Regulation of catalase expression in healthy and cancerous cells

https://doi.org/10.1016/j.freeradbiomed.2015.06.017Get rights and content

Highlights

  • Multiple transcription factors control the expression of catalase.

  • Multiple mechanisms of regulation are involved in catalase expression.

  • Catalase expression is markedly altered and variable in tumors

Abstract

Catalase is an important antioxidant enzyme that dismutates hydrogen peroxide into water and molecular oxygen. The catalase gene has all the characteristics of a housekeeping gene (no TATA box, no initiator element sequence, high GC content in promoter) and a core promoter that is highly conserved among species. We demonstrate in this review that within this core promoter, the presence of DNA binding sites for transcription factors, such as NF-Y and Sp1, plays an essential role in the positive regulation of catalase expression. Additional transcription factors, such as FoxO3a, are also involved in this regulatory process. There is strong evidence that the protein Akt/PKB in the PI3K signaling pathway plays a major role in the expression of catalase by modulating the activity of FoxO3a. Over the past decade, other transcription factors (PPARγ, Oct-1, etc.), as well as genetic, epigenetic, and posttranscriptional processes, have emerged as crucial contributors to the regulation of catalase expression. Altered expression levels of catalase have been reported in cancer tissues compared to their normal counterparts. Deciphering the molecular mechanisms that regulate catalase expression could, therefore, be of crucial importance for the future development of pro-oxidant cancer chemotherapy.

Introduction

The human catalase gene is located on the short arm of chromosome 11 [1] and has all the characteristics of a housekeeping gene, with no TATA box, no initiator element sequence, and high GC content in the promoter [2]. The complete genomic DNA coding sequence for catalase has 32,420 bp and contains 12 introns and 13 exons, generating an mRNA of 2287 bp [2] encoding a single protein of 526 amino acids. The enzyme (EC 1.11.1.6), first described by Loew more than 100 years ago [3], is a homotetramer in which each monomer (62.5 kDa) contains a heme b group responsible for the enzymatic activity [4]. The human catalase belongs to the family of typical catalases [5], which predominantly catalyze the dismutation of hydrogen peroxide (H2O2) into water and molecular oxygen. In addition to its dominant “catalatic” activity (decomposition of H2O2), catalase can also decompose peroxynitrite [6], [7], [8], oxidize nitric oxide to nitrogen dioxide [9], and exhibit marginal peroxidase (i.e., oxidation of organic substrates with concomitant reduction of a peroxide ) [10] as well as low oxidase activity (O2-dependent oxidation of organic substrates) [11].

In metazoans, catalase is expressed in all major body organs, especially in the liver, kidney, and erythrocytes [12], [13], where it plays an essential role in cell defense against oxidative stress [14]. The enzyme is mainly located in peroxisomes [15], [16], but a functional catalase has also been detected in the cytoplasm [17], [18], in the mitochondria of rat cardiomyocytes [19], and on the cytoplasmic membrane of human cancer cells [20].

Altered catalase expression has been associated with several diseases. For example, certain polymorphisms in the catalase gene have been described in diabetes, hypertension, vitiligo, Alzheimer disease, and acatalasemia, leading to decreased catalase activities [21], [22].

Catalase is frequently downregulated in human and rodent tumor tissues compared to normal tissues of the same origin [23], [24], [25], [26], [27], [28], [29], [30], [31], [32], [33], [34], [35]. The low levels of catalase expression correlate with a high production of H2O2, which is involved in the activation of signaling pathways to induce proliferation, migration, and invasion in cancer cells [36], [37], [38]. We have previously reported an important decrease in catalase activity in both human and murine cancer cells [39], [40]. These observations are consistent with a study by Sun et al., who showed that immortalization and transformation of mouse liver cells with SV40 (simian virus 40) resulted in a decrease in catalase expression that contributed to oncogenesis by increasing reactive oxygen species (ROS) levels in transformed cells [41].

Conversely, catalase levels increase in rat hepatocytes and astrocytes as well as in Chinese hamster V79 fibroblasts after short-duration exposure to H2O2 [42], [43], [44]. However, short-term exposures to oxidants (i.e., hydrogen peroxide, menadione, tert-butyl hydroperoxide) fail to induce catalase protein levels in MCF-7 breast cancer cells, MRC-9 normal lung fibroblasts, or PC12 rat pheochromocytoma cells [45], [46], [47]. The expression of other H2O2-degrading enzymes (i.e., glutathione peroxidase) is induced and may counteract the increased ROS level in MCF-7 after exposure to oxidants [45]. It is noticed that the change in catalase expression, after short-term H2O2 exposure, would be influenced by several factors: the exposure time, the H2O2 concentration, the basal antioxidant enzyme capacity of the cells, and the cellular model used.

Moreover, in patients suffering from mesothelioma and in rat glioma cells, catalase protein levels are increased, conferring cellular protection against epirubicin and ionizing radiation (137Cs γ-rays), respectively [48], [49]. Increased catalase expression has been observed in tumors from patients with gastric carcinoma, skin cancer, and chronic myeloid leukemia [50], [51], [52], [53] and in human HL-60 cancer cells rendered resistant to chronic exposure to H2O2 [54], [55], [56]. This high catalase expression has also been observed in several human cancer cell lines (e.g., gastric, oral, pancreatic, bladder) exposed to cisplatin [57], ascorbic acid [58], bleomycin [59], gemcitabine [60], mitomycin C [61], hormonal therapy [62], and ionizing radiation [63].

The importance of catalase for human life is illustrated by the diseases that are associated with mutations of its gene. For example, acatalasemia is an autosomally inherited deficiency of erythrocyte catalase due to guanine-to-adenine substitution (Japanese type A), threonine deletion (Japanese type B), or guanine–adenine insertion (Hungarian type) [64], [65], [66]. Acatalasemia is characterized by a specific catalatic activity of less than 5% compared to normal rates; it is still rare and usually benign but can sometimes be problematic, resulting in oral gangrene ulceration for Japanese patients or in essential hypertension [67], [68], [69]. Catalase polymorphism has also been associated with the occurrence of diabetes, vitiligo, or Alzheimer disease [21], [22]. Regarding catalase downregulation, it should be noted that catalase-deficient mice are viable and fertile [70]. They develop normally with a normal hematological profile, but after trauma the mitochondria show defects in oxidative phosphorylation. This phenotype could be explained by the presence of other H2O2-degrading enzymes such as glutathione peroxidases and peroxiredoxins. On the other hand, the enzyme was also overexpressed in mice. Mitochondrial catalase overexpression in mice enables a life-span increase of 20% [71]. In these mice, the mitochondrial deletions are reduced; they prevent heart disease and the onset of cataracts.

As previously cited, the first function assigned to catalase is the transformation of hydrogen peroxide into oxygen and water (2 H2O2 → 2 H2O + O2). It thus plays an important role in defending cells against oxidative damage by degrading hydrogen peroxide. However, increasing evidence suggests that catalase is also involved in various other processes.

Indeed, ROS are able to activate various signaling pathways, such as that of mitogen-activated protein kinase (MAPK) to increase the capacity for proliferation, migration, and invasion [38]. Catalase can modulate the growth rate by various mechanisms, the first obviously being its ability to detoxify H2O2. The second is its ability to bind and protect certain proteins from potential oxidative damage, such as Grb2 and SHP2, involved in integrin pathways, which in turn are themselves involved in the processes of proliferation and migration [72], [73]. As shown by many reports, catalase and mitochondrial superoxide dismutase control cell growth and migration processes in cancer cells [74], [75], [76], [77].

Surprisingly, when exposed to UVB rays, catalase can also produce ROS via NADH oxidation [78]. In addition, overexpression of catalase protects cells against DNA damage induced by UVB and X-rays [79], [80].

Catalase, as catalase-peroxidase, also possesses oxidase activity [11]. The oxidase activity requires oxygen in addition to electron donors as it is heme-dependent. Phenols (benzene derivatives), alcohols, aryl amines, and other carcinogens are substrates and/or inhibitors of oxidase function. Certain metabolites, when metabolized by catalase, are active as indole and the neurotransmitter β-phenethylamine [11]. Thus, catalase may also have additional roles such as the detoxification or activation of toxic and anti-tumor compounds. We have explored a potential role of catalase during the acquisition of cancer cell resistance to chemotherapeutic agents. To this end, we overexpressed human catalase in MCF-7 cells, a human-derived breast cancer cell line. No particular resistance against conventional chemotherapies such as doxorubicin, cisplatin, or paclitaxel was observed in cells overexpressing catalase, but cells were more resistant to the pro-oxidant combination ascorbate and menadione [74]. This association generates a redox cycling that induces ROS production (mainly H2O2) and exhibits a strong preferential ability to kill cancer cells [39], [40], [74].

Tumor cells frequently produce large amounts of reactive oxygen species [81]. This can be explained by the presence of mitochondrial defects and a decreased expression of antioxidant enzymes such as catalase or manganese superoxide dismutase [82], [83]. Instead of the phenotype observed in catalase-knockout mice and acatalasemic patients, the expression of other antioxidant enzymes is also altered in cancer cells [83]. ROS play an important role in tumorigenesis and tumor progression by inducing DNA mutations and genomic instability. On the other hand, high levels of ROS can induce cell death and alterations in redox regulation and redox signaling conducted for the development of potential anti-cancer therapies [84], [85], [86], [87].

The structure, the mechanism of enzyme activity, and the phylogeny of catalase have already been extensively reviewed [5], [88], [89], [90], [91], [92]. However, the molecular mechanisms controlling catalase expression are still poorly understood, as are those explaining the altered expression of catalase in cancer cells. Therefore, the aim of the following sections is to review current knowledge regarding the various mechanisms known to regulate catalase expression. To this end, the critical roles played by Sp1 (specificity protein 1), NF-Y (nuclear factor Y), FoxO (Forkhead box protein O), and other transcription factors such C/EBP-β (CCAAT-enhancer-binding protein β), PPARγ (peroxisome proliferator-activated receptor γ), and Oct-1 (POU2F1), as well as the pathways regulating their activities in normal and cancer cells, will be considered. Regulation at other levels, including genetic, epigenetic, and posttranscriptional modifications, will also be discussed.

Section snippets

Transcriptional regulation

The human catalase gene was first described in 1986 [2]; rat and mouse catalase genes were isolated and characterized in the next decade [93], [94]. Catalase expression is predominantly regulated at the level of transcription by transcription factors that induce or repress the transcriptional activity of human and rodent catalase promoters. Fig. 1 shows these transcription factors, specifying both the species and the exact position of the binding sites as obtained from the literature. Moreover,

Other processes regulating catalase expression

Genetic alterations and epigenetic, posttranscriptional, and posttranslational modifications have also been discussed as regulators of the transcription of the human catalase gene and are summarized in Table 2.

Developing strategies to modulate catalase protein levels

Altered expression of catalase has been associated with several diseases [21], [22]. In the particular field of cancer pathology, increased and decreased expression of catalase have both been reported.

In this context, a large body of evidence indicates that cancer cells are frequently more sensitive to oxidative stress because of their low levels of antioxidant enzymes (catalase, glutathione peroxidase, superoxide dismutase, etc.) [23], [24], [25], [26], [27], [28], [29], [30], [31], [32], [33]

Conclusions

In this review we have shown that the promoter region of mammalian catalases was highly conserved during evolutionary history allowing efficient binding of transcription factors NF-Y, Sp1, and WT1/Egr in the core region. Various other factors, mainly the Fox family members regulated by the Akt/PKB signaling pathway, possess conserved binding sites in vertebrate catalase promoters. Speciation events during evolution caused only small variations within these regions, conserving heme catalase as

References (211)

  • R. Radi et al.

    Detection of catalase in rat heart mitochondria

    J. Biol. Chem.

    (1991)
  • G. Guner et al.

    Evaluation of some antioxidant enzymes in lung carcinoma tissue

    Cancer Lett.

    (1996)
  • K.A. Kwei et al.

    Transcriptional repression of catalase in mouse skin tumor progression

    Neoplasia

    (2004)
  • R. Subapriya et al.

    Oxidant–antioxidant status in patients with oral squamous cell carcinomas at different intraoral sites

    Clin. Biochem.

    (2002)
  • D.G. Yoo et al.

    Alteration of APE1/ref-1 expression in non-small cell lung cancer: the implications of impaired extracellular superoxide dismutase and catalase antioxidant systems

    Lung Cancer

    (2008)
  • S. Sen et al.

    Maintenance of higher H2O2 levels, and its mechanism of action to induce growth in breast cancer cells: important roles of bioactive catalase and PP2A

    Free Radic. Biol. Med.

    (2012)
  • E. Rohrdanz et al.

    Alterations of antioxidant enzyme expression in response to hydrogen peroxide

    Free Radic. Biol. Med.

    (1998)
  • E. Rohrdanz et al.

    The influence of oxidative stress on catalase and MnSOD gene transcription in astrocytes

    Brain Res.

    (2001)
  • P. Sen et al.

    Enhancement of catalase activity by repetitive low-grade H2O2 exposures protects fibroblasts from subsequent stress-induced apoptosis

    Mutat. Res.

    (2003)
  • S. Desaint et al.

    Mammalian antioxidant defenses are not inducible by H2O2

    J. Biol. Chem.

    (2004)
  • V.D. Nair et al.

    Early single cell bifurcation of pro- and antiapoptotic states during oxidative stress

    J. Biol. Chem.

    (2004)
  • P.S. Smith et al.

    Inhibiting catalase activity sensitizes 36B10 rat glioma cells to oxidative stress

    Free Radic. Biol. Med.

    (2007)
  • T.S. Hwang et al.

    Differential expression of manganese superoxide dismutase, copper/zinc superoxide dismutase, and catalase in gastric adenocarcinoma and normal gastric mucosa

    Eur. J. Surg. Oncol.

    (2007)
  • I. Kasugai et al.

    High production of catalase in hydrogen peroxide-resistant human leukemia HL-60 cell lines

    Leuk. Res.

    (1992)
  • H. Xu et al.

    Concentration-dependent collateral sensitivity of cisplatin-resistant gastric cancer cell sublines

    Biochem. Biophys. Res. Commun.

    (2005)
  • L. Goth et al.

    Kalmar T. A novel catalase mutation (a GA insertion) causes the Hungarian type of acatalasemia

    Blood Cells Mol. Dis.

    (2000)
  • A. Hirono et al.

    A novel human catalase mutation (358 T-->del) causing Japanese-type acatalasemia

    Blood Cells Mol. Dis.

    (1995)
  • J.K. Wen et al.

    Molecular analysis of human acatalasemia: identification of a splicing mutation

    J. Mol. Biol.

    (1990)
  • S. Takahara

    Progressive oral gangrene probably due to lack of catalase in the blood (acatalasaemia); report of nine cases

    Lancet

    (1952)
  • Y.S. Ho et al.

    Mice lacking catalase develop normally but show differential sensitivity to oxidant tissue injury

    J. Biol. Chem.

    (2004)
  • S. Yano et al.

    SHP2 binds catalase and acquires a hydrogen peroxide-resistant phosphatase activity via integrin-signaling

    FEBS Lett.

    (2004)
  • S. Yano et al.

    Catalase binds Grb2 in tumor cells when stimulated with serum or ligands for integrin receptors

    Free Radic. Biol. Med.

    (2004)
  • C. Glorieux et al.

    Catalase overexpression in mammary cancer cells leads to a less aggressive phenotype and an altered response to chemotherapy

    Biochem. Pharmacol.

    (2011)
  • Z. Liu et al.

    Manganese superoxide dismutase induces migration and invasion of tongue squamous cell carcinoma via H2O2-dependent Snail signaling

    Free Radic. Biol. Med.

    (2012)
  • Y. Hu et al.

    Mitochondrial manganese-superoxide dismutase expression in ovarian cancer: role in cell proliferation and response to oxidative stress

    J. Biol. Chem.

    (2005)
  • A. Glasauer et al.

    Targeting antioxidants for cancer therapy

    Biochem. Pharmacol.

    (2014)
  • H.N. Kirkman et al.

    Mechanisms of protection of catalase by NADPH: kinetics and stoichiometry

    J. Biol. Chem.

    (1999)
  • C.D. Putnam et al.

    Active and inhibited human catalase structures: ligand and NADPH binding and catalytic mechanism

    J. Mol. Biol.

    (2000)
  • M. Zamocky et al.

    Understanding the structure and function of catalases: clues from molecular evolution and in vitro mutagenesis

    Prog. Biophys. Mol. Biol.

    (1999)
  • M. Zamocky et al.

    Evolution of structure and function of Class I peroxidases

    Arch. Biochem. Biophys.

    (2010)
  • H. Nakashima et al.

    Isolation and characterization of the rat catalase-encoding gene

    Gene

    (1990)
  • D.L. Reimer et al.

    Complete cDNA and 5′ genomic sequences and multilevel regulation of the mouse catalase gene.

    Genomics

    (1994)
  • L. Li et al.

    Positive regulation of human alpha 1 (I) collagen promoter activity by transcription factor Sp1

    Gene

    (1995)
  • K.C. Brown et al.

    Nicotine induces the up-regulation of the alpha7-nicotinic receptor (alpha7-nAChR) in human squamous cell lung cancer cells via the Sp1/GATA protein pathway

    J. Biol. Chem.

    (2013)
  • R. Huber et al.

    Multiple Sp1 sites efficiently drive transcription of the TATA-less promoter of the human glypican 3 (GPC3) gene

    Gene

    (1998)
  • P. Wieacker et al.

    Assignment of the gene coding for human catalase to the short arm of chromosome 11

    Ann. Genet.

    (1980)
  • F. Quan et al.

    Isolation and characterization of the human catalase gene

    Nucleic Acids Res.

    (1986)
  • O. Loew

    A new enzyme of general occurrence in organisms

    Science

    (1900)
  • R.A. Nagem et al.

    Crystallization and preliminary X-ray diffraction studies of human catalase

    Acta Crystallogr. D Biol. Crystallogr

    (1999)
  • M. Zamocky et al.

    Evolution of catalases from bacteria to humans

    Antioxid. Redox Signaling

    (2008)
  • Cited by (0)

    View full text