Tanaka S., Suzuki N., Tomizawa-Murasawa M., Sone A., Najima Y., Takagi S., Aoki Y., Wake A. et al. (2010). Induction of cell cycle entry eliminates human leukemia stem cells in a mouse model of AML. Nat. Biotechnol. 28, 275-280.
    1. Saitoh M., Nishitoh H., Fujii M., Takeda K., Tobiume K., Sawada Y., Kawabata M., Miyazono K. and Ichijo H. (1998). Mammalian thioredoxin is a direct inhibitor of apoptosis signal-regulating kinase (ASK) 1. EMBO J. 17, 2596-2606 10.1093/emboj/17.9.2596
    1. Salmeen A., Andersen J. N., Myers M. P., Meng T.-C., Hinks J. A., Tonks N. K. and Barford D. (2003). Redox regulation of protein tyrosine phosphatase 1B involves a sulphenyl-amide intermediate. Nature 423, 769-773 10.1038/nature01680
    1. Sarbassov D. D. and Sabatini D. M. (2005). Redox regulation of the nutrient-sensitive raptor-mTOR pathway and complex. J. Biol. Chem. 280, 39505-39509 10.1074/jbc.M506096200
    1. Schieke S. M., Ma M., Cao L., McCoy J. P., Liu C., Hensel N. F., Barrett A. J., Boehm M. and Finkel T. (2008). Mitochondrial metabolism modulates differentiation and teratoma formation capacity in mouse embryonic stem cells. J. Biol. Chem. 283, 28506-28512 10.1074/jbc.M802763200
    1. Signer R. A. J. and Morrison S. J. (2013). Mechanisms that regulate stem cell aging and life span. Cell Stem Cell 12, 152-165 10.1016/j.stem.2013.01.001
    1. Simsek T., Kocabas F., Zheng J., Deberardinis R. J., Mahmoud A. I., Olson E. N., Schneider J. W., Zhang C. C. and Sadek H. A. (2010). The distinct metabolic profile of hematopoietic stem cells reflects their location in a hypoxic niche. Cell Stem Cell 7, 380-390 10.1016/j.stem.2010.07.011
    1. Singh S. K., Williams C. A., Klarmann K., Burkett S. S., Keller J. R. and Oberdoerffer P. (2013). Sirt1 ablation promotes stress-induced loss of epigenetic and genomic hematopoietic stem and progenitor cell maintenance. J. Exp. Med. 210, 987-1001 10.1084/jem.20121608
    1. Song J. S., Cho H. H., Lee B.-J., Bae Y. C. and Jung J. S. (2011). Role of thioredoxin 1 and thioredoxin 2 on proliferation of human adipose tissue-derived mesenchymal stem cells. Stem Cells Dev. 20, 1529-1537 10.1089/scd.2010.0364
    1. Sundaresan M., Yu Z.-X., Ferrans V. J., Irani K. and Finkel T. (1995). Requirement for generation of H2O2 for platelet-derived growth factor signal transduction. Science 270, 296-299 10.1126/science.270.5234.296
    1. Sutendra G., Kinnaird A., Dromparis P., Paulin R., Stenson T. H., Haromy A., Hashimoto K., Zhang N., Flaim E. and Michelakis E. D. (2014). A nuclear pyruvate dehydrogenase complex is important for the generation of acetyl-CoA and histone acetylation. Cell 158, 84-97 10.1016/j.cell.2014.04.046
    1. Tahara E. B., Navarete F. D. T. and Kowaltowski A. J. (2009). Tissue-, substrate-, and site-specific characteristics of mitochondrial reactive oxygen species generation. Free Radic. Biol. Med. 46, 1283-1297 10.1016/j.freeradbiomed.2009.02.008
    1. Tahiliani M., Koh K. P., Shen Y., Pastor W. A., Bandukwala H., Brudno Y., Agarwal S., Iyer L. M., Liu D. R., Aravind L. et al. (2009). Conversion of 5-methylcytosine to 5-hydroxymethylcytosine in mammalian DNA by MLL partner TET1. Science 324, 930-935 10.1126/science.1170116
    1. Tai-Nagara I., Matsuoka S., Ariga H. and Suda T. (2014). Mortalin and DJ-1 coordinately regulate hematopoietic stem cell function through the control of oxidative stress. Blood 123, 41-50 10.1182/blood-2013-06-508333
    1. Takahashi K. and Yamanaka S. (2006). Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 126, 663-676 10.1016/j.cell.2006.07.024
    1. Takubo K., Goda N., Yamada W., Iriuchishima H., Ikeda E., Kubota Y., Shima H., Johnson R. S., Hirao A., Suematsu M. et al. (2010). Regulation of the HIF-1alpha level is essential for hematopoietic stem cells. Cell Stem Cell 7, 391-402 10.1016/j.stem.2010.06.020
    1. Takubo K., Nagamatsu G., Kobayashi C. I., Nakamura-Ishizu A., Kobayashi H., Ikeda E., Goda N., Rahimi Y., Johnson R. S., Soga T. et al. (2013). Regulation of glycolysis by pdk functions as a metabolic checkpoint for cell cycle quiescence in hematopoietic stem cells. Cell Stem Cell 12, 49-61 10.1016/j.stem.2012.10.011
    1. TeKippe M., Harrison D. E. and Chen J (2003). Expansion of hematopoietic stem cell phenotype and activity in Trp53-null mice. Exp. Hematol. 31, 521-527. 10.1016/S0301-472X(03)00072-9
    1. Tohyama S., Hattori F., Sano M., Hishiki T., Nagahata Y., Matsuura T., Hashimoto H., Suzuki T., Yamashita H., Satoh Y. et al. (2012). Distinct metabolic flow enables large-scale purification of mouse and human pluripotent stem cell-derived cardiomyocytes. Cell Stem Cell 12, 127-137 10.1016/j.stem.2012.09.013
    1. Toledano M. B. and Leonard W. J. (1991). Modulation of transcription factor NF-kappa B binding activity by oxidation-reduction in vitro. Proc. Natl. Acad. Sci. USA 88, 4328-4332 10.1073/pnas.88.10.4328
    1. Tomko R. J., Bansal P. and Lazo J. S. (2006). Airing out an antioxidant role for the tumor suppressor p53. Mol. Interv. 6, 23-25, 2 10.1124/mi.6.1.5
    1. Tormos K. V., Anso E., Hamanaka R. B., Eisenbart J., Joseph J., Kalyanaraman B. and Chandel N. S. (2011). Mitochondrial complex III ROS regulate adipocyte differentiation. Cell Metabol. 14, 537-544 10.1016/j.cmet.2011.08.007
    1. Tothova Z., Kollipara R., Huntly B. J., Lee B. H., Castrillon D. H., Cullen D. E., McDowell E. P., Lazo-Kallanian S., Williams I. R., Sears C. et al. (2007). FoxOs are critical mediators of hematopoietic stem cell resistance to physiologic oxidative stress. Cell 128, 325-339 10.1016/j.cell.2007.01.003
    1. Trowbridge J. J., Snow J. W., Kim J. and Orkin S. H. (2009). DNA methyltransferase 1 is essential for and uniquely regulates hematopoietic stem and progenitor cells. Cell Stem Cell 5, 442-449 10.1016/j.stem.2009.08.016
    1. Tsai W. B., Chung Y. M., Takahashi Y., Xu Z., Hu, M. C., (2008). Functional interaction between FOXO3a and ATM regulates DNA damage response. Nat. Cell Biol. 10, 460-467 10.1038/ncb1709
    1. Tsai J. J., Dudakov J. A., Takahashi K., Shieh J.-H., Velardi E., Holland A. M., Singer N. V., West M. L., Smith O. M., Young L. F. et al. (2013). Nrf2 regulates haematopoietic stem cell function. Nat. Cell Biol. 15, 309-316 10.1038/ncb2699
    1. Unwin R. D., Smith D. L., Blinco D., Wilson C. L., Miller C. J., Evans C. A., Jaworska E., Baldwin S. A., Barnes K., Pierce A., et al. . (2006). Quantitative proteomics reveals posttranslational control as a regulatory factor in primary hematopoietic stem cells. Blood 107, 4687-4694 10.1182/blood-2005-12-4995
    1. Urao N. and Ushio-Fukai M. (2013). Redox regulation of stem/progenitor cells and bone marrow niche. Free Radic. Biol. Med. 54, 26-39 10.1016/j.freeradbiomed.2012.10.532
    1. Vander Heiden M. G., Chandel N. S., Williamson E. K., Schumacker P. T. and Thompson C. B. (1997). Bcl-xL regulates the membrane potential and volume homeostasis of mitochondria. Cell 91, 627-637 10.1016/S0092-8674(00)80450-X
    1. Van der Horst A., Tertoolen L. G. J., de Vries-Smits L. M. M., Frye R. A., Medema R. H. and Burgering B. M. T. (2004). FOXO4 is acetylated upon peroxide stress and deacetylated by the longevity protein hSir2(SIRT1). J. Biol. Chem. 279, 28873-28879 10.1074/jbc.M401138200
    1. Vaziri H., Dessain S. K., Ng Eaton E., Imai S.-I., Frye R. A., Pandita T. K., Guarente L. and Weinberg R. A. (2001). hSIR2(SIRT1) functions as an NAD-dependent p53 deacetylase. Cell 107, 149-159 10.1016/S0092-8674(01)00527-X
    1. Velu C. S., Niture S. K., Doneanu C. E., Pattabiraman N. and Srivenugopal K. S. (2007). Human p53 is inhibited by glutathionylation of cysteines present in the proximal DNA-binding domain during oxidative stress. Biochemistry 46, 7765-7780 10.1021/bi700425y
    1. Venugopal R. and Jaiswal A. K. (1996). Nrf1 and Nrf2 positively and c-Fos and Fra1 negatively regulate the human antioxidant response element-mediated expression of NAD(P)H:quinone oxidoreductase1 gene. Proc. Natl. Acad. Sci. USA 93, 14960-14965 10.1073/pnas.93.25.14960
    1. Walker L. J., Robson C. N., Black E., Gillespie D. and Hickson I. D. (1993). Identification of residues in the human DNA repair enzyme HAP1 (Ref-1) that are essential for redox regulation of Jun DNA binding. Mol. Cell. Biol. 13, 5370-5376.
    1. Wang N., Xie K., Huo S., Zhao J., Zhang S. and Miao J. (2007). Suppressing phosphatidylcholine-specific phospholipase C and elevating ROS level, NADPH oxidase activity and Rb level induced neuronal differentiation in mesenchymal stem cells. J. Cell. Biochem. 100, 1548-1557 10.1002/jcb.21139
    1. Wang J., Alexander P., Wu L., Hammer R., Cleaver O. and McKnight S. L. (2009). Dependence of mouse embryonic stem cells on threonine catabolism. Science 325, 435-439 10.1126/science.1173288
    1. Wang Y.-H., Israelsen W. J., Lee D., Yu V. W. C., Jeanson N. T., Clish C. B., Cantley L. C., Vander Heiden M. G. and Scadden D. T. (2014). Cell-state-specific metabolic dependency in hematopoiesis and leukemogenesis. Cell 158, 1309-1323 10.1016/j.cell.2014.07.048
    1. Ward P. S. and Thompson C. B. (2012). Metabolic reprogramming: a cancer hallmark even warburg did not anticipate. Cancer Cell 21, 297-308 10.1016/j.ccr.2012.02.014
    1. Will B., Vogler T. O., Bartholdy B., Garrett-Bakelman F., Mayer J., Barreyro L., Pandolfi A., Todorova T. I., Okoye-Okafor U. C., Stanley R. F. et al. (2013). Satb1 regulates the self-renewal of hematopoietic stem cells by promoting quiescence and repressing differentiation commitment. Nat. Immunol. 14, 437-445 10.1038/ni.2572
    1. Wilson A., Laurenti E., Oser G., van der Wath R. C., Blanco-Bose W., Jaworski M., Offner S., Dunant C. F., Eshkind L., Bockamp E. et al. (2008). Hematopoietic stem cells reversibly switch from dormancy to self-renewal during homeostasis and repair. Cell 135, 1118-1129 10.1016/j.cell.2008.10.048
    1. Winterbourn C. (2014). Current methods to study reactive oxygen species--pros and cons. Preface. Biochim. Biophys. Acta 1840, 707 10.1016/j.bbagen.2013.09.021
    1. Xiao Q., Luo Z., Pepe A. E., Margariti A., Zeng L. and Xu Q. (2009). Embryonic stem cell differentiation into smooth muscle cells is mediated by Nox4-produced H2O2. Am. J. Physiol. Cell Physiol. 296, C711-C723 10.1152/ajpcell.00442.2008
    1. Xiao M., Yang H., Xu W., Ma S., Lin H., Zhu H., Liu L., Liu Y., Yang C., Xu Y. et al. (2012). Inhibition of α-KG-dependent histone and DNA demethylases by fumarate and succinate that are accumulated in mutations of FH and SDH tumor suppressors. Genes Dev. 26, 1326-1338 10.1101/gad.191056.112
    1. Yahata T., Takanashi T., Muguruma Y., Ibrahim A. A., Matsuzawa H., Uno T., Sheng Y., Onizuka M., Ito M., Kato S. et al. (2011). Accumulation of oxidative DNA damage restricts the self-renewal capacity of human hematopoietic stem cells. Blood 118, 2941-2950 10.1182/blood-2011-01-330050
    1. Yalcin S., Zhang X., Luciano J. P., Mungamuri S. K., Marinkovic D., Vercherat C., Sarkar A., Grisotto M., Taneja R. and Ghaffari S. (2008). Foxo3 is essential for the regulation of ataxia telangiectasia mutated and oxidative stress-mediated homeostasis of hematopoietic stem cells. J. Biol. Chem. 283, 25692-25705 10.1074/jbc.M800517200
    1. Yalcin S., Marinkovic D., Mungamuri S. K., Zhang X., Tong W., Sellers R. and Ghaffari S. (2010). ROS-mediated amplification of AKT/mTOR signalling pathway leads to myeloproliferative syndrome in Foxo3(-/-) mice. EMBO J. 29, 4118-4131 10.1038/emboj.2010.292
    1. Yanes O., Clark J., Wong D. M., Patti G. J., Sánchez-Ruiz A., Benton H. P., Trauger S. A., Desponts C., Ding S. and Siuzdak G. (2010). Metabolic oxidation regulates embryonic stem cell differentiation. Nat. Chem. Biol. 6, 411-417 10.1038/nchembio.364
    1. Yang W., Xia Y., Ji H., Zheng Y., Liang J., Huang W., Gao X., Aldape K. and Lu Z. (2011). Nuclear PKM2 regulates β-catenin transactivation upon EGFR activation. Nature 478, 118-122 10.1038/nature10598
    1. Yeo H., Lyssiotis C. A., Zhang Y., Ying H., Asara J. M., Cantley L. C. and Paik J.-H. (2013). FoxO3 coordinates metabolic pathways to maintain redox balance in neural stem cells. EMBO J. 32, 2531-2656 10.1038/emboj.2013.186
    1. Yoshida S., Hong S., Suzuki T., Nada S., Mannan A. M., Wang J., Okada M., Guan K.-L. and Inoki K. (2011). Redox regulates mammalian target of rapamycin complex 1 (mTORC1) activity by modulating the TSC1/TSC2-Rheb GTPase pathway. J. Biol. Chem. 286, 32651-32660 10.1074/jbc.M111.238014
    1. Youle R. J. and van der Bliek A. M. (2012). Mitochondrial fission, fusion, and stress. Science 337, 1062-1065 10.1126/science.1219855
    1. Yu W.-M., Liu X., Shen J., Jovanovic O., Pohl E. E., Gerson S. L., Finkel T., Broxmeyer H. E. and Qu C.-K. (2013). Metabolic regulation by the mitochondrial phosphatase PTPMT1 is required for hematopoietic stem cell differentiation. Cell Stem Cell 12, 62-74 10.1016/j.stem.2012.11.022
    1. Zee R. S., Yoo C. B., Pimentel D. R., Perlman D. H., Burgoyne J. R., Hou X., McComb M. E., Costello C. E., Cohen R. A. and Bachschmid M. M. (2010). Redox regulation of sirtuin-1 by S-glutathiolation. Antioxid. Redox Signal. 13, 1023-1032 10.1089/ars.2010.3251
    1. Zhang X., Yalcin S., Lee D.-F., Yeh T.-Y. J., Lee S.-M., Su J., Mungamuri S. K., Rimmelé P., Kennedy M., Sellers R. et al. (2011a). FOXO1 is an essential regulator of pluripotency in human embryonic stem cells. Nat. Cell Biol. 13, 1092-1099 10.1038/ncb2293
    1. Zhang J., Khvorostov I., Hong J. S., Oktay Y., Vergnes L., Nuebel E., Wahjudi P. N., Setoguchi K., Wang G., Do A. et al. (2011b). UCP2 regulates energy metabolism and differentiation potential of human pluripotent stem cells. EMBO J. 30, 4860-4873 10.1038/emboj.2011.401
    1. Zhang X., Rielland M., Yalcin S. and Ghaffari S. (2011c). Regulation and function of FoxO transcription factors in normal and cancer stem cells: what have we learned? Curr. Drug Targets 12, 1267-1283 10.2174/138945011796150325
    1. Zhang J., Nuebel E., Daley G. Q., Koehler C. M. and Teitell M. A. (2012). Metabolic regulation in pluripotent stem cells during reprogramming and self-renewal. Cell Stem Cell 11, 589-595 10.1016/j.stem.2012.10.005
    1. Zhang X., Campreciós G., Rimmelé P., Liang R., Yalcin S., Mungamuri S. K., Barminko J., D'Escamard V., Baron M. H., Brugnara C. et al. (2014). FOXO3-mTOR metabolic cooperation in the regulation of erythroid cell maturation and homeostasis. Am. J. Hematol. 89, 954-963 10.1002/ajh.23786
    1. Zhou W., Choi M., Margineantu D., Margaretha L., Hesson J., Cavanaugh C., Blau C. A., Horwitz M. S., Hockenbery D., Ware C. et al. (2012). HIF1α induced switch from bivalent to exclusively glycolytic metabolism during ESC-to-EpiSC/hESC transition. EMBO J. 31, 2103-2116 10.1038/emboj.2012.71
    1. Zhu S., Li W., Zhou H., Wei W., Ambasudhan R., Lin T., Kim J., Zhang K. and Ding S. (2010). Reprogramming of human primary somatic cells by OCT4 and chemical compounds. Cell Stem Cell 7, 651-655 10.1016/j.stem.2010.11.015
    1. Zou G.-M., Luo M.-H., Reed A., Kelley M. R. and Yoder M. C. (2007). Ape1 regulates hematopoietic differentiation of embryonic stem cells through its redox functional domain. Blood 109, 1917-1922 10.1182/blood-2006-08-044172

Source: PubMed

3
Subskrybuj