Sei sulla pagina 1di 14

Biol. Chem.

2017; 398(10): 1095–1108

Review

Christophe Glorieuxa and Pedro Buc Calderon*

Catalase, a remarkable enzyme: targeting


the oldest antioxidant enzyme to find a new
cancer treatment approach
DOI 10.1515/hsz-2017-0131
Received March 13, 2017; accepted April 4, 2017; previously published
Historical view
online April 6, 2017
The origin of catalase (EC 1.11.1.6), the enzyme that metab-
Abstract: This review is centered on the antioxidant enzyme olizes H2O2 but also reacts with a multitude of other sub-
catalase and will present different aspects of this particu- strates, can be traced back to the 19th century (Figure 1)
lar protein. Among them: historical discovery, biological when Thénard discovered H2O2 and suspected that its tissue
functions, types of catalases and recent data with regard degradation in living organisms was the result of a ‘special’
to molecular mechanisms regulating its expression. The substance activity (Thénard, 1811). Schönbein showed that
main goal is to understand the biological consequences of a ‘ferment’ can detoxify H2O2 (Schönbein, 1863) and later
chronic exposure of cells to hydrogen peroxide leading to on Loew gave the name ‘catalase’ to the enzyme converting
cellular adaptation. Such issues are of the utmost impor- H2O2 into water and oxygen (Loew, 1900). Loew found the
tance with potential therapeutic extrapolation for various presence of this enzyme in many organisms ranging from
pathologies. Catalase is a key enzyme in the metabolism of plants to mammals. In the 1920s, Warburg and c­ oworkers
H2O2 and reactive nitrogen species, and its expression and demonstrated, by using cyanide as enzyme inhibitor, that
localization is markedly altered in tumors. The molecular the active site of catalase contains an iron atom (Warburg,
mechanisms regulating the expression of catalase, the old- 1923). Furthermore, Stern showed that the hemin group of
est known and first discovered antioxidant enzyme, are the enzyme can react with compounds such as cyanides,
not completely elucidated. As cancer cells are character- sulfides, fluorides and he also showed that the enzyme
ized by an increased production of reactive oxygen spe- active group had a ferric complex identical to the protopor-
cies (ROS) and a rather altered expression of antioxidant phyrin found in the hemoglobin of red blood cells (Stern,
enzymes, these characteristics represent an advantage in 1937). Thereafter, Sumner and Dounce (1937) purified and
terms of cell proliferation. Meanwhile, they render cancer crystallized bovine catalase.
cells particularly sensitive to an oxidant insult. In this con- The advances in biochemistry facilitated further elu-
text, targeting the redox status of cancer cells by modulat- cidation about the mechanisms of the ‘catalase’ enzy-
ing catalase expression is emerging as a novel approach to matic reaction. Thus, the work carried out by Chance’s
potentiate chemotherapy. lab led to the discovery of the formation of the ‘Com-
pound I’ that occurred during the reaction between cata-
Keywords: antioxidant enzyme; cancer; catalase; catalase
lase and the first molecule of H2O2. A few years later, he
regulation; hydrogen peroxide; pro-oxidant therapy.
also discovered the ‘compounds II and III’ (Chance, 1948,
1949; Chance et al., 1952). In the 1960s, it was elucidated
a
Present address: Sun Yat-Sen University Cancer Center, State Key the role of key residues in the active site of the enzyme,
Laboratory of Oncology in South China, Collaborative Innovation
such as the distal histidine, and their importance for the
Center of Cancer Medicine, 510275 Guangzhou, China.
*Corresponding author: Pedro Buc Calderon, Metabolism and stabilization of the tertiary structure was discussed by
Nutrition Research Group, Louvain Drug Research Institute, Nakatani (1961). In 1970, catalase Compound I was iden-
Université Catholique de Louvain, Avenue E. Mounier 73, 1200 tified in intact eukaryotic cells, proving the existence of
Brussels, Belgium; and Facultad de Ciencias de la Salud, H2O2 in normal aerobic metabolism (Sies and Chance,
Universidad Arturo Prat, 1100000 Iquique, Chile, e-mail: pedro.
1970). Kirkman and Gaetani reported that NADPH was
buccalderon@uclouvain.be
Christophe Glorieux: Metabolism and Nutrition Research Group,
the enzyme cofactor bound to catalase (1984), which
Louvain Drug Research Institute, Université Catholique de Louvain, was further confirmed following X-ray analysis of cata-
Avenue E. Mounier 73, 1200 Brussels, Belgium lase structures (Fita and Rossmann, 1985a). Finally,

Brought to you by | Bibliothèques de l'UCL


Authenticated
Download Date | 9/8/17 1:29 PM
1096      C. Glorieux and P. B. Calderon: Targeting catalase as new cancer treatment approach

Figure 1: Time line of catalase main discoveries and findings.


Modified from Sies (2017).

recombinant phage clones containing the human cata- catalase-peroxidases, whereas the third group contains
lase gene were isolated and characterized (Quan et  al., (non-heme) manganese catalases (Zamocky and Koller,
1986) and the use of molecular biology techniques have 1999; Zamocky et al., 2008, 2010).
led to major advances in the understanding of catalase
regulation mechanisms and the role that this enzyme
Typical catalases
could play in numerous biological processes.
In this context, Nenoi et al. (2001) reported that in the
The members of this largest group are found in aerobically
upstream catalase promoter, there is a region containing
respiring organisms. In contrast, in anaerobic ­bacteria,
CCAAT and GGGCGG boxes where transcription factors
catalases proteins are generally not expressed. Most of
like Nuclear factor Y (NF-Y) and Specificity protein 1 (Sp1)
these catalases are homotetramers, between 200 and
can bind to the catalase promoter regulating the transcrip-
340 kDa in size and contain four prosthetic groups. In the
tional activation of the human catalase gene. Recently,
majority of true catalases, a ferric protoporphyrin IX was
Glorieux et al. (2016a) identified a new regulatory region
found in the active center (namely heme b, similar to the
in the human catalase promoter, in which chromatin
prosthetic group of human hemoglobin). Some variants
remodeling is required to regulate catalase expression by
do exist such as ‘heme d’ groups which can reside in some
retinoic acid receptor alpha (RARα) and JunB transcrip-
typical catalases.
tion factors. This novel regulatory mechanism is involved
Following phylogenetic analyses the typical cata-
during cancer cells adaptation to chronic exposure to H2O2
lases can also be divided into three main clades. Clade 1
and may have therapeutic consequences for various dis-
contains bacterial, algal and plant catalases with small-
eases and metabolic disorders.
sub­unit size (55–69 kDa) using heme b as the prosthetic
group. Clade 2 regroups bacterial and fungal catalases
and has a large subunit size (75–84  kDa), with heme d
Types of catalases as the prosthetic group and an additional ‘flavodoxin-
like’ domain. Clade 3 is the most abundant subfamily;
With the increasing number of complete sequences availa- catalases from this subgroup are found in archaebacte-
ble, distinct homologies were detected and catalases came rial, fungi, protists, plants and animals. The human cata-
to be classified in three groups based on their structure and lase belongs to this clade and is characterized by a small
function. The first and the second group contain heme- subunit (62 kDa), with heme b as its prosthetic group and
containing enzymes, namely typical or true catalases and NADPH as cofactor.

Brought to you by | Bibliothèques de l'UCL


Authenticated
Download Date | 9/8/17 1:29 PM
C. Glorieux and P. B. Calderon: Targeting catalase as new cancer treatment approach      1097

Catalase-peroxidases β-barrel domain, the connection domain and the zone


neighboring the distal histidine are highly conserved.
These proteins have been found in fungi, archeobacteria The α-helical domain is moderately conserved between
and bacteria. Their molecular weight varies between 120 species and some typical catalases do not bind the cofac-
and 340 kDa and they are generally homodimers. The cat- tor NADPH.
alase activity (degrading hydrogen peroxide) is less effi- The investigation of inhibitory mechanisms by
cient than in typical catalases but catalase-peroxidases which cyanide and 3-amino-1,2,4-triazole (ATA) inhibit
have a better affinity for their substrate H2O2. Catalase-per- human catalase allowed to understand the enzyme activ-
oxidases are also significantly more sensitive than typical ity (Putnam et  al., 2000). Cyanide nitrogen blocks heme
catalases to inactivation by pH and temperature. The access to other potential ligands. It interacts with the
well-known horseradish peroxidase, currently employed distal histidine and an asparagine residue suggesting that
in immunoblotting experiments, is one example of it competes with hydrogen peroxide for heme binding.
catalase-peroxidase. Meanwhile, ATA interacts with the distal histidine leading
to an adduct formation and thereby blocks the catalase
reaction.
Manganese catalases

These enzymes have been found exclusively in bacteria.


Manganese catalases utilize two manganese ions in the Mechanism of ‘catalase’ reaction
active site, they can form oligomeric structures measuring
between 170 and 210 kDa and have no significant homol- During the enzymatic reaction leading to H2O2 destruction,
ogy with either typical catalases or catalase-peroxidases. catalase is first oxidized to a hypervalent iron intermedi-
The catalytic reaction is completely different to other ate, known as compound I (Cpd I), which is then reduced
types of catalases. The dimanganese core is equally stable back to the resting state by a second H2O2 molecule.
Mn2+–Mn2+ or Mn3+–Mn3+. Like typical catalases, the cata-
1st reaction: Enz (Porf-FeIII ) + H2 O2 →
lase reaction occurs in two-step.
Cpd I (Porf •+ -FeIV =O) + H2O
1st step: H2O2 + Mn −Mn (2H ) → Mn −Mn + 2 H2O
2+ 2+ + 3+ 3+
2nd reaction: Cpd I (Porf •+ -FeIV =O) + H2O2 →
2nd step: H2O2 + Mn −Mn → Mn −Mn (2H ) + O2
3+ 3+ 2+ 2+ +
Enz (Porf-FeIII ) + H2 O + O2
Sum of reactions: H2 O2 + H2 O2 → 2 H2O + O2 Sum of reactions: H2 O2 + H2O2 → 2 H2O + O2

The first reaction is characterized by the oxidation of the


heme protein by a single H2O2 molecule leading to the
Structure of human catalase formation of Cpd I, an oxoferryl porphyrin cation radical
(Jones and Dunford, 2005). Once Cpd I is formed, it reacts
Human catalase contains four identical subunits of rapidly with a second molecule of H2O2 to generate H2O
62 kDa, each subunit containing four distinct domains and and O2 in a two-electron redox process. This second reac-
one prosthetic heme group (Nagem et al., 1999; Zamocky tion is particularly efficient in some catalases compared
and Koller, 1999; Putnam et al., 2000). The four domains to other heme proteins such as myoglobin (Matsui et al.,
include: (1) a N-terminal arm which contains a distal histi- 1999). Labeling studies have shown that both H2O and
dine, an essential amino acid for the catalase reaction; (2) O2 molecules are formed from the same molecule of H2O2
a β-barrel domain that contains eight β-barrels arranged (Vlasits et al., 2007). Fita and Rossmann (1985b) proposed
in an antiparallel fashion with six α-helical insertions, that two molecules of H2O2 were sequentially transferred
conferring the hydrophobic core of the protein neces- to the oxoferryl group of Cpd I, where the distal histidine
sary for the tri-dimensional structure of the enzyme; (3) residue plays a role of acid-base catalyst. Although muta-
a connection domain which contains the tyrosine residue tion of distal histidine suppresses the ability to form Cpd I
that binds the heme group; and finally (4) an α-helical (Nakatani, 1961), some authors claimed that reaction
domains, which is important for NADPH binding. occurs as a direct mechanism and histidine does not play
Although amino acid sequences do not have high a crucial role in catalysis (Kato et al., 2004).
identities between all typical catalases, the tridimensional In the presence of one-electron donors (such as
structure is highly conserved. The tertiary structure of the phenols, ferrocyanide, salicylic acid, NO, superoxide

Brought to you by | Bibliothèques de l'UCL


Authenticated
Download Date | 9/8/17 1:29 PM
1098      C. Glorieux and P. B. Calderon: Targeting catalase as new cancer treatment approach

anions) and low H2O2 concentrations, Cpd I may undergo signal (Purdue and Lazarow, 1996). Proteins bearing these
a one-electron reduction towards the inactive compound signals are recognized by the PTS1 receptor, called PEX5p
II (Cpd II) intermediate, which transforms back to the for humans. Some diseases related to defects in peroxi-
resting state through another one-electron reduction step some biogenesis, such as Zellweger syndrome, are char-
(reviewed in Bauer, 2015). Both Cpd I and Cpd II are gen- acterized by mutations in the PEX5p receptor and cellular
erally described as an oxoferryl-heme species (Rovira, H2O2 overproduction due to a catalase default import in
2005). In the case of Cpd I, the porphyrin bears a cation peroxisomes (Wanders et al., 1984). This syndrome is most
radical (O=FeIV-heme˙+), while Cpd II lacks the porphy- commonly called ‘the syndrome of empty peroxisome’. It
rin cation radical (O=FeIV-heme). Thus, Cpd II is best has been shown that overexpressing receptor mutants in
described as a hydroxoferryl bond (HO-FeIV-heme) instead PEX5-deficient CHO cells, drastically reduce the import of
of the traditional oxoferryl species, consistent with the proteins such as catalase (Shimozawa et al., 1999). Some
fact that a proton is released upon conversion of Cpd I to authors have observed that catalase with the SKL signal
Cpd II: and not KANL had a better import capacity and could be
transported into the peroxisome even in the case of muta-
Cpd I + AH (one-electron donor) →
tions in the PTS1 receptor (Koepke et al., 2007). It has been
Cpd II(Por-FeIV = O and Por-FeIV -OH)
shown that PEX5p recognizes catalase which is already
folded and interacts with other receptors such as PEX13p
At higher H2O2 concentrations, NADPH prevents the gen-
for proper import into the peroxisome (Otera and Fujiki,
eration of Cpd II by participating in a two-electron reduc-
2012). Studies are underway to validate, through clinical
tion process (Kirkman et  al., 1999). In the presence of
trials, whether catalase SKL has a therapeutic potential for
another one-electron donor, Cpd II will return to a resting
diseases with peroxisome biogenesis disorders. Recently,
state. But in presence of a H2O2 molecule, Cpd II will be
it has been reported that PEX19p, an essential protein for
transformed back to compound III (Cpd III), an inactive
peroxisome biogenesis, interacts with Valosin-contain-
intermediate (Gabdoulline et  al., 2003). In this interme-
ing protein (VCP) and regulates the catalase cytoplasmic
diate state, iron is at an oxyferrous state (O2-FeII-heme).
localization, a potential feedback mechanism modulating
Then, Cpd III goes back to a resting state or leads to the
H2O2 levels (Murakami et al., 2013).
inactivation of the catalase.
Interestingly, the existence of a cytosolic catalase,
Cpd II + AH → Enz (Por-FeIII ) resting state in its active tetrameric conformation, has been reported
Cpd II + H2O2 → Cpd III (Por-FeII -O2 ) (­Middelkoop et al., 1993) with varied percentage from one
Cpd III → Enz (Por-FeIII ) or inactivation cell type to another, and different function to peroxisome
catalase. Indeed, catalase may bind cytosolic proteins such
as Grb2 and SHP2, to protect them from potential oxidative
damage (Yano et al., 2004a,b). These proteins are linked to

Cellular and tissue catalase the membrane by a pleckstrin homology (PH) domain, so
it is common to find catalase in fractions including mem-
distribution brane proteins and membrane-associated proteins. In this
context, it has been shown that catalase can be localized
Regarding catalase localization within the cell, it should at the cytoplasmic membrane, specifically at the surface
be noted that catalase is mainly located in peroxisomes of cancer cells (Bauer, 2012). This locally high expression
because it contains a sequence signal recognized by some of catalase on the membrane of tumor cells is in line with
peroxisome receptors. Contrary to mitochondria, proteins the findings by Deichman’s group that showed that tumor
located within peroxisomes are all of nuclear origin and progression in vivo is dependent on increased resistance
should be imported. Indeed, it is generally accepted that towards exogenous H2O2 (Deichman, 2000, 2002). Further-
catalase monomers are imported into peroxisomes where more, localized expression of catalase on the membrane
tetramerization and heme addition occurs (Lazarow and of tumor cells is not in disagreement with the finding of
De Duve, 1973). The apoprotein (monomer) enters into the lower total catalase concentration in malignant cells, as
peroxisome by a peroxisome-targeting signal sequence the membrane comprises a minority of the total cellular
(PTS) present on the carboxy-terminal tail of catalase. The material.
most common targeting signal is the SKL (serine-lysine- These observations open the way for new anti-can-
leucine) but catalase is characterized by a different signal, cer therapies targeting catalase with specific antibodies
namely the KANL (lysine-alanine-asparagine-leucine) (Bauer, 2012; Bauer and Motz, 2016) or exogenous singlet

Brought to you by | Bibliothèques de l'UCL


Authenticated
Download Date | 9/8/17 1:29 PM
C. Glorieux and P. B. Calderon: Targeting catalase as new cancer treatment approach      1099

oxygen (Riethmüller et al., 2015; Bauer and Graves, 2016) et al., 2001). A discrete balance between oxidation of NO
to induce apoptosis by reactivation of intercellular HOCl by compound I of catalase and inhibition of catalase by
and/or NO/peroxynitrite signaling after catalase inhibi- NO through formation of a CAT-FeIIINO complex has been
tion or inactivation. Furthermore, the modulation of the reported (Brown, 1995). Catalase also exhibits low oxidase
intracellular NO concentration has been shown to lead to activity (O2-dependent oxidation of organic substrates)
the generation of cell-derived singlet oxygen that inacti- (Vetrano et al., 2005). Thus, catalase may also have addi-
vates tumor cell protective catalase and reactivates inter- tional roles such as the detoxification or activation of toxic
cellular ROS/RNS-mediated apoptosis-inducing signaling and anti-tumor compounds. For instance, catalase has
(Bauer, 2015; Scheit and Bauer, 2015). been detected in mouse oocytes most likely to protect the
In addition to classical intracellular catalase and genome from oxidative damage during meiotic matura-
membrane-associated catalase of tumor cells (Heinzel- tion (Park et al., 2016).
mann and Bauer, 2010; Böhm et al., 2015), the release of Within this framework, several studies have shown
soluble catalase from tumor cells has also been reported a change in catalase expression in cancer cells became
(Sandstrom and Buttke, 1993; Moran et  al., 2002; Böhm resistant to chemotherapies (Akman et  al., 1990; Kim
et al., 2015). This soluble extracellular catalase was pro- et al., 2001; Kalinina et al., 2006). Thus, a potential role
tective for the tumor cells. of catalase during the acquisition of cancer cell resistance
Finally, catalase has been also localized in the mito- to chemotherapeutic agents was explored by overexpress-
chondria of rat cardiomyocytes (Radi et al., 1991). ing the human enzyme in MCF-7 cells, a human derived
The human catalase is expressed in every organ and breast cancer cell line (Glorieux et al., 2011). No ­particular
the highest levels of activity are measured in the liver, resistance against conventional chemotherapies like
kidney and red blood cells (Winternitz and Meloy, 1908). doxorubicin, cisplatin and paclitaxel was observed in
­
In erythrocytes, a high production of H2O2 is generated cells overexpressing catalase but they were more resistant
due to oxygen transport and catalase is responsible for to the pro-oxidant effect induced by an H2O2-generating
more than 50% of the H2O2 turnover (Mueller et al., 1997). system (Glorieux et al., 2011).
In addition, catalase mitochondria overexpression
in mice enables an increase of lifespan by 20% (Schriner

Biological functions of catalase et  al., 2005). In such animals, the development of mito-
chondrial deletions was reduced and heart disease and
and related diseases the onset of cataracts were delayed.
Regarding catalase down-regulation, no particular
The first function assigned to catalase is the dismutation sensitivity was observed as catalase-deficient mice are
of H2O2 into oxygen and water without consummation of viable and fertile (Ho et al., 2004). They develop normally
endogenous reducing equivalents, an important role in with a normal hematological profile, but after trauma the
cell defense against oxidative damage by H2O2. To note mitochondria shows defects in the oxidative phosphoryla-
that H2O2 is not only toxic by its ability to form other ROS, tion. Note that humans may also be deficient in catalase, a
like hydroxyl radical through the Fenton reaction (Fenton, condition known as acatalasemia that is characterized by
1894) but as was nicely recently reviewed by Sies, H2O2 a low catalase rate, but it is still rare and usually benign
acting as a second messenger is involved in many biologi- (Goth et al., 2004).
cal processes including changes of morphology, prolifera- In this context, there are benign polymorphisms of
tion, signaling (i.e. NF-κB), apoptosis, etc (Sies, 2017). In the catalase gene for which no change of catalase expres-
addition to its dominant ‘catalatic’ activity (decomposi- sion or activity was detected (Goth et  al., 2004). They
tion of H2O2), catalase can also act in its peroxidatic mode, include single nucleotide substitutions in the promoter,
i.e. decomposition of small substrates such as methanol, 5′ untranslated region, intron 1, exon 1, exon 9 and 10
formate, azide, hydroperoxides (Sies, 1974; Chance et al., (Goth et al., 2004). To note that the catalase gene encodes
1979; Johansson and Borg, 1988), and in case of ethanol, one single protein of 526 amino acids and the single locus
it is also capable of oxidize it to acetaldehyde contributing has been mapped to chromosome 11p13 (Wieacker et al.,
to its liver metabolism (Keilin and Hartree, 1945; Thurman 1980). The length of the catalase gene is 34 kb, it contains
et al., 1972; Oshino et al., 1973). It has also been reported 12 introns and 13 exons generating a mRNA of 2286 bp
that catalase may decompose peroxynitrite (Gebicka and (Quan et al., 1986).
Didik, 2009; Heinzelmann and Bauer, 2010), oxidize Conversely, some catalase mutations provoke changes
nitric oxide to nitrite (Wink and Mitchell, 1998; Brunelli in either catalase expression or activity and may be

Brought to you by | Bibliothèques de l'UCL


Authenticated
Download Date | 9/8/17 1:29 PM
1100      C. Glorieux and P. B. Calderon: Targeting catalase as new cancer treatment approach

associated with some diseases. For instance, a higher tran- (i.e. loss of heterozygosity) of the catalase gene in non-
scriptional activity in two human cancer cell lines, namely small-cell lung cancer cells (Ludwig et  al., 1991; Fong
HepG2 and K562 cells, was observed in the case of common et al., 1994; Shipman et al., 1998) or deletion of chromo-
functional C-T substitution polymorphism in the promoter some 11p, as has been observed in children affected by
region (−262) of the human catalase gene (­Forsberg et al., Wilms’ tumor, aniridia, gonadoblastoma and retardation
2001) but no mutations have been detected in the coding (WAGR) syndrome (Dufier et al., 1981; Gregoire et al., 1983;
sequence, to our knowledge, in patients suffering from Barletta et al., 1985; Michalopoulos et al., 1985; van Hey-
cancer. Recent studies have focused on the associations ningen et al., 1985).
of catalase polymorphisms with various types of cancer
but many inconsistent results about the relationship
between the catalase gene polymorphism and cancer risk
were reported. Recently, two meta-analyses pointed out a
Regulation of catalase expression
correlation exists between this polymorphism C-262T and in cancer cells
prostate cancer (Liu et al., 2016; Wang et al., 2016).
Different catalase mutations in patients can cause It is generally accepted that the cellular maintenance of
decreased catalase activity leading to increased H2O2 con- redox homeostasis is controlled by a complex network
centrations in the blood and tissues. Table 1 shows the of antioxidant enzymes (i.e. superoxide dismutases and
polymorphism of the catalase gene in patients suffering glutathione peroxidases) whose expression is under the
from hypocatalasemia (about 50% of catalase activities) fine-tuning control of the Keap1-Nrf2  signaling pathway
or acatalasemia (less than 10% of catalase activities). (Menegon et al., 2016). Nevertheless, the molecular mech-
Depending on the mutations, such patients may be anisms regulating the expression of catalase – the oldest
subject to an increased risk of type 2 diabetes, vitiligo and known and first discovered antioxidant enzyme – are
increased blood pressure (Goth et  al., 2004). When the independent of this pathway and not totally elucidated.
mutations are located in exons as in Japanese and Hun- Therefore, the fine-tuning regulation of this enzyme
garian acatalasemia, a truncated or mutated catalase is should be prior elucidated in order to find a new approach
synthetized and functionally less active. Takahara was the to modulate the antioxidant status in cancer cells.
first to describe acatalasemia and this pathology is often Interestingly, despite the existence of diverse protec-
benign but Japanese patients can suffer from oral gan- tion mechanisms against oxidant injuries, a consensus
grenes and esophageal ulcerations (Takahara’s disease), emerged in the scientific literature about an alteration
probably promoted by H2O2 generated by phagocytic cells of redox homeostasis within tumor cells. Indeed, they
and bacterial actions (Takahara, 1952). produce large amounts of ROS that are involved in the
Decreased activity of catalase has also been observed maintenance of genetic instability favoring cancer cell
in various genetic alterations, for example, loss of alleles proliferation. Meanwhile, altered expression levels of
catalase have been reported in cancer tissues as com-
pared to their normal counterparts. Thus, as compared to
Table 1: Catalase gene polymorphisms in patients suffering from
hypocatalasemia or acatalasemia.
normal tissues of the same origin, some authors reported
an increased catalase expression in tumors (Sander
Mutations Position Region Associated disease et  al., 2003; Hwang et  al., 2007; Rainis et  al., 2007),
whereas other studies showed a catalase down-regula-
G insertion Codon 48 Exon 2 Diabetes type 2
T→A Codon 53 Exon 2 Diabetes type 2
tion (Marklund et al., 1982; Baker et al., 1997; Lauer et al.,
G→C Codon 66 Exon 2 Diabetes type 2 1999; Chung-man et  al., 2001; Cullen et  al., 2003; Kwei
C→T −773a Promoter Hypertension et  al., 2004), indicating that cancer cells are frequently
T→C Codon 388 Exon 9 Vitiligo more sensitive to an oxidative stress. For instance, we
G→A 5b Intron 4 Japanese hypocatalasemia (A) have reported an important decrease of catalase activity
T deletion Codon 134 Exon 4 Japanese hypocatalasemia (B)
in different cancer cell lines, as shown in Table 2 (Verrax
GA insertion Codon 67 Exon 2 Hungarian hypocatalasemia (A)
G insertion Codon 58 Exon 2 Hungarian hypocatalasemia (B) et  al., 2009; Beck et  al., 2011a; Glorieux et  al., 2011).
G→T 5b Intron 7 Hungarian hypocatalasemia (C) Briefly, catalase levels can vary after short treatments to
G→A Codon 354 Exon 9 Hungarian hypocatalasemia (D) H2O2 (Rohrdanz and Kahl, 1998; Rohrdanz et al., 2001; Sen
This table was adapted from Goth et al. (2004).
et al., 2003) and catalase expression is modified in cancer
a
Position from the ATG translation start. cell lines rendered resistant to chronic exposures to H2O2
b
Position from the first nucleotide in the intron sequence. (Kasugai and Yamada, 1992; Nenoi et al., 2001) or certain

Brought to you by | Bibliothèques de l'UCL


Authenticated
Download Date | 9/8/17 1:29 PM
C. Glorieux and P. B. Calderon: Targeting catalase as new cancer treatment approach      1101

Table 2: Catalase enzyme activity in cells from diver origins.

Type of cells Normal origin Cancer origin References

Mouse hepatocytes 96.45 ± 6.32 11.12 ± 4.43a,f Verrax et al. (2009)


Human leukocytes 44.55 ± 1.80 16.36 ± 3.60b,e Beck et al. (2011a)
Human mammary epithelial cells 13.14 ± 3.04c 5.44 ± 0.88d,e Glorieux et al. (2016a)

Catalase activity (U/mg proteins) in normal and cancer cell lines. aTLT mouse hepatocarcinoma cells; bK562 human CML cells; c250MK cells;
and dMCF-7 human mammary cells. Data were analyzed by unpaired t-test. ep-Value ≤ 0.01; fp-value ≤ 0.001.

chemotherapeutic compounds such as doxorubicin contributors to the negative regulation of catalase expres-
(Ramu et  al., 1984; Akman et  al., 1990; Kim et  al., 2001; sion (Glorieux et  al., 2015). Specifically, we investigated
Kalinina et  al., 2006). Although mechanisms controlling the transcriptional regulatory mechanism controlling
catalase expression have been partially elucidated, the catalase expression in human mammary cell lines. To
decreased catalase expression in cancer cells still remains this end, we have made a human breast MCF-7 cancer
an unanswered question. cell line resistant to oxidative stress, the so-called Resox
The regulation of catalase expression in cancer cells cells. These cells show decreased ROS basal levels and an
is a complex process because different levels of regula- increased activity of some antioxidant enzymes, notably
tion are thought to be involved. In a recent review, we catalase (Dejeans et al., 2012; Glorieux et al., 2015, 2016b).
discussed the different mechanisms playing a potential A novel promoter region, responsible for the regulation of
role in the regulation of its expression in both healthy catalase expression, was identified at −1518/−1226 locus
and tumor cells (Glorieux et al., 2015). They include tran- in Resox cells (Glorieux et  al., 2016a). The AP-1 family
scriptional regulation, represented by the activity of tran- member JunB and retinoic acid receptor alpha (RARα)
scription factors that induce or repress the transcriptional mediate catalase transcriptional activation and repres-
activity of catalase promoters, post-transcriptional regu- sion, respectively, by controlling chromatin remodeling
lation (mRNA stability) and post-translational modifica- through a histone deacetylases-dependent mechanism
tion (phosphorylation and ubiquitination of the protein). (Glorieux et  al., 2016a). Indeed, RARα and JunB act in
In addition, epigenetic (DNA methylation, modifications collaboration with transcription factors or other proteins
of histones) changes or genetic alterations can also be on either a closed- or an open-promoter chromatin status
involved playing a role in governing proper levels of cata- regulating the expression of catalase in breast cancer
lase activity in these cells. cells (Figure 2). Thus, cancer adaptation to oxidative
Regarding transcription it should be noted that the stress, regulated by transcriptional factors through chro-
­catalase gene has all the characteristics of a housekeeping matin remodeling appears as a new mechanism to target
gene (no TATA box, no INR sequence, high GC content in cancer cells.
promoter) and a core promoter which is highly conserved
among species (Quan et al., 1986; Nakashima et al., 1990;
Reimer et  al., 1994). In this core promoter, the presence
of DNA binding sites for transcription factors like NF-Y
Oxidative stress-based therapies
and Sp1 has an essential role in the positive regulation of against cancer
catalase expression (Nenoi et al., 2001). Additional tran-
scription factors have also been involved in this regulatory As previously mentioned, cancer cells are generally defi-
process. In fact, there is strong evidence that the protein cient in antioxidant enzymes, thus, any increase in ROS
Akt/PKB in the PI3K signaling pathway plays a major role levels would be a menace to the precarious redox balance
in the expression of catalase by modulating the activity of of cancer cells making them vulnerable to an additional
FoxO3a (Turdi et al., 2007; Venkatesan et al., 2007; Akca oxidative stress. Given that weakness, the loss of redox
et al., 2013). Therefore, targeting PI3K/Akt/mTOR (LoPic- homeostasis represents an interesting target for research
colo et  al., 2008; Rodon et  al., 2013) may be an efficient and development of new molecules with antitumor activ-
way to increase the expression of catalase in tumors and ity and numerous drugs are currently being clinically eval-
inhibit tumor cell growth. uated. Therefore, several strategies have been developed
In this last decade, other transcription factors looking for the disruption of tumor cell redox homeostasis
(PPARγ, Oct-1, etc.) as well as genetic, epigenetic and and a subsequent cancer cell death (Demizu et al., 2008;
post-transcriptional processes are emerging as crucial Trachootham et al., 2009; Verrax et al., 2009).

Brought to you by | Bibliothèques de l'UCL


Authenticated
Download Date | 9/8/17 1:29 PM
1102      C. Glorieux and P. B. Calderon: Targeting catalase as new cancer treatment approach

is supposed to be the main mechanism of both chemo-


therapeutic drugs (Thayer, 1977; Pelicano et al., 2003). A
decrease in antioxidant levels has also been proposed.
For instance, glutathione (GSH) levels may be decreased
either by its direct binding with phenylethylisothiocy-
anate (PEITC) and sulphoraphane (Trachootham et  al.,
2009) or by inhibition of γ-glutamylcysteine synthetase
activity (a key enzyme involved in GSH synthesis) as done
by Buthionine sulfoximine (Griffith and Meister, 1979).
The development of specific inhibitors of thioredoxin and
thioredoxin reductases was also carried out. Inhibitors
of thioredoxin, such as PX-12, were shown to have potent
antitumor activities (Welsh et al., 2003; Baker et al., 2006).
Figure 2: Hypothetical mechanism of catalase regulation in breast
cancer cell lines.
Conversely, the overexpression of Trx1 is correlated with
A human breast MCF-7 cancer cell line resistant to oxidative stress resistance to anti-cancer drugs (Baker et al., 2006; Kaimul
was generated (Resox cells) by chronically exposing MCF-7 cells with et al. 2007).
an H2O2-generating system (Dejeans et al., 2012). These cells show As quinones display redox cycling abilities thus gen-
increased expression of catalase. A novel promoter region, respon- erating ROS (Kappus and Sies, 1981), the association of
sible for the regulation of catalase expression, was identified at
menadione (a naphthoquinone derivative) and ascor-
−1518/−1226 locus in Resox cells (Glorieux et al., 2016a). The AP-1
family member JunB and retinoic acid receptor alpha (RARα) mediate bate (Figure 3), was employed to trigger tumor cell death
catalase transcriptional activation and repression, respectively, by (Verrax et al., 2011b). This association (Asc/Men) has been
controlling chromatin remodeling through a histone deacetylases shown to have potent in vitro and in vivo antitumor activi-
(HDAC)-dependent mechanism. (HAT: Histobe acetyltransferase). ties enhancing as well the therapeutic effect of currently
Adapted from Glorieux et al. (2016a).
used anticancer drugs (Taper et  al., 1987; Buc C ­ alderon
et  al., 2002; Verrax et  al., 2003). We hypothesized that
H2O2, issued from the redox cycling, is the oxidant species
One of these strategies is the use of ROS-generating responsible for antitumor effects observed both in vitro
compounds such as arsenic trioxide (ATO), currently and in vivo (Verrax et  al., 2006, 2007; Verrax and Buc
employed against promyelocytic leukemia (Valenzuela Calderon, 2009). The induced oxidative stress provokes
et al., 2014; Lo-Coco et al., 2016), or doxorubicin, an anthra- cell necrosis by a wide variety of processes including
cycline used for the treatment of several types of cancer ATP depletion (Verrax et  al., 2004, 2005, 2011a), disrup-
(i.e. breast cancer). Indeed, the impairment of mitochon- tion of Ca2+ homeostasis (Dejeans et al., 2010) and loss of
drial function due to increased levels of superoxide anion HSP90 chaperone activity (Beck et al., 2009, 2011b, 2012).

Figure 3: The quinone redox cycling hypothesis.


The biological activity showed by menadione mainly relies upon its ability to accept one electron from ascorbate to form a semiquinone
radical. When the semiquinone is reduced back to its quinone form, superoxide anion is produced which by dismutation generates hydro-
gen peroxide.

Brought to you by | Bibliothèques de l'UCL


Authenticated
Download Date | 9/8/17 1:29 PM
C. Glorieux and P. B. Calderon: Targeting catalase as new cancer treatment approach      1103

a change of its activity or expression will lead to patho-


logical processes as Zellweger syndrome, acatalasemia or
WAGR syndrome. The subcellular localization of catalase
is mainly peroxisomal but a shuttle between this organelle
and cytoplasm exists and may be involved in the protec-
tion of key cellular elements (i.e. proteins, chromosomes)
against an oxidative damage.
Our group and others demonstrated that catalase
expression is also altered in cancer cells, most likely to
favor cell proliferation by inducing genetic instability
and activation of oncogenes. The regulation of catalase
expression appears to be mainly controlled at transcrip-
tional levels although other mechanisms may also be
involved. In addition of transcription factors like Sp1 and
NF-Y, JunB and RARα transcription factors are crucial
Figure 4: Arsenic trioxide (ATO) decreases catalase expression in
regulators in breast cancer cells by recruiting proteins
breast cancer cells and sensitizes them to pro-oxidant drugs.
involved in transcriptional complexes and chromatin
remodeling.
Therefore, catalase can be a future therapeutic target
Based on the vulnerability of tumor cells to an oxidative
in the context of cancer by using pro-oxidant approaches.
stress, we have induced the alteration of their intracellu-
lar redox homeostasis as a new strategy in the research
Acknowledgments: The authors thank Professor Helmut
and development of new antitumor drugs (Benites et al.,
Sies for the splendid discussion and his precious input.
2010; Arenas et  al., 2013; Valderrama et  al., 2015). A
This project was funded by FNRS-Televie Grant (grant n°
similar approach has been recently developed by using
7.4575.12F).
the SnFe2O4 nanocrystals, a heterogeneous Fenton cata-
lyst, which once internalized into the cancer cells, convert
H2O2 into hydroxyl radicals inducing apoptotic cell death.
In normal cells, the oxidative injury induced by SnFe2O4 is References
prevented by catalase (Lee et al., 2017).
As ATO increases the levels of ROS (Valenzuela et al., Akca, H., Demiray, A., Aslan, M., Acikbas, I., and Tokgun, O. (2013).
2014) and Asc/Men increases the cytotoxicity of chemo- Tumour suppressor PTEN enhanced enzyme activity of GPx,
therapeutic drugs (Taper et al., 1987), we formulated the SOD and catalase by suppression of PI3K/AKT pathway in non-
following hypothesis: Asc/Men potentiates ATO-mediated small cell lung cancer cell lines. J. Enzyme Inhib. Med. Chem.
28, 539–544.
cytotoxicity (Figure 4). ATO decreases catalase expression
Akman, S.A., Forrest, G., Chu, F.F., Esworthy, R.S., and Doroshow,
most likely by activating Akt pathway and/or inducing J.H. (1990). Antioxidant and xenobiotic-metabolizing enzyme
RARα, meanwhile Asc/Men generates H2O2. Despite that gene expression in doxorubicin-resistant MCF-7 breast cancer
Resox cells display high antioxidant defenses (Dejeans cells. Cancer Res. 50, 1397–1402.
et al., 2012; Glorieux et al., 2015, 2016b), an enhanced cell Arenas, P., Pena, A., Rios, D., Benites, J., Muccioli, G.G., Buc Calde-
ron, P., and Valderrama, J.A. (2013). Echo-friendly synthesis
death was observed when they were exposed to a mixture
and antiproliferative evaluation of oxygen substituted diaryl
containing sublethal doses of Asc/Men and ATO. This is ketones. Molecules 18, 9818–9832.
likely due to a decreased transcriptional activity of the Baker, A.M., Oberley, L.W., and Cohen, M.B. (1997). Expression of
human catalase −1518/+16 promoter caused by ATO result- antioxidant enzymes in human prostatic adenocarcinoma.
ing in less catalase protein (Glorieux et al., 2017, unpub- Prostate 32, 229–233.
lished results). Baker, A.F., Dragovich, T., Tate, W.R., Ramanathan, R.K., Roe, D.,
Hsu, C.H., Kirkpatrick, D.L., and Powis, G. (2006). The antitu-
mor thioredoxin-1 inhibitor PX-12 (1-methylpropyl 2-imidazolyl
disulfide) decreases thioredoxin-1 and VEGF levels in cancer

Conclusions patient plasma. J Lab Clin Med. 147, 83–90.


Barletta, C., Castello, M.A., Ferrante, E., Mavelli, I., Clerico, A.,
­Ciriolo, M.R., and Vignetti, P. (1985). 11p13 deletion and
Under stress conditions, the antioxidant enzyme catalase reduced RBC catalase in a patient with aniridia, glaucoma and
plays a major role by detoxifying H2O2. As consequences, bilateral Wilms’ tumor. Tumori 71, 119–121.

Brought to you by | Bibliothèques de l'UCL


Authenticated
Download Date | 9/8/17 1:29 PM
1104      C. Glorieux and P. B. Calderon: Targeting catalase as new cancer treatment approach

Bauer, G. (2012). Tumor cell-protective catalase as a novel target for dismutase and catalase in lung cancer. Cancer Res. 61,
rational therapeutic approaches based on specific intercellular 8578–8585.
ROS signaling. Anticancer Res. 32, 2599–2624. Cullen, J.J., Mitros, F.A., and Oberley, L.W. (2003). Expression of anti-
Bauer, G. (2015). Increasing the endogenous NO level causes oxidant enzymes in diseases of the human pancreas: another
catalase inactivation and reactivation of intercellular apoptosis link between chronic pancreatitis and pancreatic cancer.
signaling specifically in tumor cells. Redox Biol. 6, 353–371. Pancreas 26, 23–27.
Bauer, G. and Graves, D.B. (2016). Mechanisms of selective Deichman, G.I. (2000). Natural selection and early changes of
antitumor action of cold atmospheric plasma-derived reac- phenotype of tumor cells in vivo: acquisition of new defense
tive oxygen and nitrogen species. Plasma Process Polym. 13, mechanisms. Biochemistry (Mosc). 65, 78–94.
1157–1178. Deichman, G.I. (2002). Early phenotypic changes of in vitro trans-
Bauer, G. and Motz, M. (2016). The antitumor effect of single- formed cells during in vivo progression: possible role of the
domain antibodies directed towards membrane-associated host innate immunity. Semin. Cancer Biol. 12, 317–326.
catalase and superoxide dismutase. Anticancer Res. 36, Dejeans, N., Tajeddine, N., Beck, R., Verrax, J., Taper, H., Gailly, P.,
5945–5956. and Buc Calderon, P. (2010). Endoplasmic reticulum calcium
Beck, R., Verrax, J., Gonze, T., Zappone, M., Curi Pedrosa, R., Taper, release potentiates the ER stress and cell death caused by
H., Feron, O., and Buc Calderon, P. (2009). Hsp90 cleavage by an oxidative stress in MCF-7 cells. Biochem. Pharmacol. 79,
an oxidative stress leads to Bcr-Abl degradation and leukemia 1221–1230.
cell death. Biochem. Pharmacol. 77, 375–383. Dejeans, N., Guenin, S., Glorieux, C., Beck, R., Sid, B., Rousseau, R.,
Beck, R., Pedrosa, R.C., Dejeans, N., Glorieux, C., Leveque, P., Gal- Bisig, B., Delvenne, P., Buc Calderon, P., and Verrax, J. (2012).
lez, B., Taper, H., Eeckhoudt, S., Knoops, L., Buc Calderon, P., Overexpression of GRP94 in breast cancer cells resistant to oxi-
et al. (2011a). Ascorbate/menadione-induced oxidative stress dative stress promotes high levels of cancer cell proliferation
kills cancer cells that express normal or mutated forms of the and migration: implications for tumor recurrence. Free Radic.
oncogenic protein Bcr-Abl. An in vitro and in vivo mechanistic Biol. Med. 52, 993–1002.
study. Invest New Drugs 29, 891–900. Demizu, Y., Sasaki, R., Trachootham, D., Pelicano, H., Colacino, J.A.,
Beck, R., Dejeans, N., Glorieux, C., Curi Pedrosa, R., Vasquez, D., Liu, J., and Huang, P. (2008). Alterations of cellular redox state
Valderrama, J.A., Buc Calderon, P., and Verrax, J. (2011b). Mole- during NNK-induced malignant transformation and resistance
cular chaperone Hsp90 as a target for oxidant-based anticancer to radiation. Antiox. Redox Signal. 10, 951–961.
therapies. Curr. Med. Chem. 18, 2816–2825. Dufier, J.L., Phug, L.H., Schmelck, P., Fekete, C., de Grouchy, G.J.,
Beck, R., Dejeans, N., Glorieux, C., Creton, M., Delaive, E., Dieu, M., Turleau, C., and Haye, C. (1981). [Intercalary deletion of the
Raes, M., Leveque, P., Gallez, B., Depuydt, M., et al. (2012). short arm of chromosome 11: aniridia, glaucoma, staturopon-
Hsp90 is cleaved by reactive oxygen species at a highly con- deral and mental retardation, sexual ambiguity, gonadoblas-
served N-terminal amino acid motif. PLoS One 7, e40795. toma and catalase deficiency]. Bull. Soc. Ophtalmol. Fr. 81,
Benites, J., Valderrama, J.A., Bettega, K., Curi Pedrosa, R., Buc 747–749.
Calderon, P., and Verrax, J. (2010). Biological evaluation of Fenton, H.J.H. (1894). Oxidation of tartaric acid in presence of iron. J.
donor-acceptor aminonaphthoquinones as antitumor agents. Chem. Soc. Trans. 65, 899–911.
Eur. J. Med. Chem. 45, 6052–6057. Fita, I. and Rossmann, M.G. (1985a). The NADPH binding site on beef
Böhm, B., Heinzelmann, S., Motz, M., and Bauer, G. (2015). Extracel- liver catalase. Proc. Natl. Acad. Sci. USA 82, 1604–1608.
lular localization of catalase is associated with the transformed Fita, I. and Rossmann, M.G. (1985b). The active center of catalase. J.
state of malignant cells. Biol Chem. 396, 1339–1356. Mol. Biol. 185, 21–37.
Brown, G.C. (1995). Reversible binding and inhibition of catalase by Fong, K.M., Zimmerman, P.V., and Smith, P.J. (1994). Correlation
nitric oxide. Eur J Biochem. 232, 188–191. of loss of heterozygosity at 11p with tumour progression and
Brunelli, L., Yermilov, V., and Beckman, J.S. (2001). Modulation of survival in non-small cell lung cancer. Genes Chromosomes
catalase peroxidatic and catalatic activity by nitric oxide. Free Cancer 10, 183–189.
Radic. Biol. Med. 30, 709–714. Forsberg, L., Lyrenas, L., de Faire, U., and Morgenstern, R. (2001).
Buc Calderon, P., Cadrobbi, J., Marquez, C., Hong-Ngo, N., Jamison, A common functional C-T substitution polymorphism in the
J.M., Gilloteaux, J., Summers, J.L., and Taper, H.S. (2002). promoter region of the human catalase gene influences
Potential therapeutic application of the association of vitamins transcription factor binding, reporter gene transcription and is
C and K3 in cancer treatment. Curr. Med. Chem. 9, 2271–2285. correlated to blood catalase levels. Free Radic. Biol. Med. 30,
Chance, B. (1948). The enzyme-substrate compounds of catalase 500–505.
and peroxides. Nature 161, 914–917. Gabdoulline, R.R., Kummer, U., Olsen, L.F., and Wade, R.C. (2003).
Chance, B. (1949). The primary and secondary compounds of Concerted simulations reveal how peroxidase compound
catalase and methyl or ethyl hydrogen peroxide; kinetics and III formation results in cellular oscillations. Biophys. J. 85,
activity. J. Biol. Chem. 179, 1341–1369. 1421–1428.
Chance, B., Greenstein, D.S., and Roughton, F.J. (1952). The Gebicka, L. and Didik, J. (2009). Catalytic scavenging of peroxyni-
mechanism of catalase action. I. Steady-state analysis. Arch. trite by catalase. J. Inorg. Biochem. 103, 1375–1379.
Biochem. Biophys. 37, 301–321. Glorieux, C., Dejeans, N., Sid, B., Beck, R., Buc Calderon, P.,
Chance, B., Sies, H., and Boveris, A. (1979). Hydroperoxide meta- and Verrax, J. (2011). Catalase overexpression in mammary
bolism in mammalian organs. Physiol Rev. 59, 527–605. cancer cells leads to a less aggressive phenotype and an
Chung-man, H.J., Zheng, S., Comhair, S.A., Farver, C., and Erzurum, altered response to chemotherapy. Biochem. Pharmacol. 82,
S.C. (2001). Differential expression of manganese ­superoxide 1384–1390.

Brought to you by | Bibliothèques de l'UCL


Authenticated
Download Date | 9/8/17 1:29 PM
C. Glorieux and P. B. Calderon: Targeting catalase as new cancer treatment approach      1105

Glorieux, C., Zamocky, M., Sandoval, C.J.M., Verrax, J., and Buc Cal- Kirkman, H.N. and Gaetani G.F. (1984). Catalase: a tetrameric
deron, P. (2015). Regulation of catalase expression in healthy enzyme with four tightly bound molecules of NADPH. Proc.Natl.
and cancer cells. Free Radic. Biol. Med. 87, 84–97. Acad. Sci. USA 81, 4343–4347.
Glorieux, C., Sandoval, J.M., Fattaccioli, A., Dejeans, N., Garbe, J.C., Kirkman, H.N., Rolfo, M., Ferraris, A.M., and Gaetani, G.F. (1999).
Dieu, M., Verrax, J., Renard, P., Huang, P., and Buc Calderon, P. Mechanisms of protection of catalase by NADPH. Kinetics and
(2016a). Chromatin remodeling regulates catalase expression stoichiometry. J. Biol. Chem. 274, 13908–13914.
during cancer cells adaptation to chronic oxidative stress. Free Koepke, J.I., Nakrieko, K.A., Wood, C.S., Boucher, K.K., Terlecky, L.J.,
Radic. Biol. Med. 99, 436–450. Walton, P.A., and Terlecky, S.R. (2007). Restoration of peroxiso-
Glorieux, C., Sandoval, J.M., Dejeans, N., Ameye, G., Poirel, H.A., mal catalase import in a model of human cellular aging. Traffic
Verrax, J., and Buc Calderon P. (2016b). Overexpression of 8, 1590–1600.
NAD(P)H:quinone oxidoreductase 1 (NQO1) and genomic gain Kwei, K.A., Finch, J.S., Thompson, E.J., and Bowden, G.T. (2004).
of the NQO1 locus modulates breast cancer cell sensitivity to Transcriptional repression of catalase in mouse skin tumor
quinones. Life Sci. 145, 57–65. progression. Neoplasia 6, 440–448.
Goth, L., Rass, P., and Pay, A. (2004). Catalase enzyme mutations Lauer, C., Volkl, A., Riedl, S., Fahimi, H.D., and Beier, K. (1999).
and their association with diseases. Mol. Diagn. 8, 141–149. Impairment of peroxisomal biogenesis in human colon carci-
Gregoire, M.J., Pernot, C., Himont, F., Pierson, M., and Gilgenkrantz, noma. Carcinogenesis 20, 985–989.
S. (1983). [Chromosome 11 and cancer]. J.Genet.Hum. 31, Lazarow, P.B. and De Duve, C. (1973). The synthesis and turnover
31–36. of rat liver peroxisomes. V. Intracellular pathway of catalase
Griffith, O.W. and Meister, A. (1979). Potent and specific inhibition synthesis. J. Cell Biol. 59, 507–524.
of glutathione synthesis by buthionine sulfoximine (S-n-butyl Lee, K.T., Lu, Y.J., Mi, F.L., Burnouf, T., Wei, Y.T., Chiu, S.C., Chuang,
homocysteine sulfoximine). J. Biol. Chem. 254, 7558–7560. E.Y., and Lu, S.Y. (2017). Catalase-modulated heterogeneous
Heinzelmann, S. and Bauer, G. (2010). Multiple protective functions Fenton reaction for selective cancer cell eradication: SnFe2O4
of catalase against intercellular apoptosis-inducing ROS signal- nanocrystals as an effective reagent for treating lung cancer
ling of human tumor cells, Biol. Chem. 391, 675–693. cells. ACS Appl. Mater. Interfaces 9, 1273–1279.
Ho, Y.S., Xiong, Y., Ma, W., Spector, A., and Ho, D.S. (2004). Mice Liu, K., Liu, X., Wang, M., Wang, X., Kang, H., Lin, S., Yang, P., Dai,
lacking catalase develop normally but show differential sensi- C., Xu, P., Li, S., et al. (2016). Two common functional catalase
tivity to oxidant tissue injury. J. Biol. Chem. 279, 32804–32812. gene polymorphisms (rs1001179 and rs794316) and cancer
Hwang, T.S., Choi, H.K., and Han, H.S. (2007). Differential expres- susceptibility: evidence from 14,942 cancer cases and 43,285
sion of manganese superoxide dismutase, copper/zinc super- controls. Oncotarget 7, 62954–62965.
oxide dismutase, and catalase in gastric adenocarcinoma and Lo-Coco, F., Cicconi, L., and Breccia, M. (2016). Current standard
normal gastric mucosa. Eur. J. Surg. Oncol. 33, 474–479. treatment of adult acute promyelocytic leukaemia. Br. J. Hae-
Johansson, L.H. and Borg, L.A. (1988). A spectrophotometric method matol. 172, 841–854.
for determination of catalase activity in small tissue samples, Loew, O. (1900). A new enzyme of general occurrence in organisms.
Anal. Biochem. 174, 331–336. Science 11, 701–702.
Jones, P. and Dunford, H.B. (2005). The mechanism of Compound I LoPiccolo, J., Blumenthal, G.M., Bernstein, W.B., and Dennis, P.A.
formation revisited. J. Inorg. Biochem. 99, 2292–2298. (2008). Targeting the PI3K/Akt/mTOR pathway: effective
Kaimul, A.M., Nakamura, H., Masutani, H., and Yodoi, J. (2007). combinations and clinical considerations. Drug Resist. Updat.
Thioredoxin and thioredoxin-binding protein-2 in cancer and 11, 32–50.
metabolic syndrome. Free Radic. Biol. Med. 43, 861–868. Ludwig, C.U., Raefle, G., Dalquen, P., Stulz, P., Stahel, R., and
Kalinina, E.V., Chernov, N.N., Saprin, A.N., Kotova, Y.N., Andreev, Obrecht, J.P. (1991). Allelic loss on the short arm of chro-
Y.A., Solomka, V.S., and Scherbak, N.P. (2006). Changes in mosome 11 in non-small-cell lung cancer. Int. J. Cancer 49,
expression of genes encoding antioxidant enzymes, heme 661–665.
oxygenase-1, Bcl-2, and Bcl-xl and in level of reactive oxygen Marklund, S.L., Westman, N.G., Lundgren, E., and Roos, G.
species in tumor cells resistant to doxorubicin. Biochemistry (1982). Copper- and zinc-containing superoxide dismutase,
(Mosc.) 71, 1200–1206. manganese-containing superoxide dismutase, catalase, and
Kappus, H. and Sies, H. (1981). Toxic drug effects associated with glutathione peroxidase in normal and neoplastic human cell
oxygen metabolism: redox cycling and lipid peroxidation. Expe- lines and normal human tissues. Cancer Res. 42, 1955–1961.
rientia 37, 1233–1241. Matsui, T., Ozaki, S., Liong, E., Phillips, G.N., and Watanabe, Jr., Y.
Kasugai, I. and Yamada, M. (1992). High production of catalase in (1999). Effects of the location of distal histidine in the reac-
hydrogen peroxide-resistant human leukemia HL-60 cell lines. tion of myoglobin with hydrogen peroxide. J. Biol. Chem. 274,
Leuk. Res. 16, 173–179. 2838–2844.
Kato, S., Ueno, T., Fukuzumi, S., and Watanabe, Y. (2004). Catalase Menegon, S., Columbano, A., and Giordano, S. (2016). The dual
reaction by myoglobin mutants and native catalase: mechanis- roles of Nrf2 in cancer. Trends Mol. Med. 22, 578–593.
tic investigation by kinetic isotope effect. J. Biol. Chem. 279, Michalopoulos, E.E., Bevilacqua, P.J., Stokoe, N., Powers, V.E., Wil-
52376–52381. lard, H.F., and Lewis, W.H. (1985). Molecular analysis of gene
Keilin, D. and Hartree, E.F. (1945). Properties of catalase. Catalysis of deletion in aniridia-Wilms tumor association. Hum. Genet. 70,
coupled oxidation of alcohols. Biochem. J. 39, 293–301. 157–162.
Kim, H.S., Lee, T.B., and Choi, C.H. (2001). Down-regulation of cata- Middelkoop, E., Wiemer, E.A., Schoenmaker, D.E., Strijland, A., and
lase gene expression in the doxorubicin-resistant AML subline Tager, J.M. (1993). Topology of catalase assembly in human
AML-2/DX100. Biochem. Biophys. Res. Commun. 281, 109–114. skin fibroblasts. Biochim. Biophys. Acta 1220, 15–20.

Brought to you by | Bibliothèques de l'UCL


Authenticated
Download Date | 9/8/17 1:29 PM
1106      C. Glorieux and P. B. Calderon: Targeting catalase as new cancer treatment approach

Moran, E.C., Kamiguti, A.S., Cawley, J.C., and Pettitt, A.R. (2002). Reimer, D.L., Bailley, J., and Singh, S.M. (1994). Complete cDNA and
Cytoprotective antioxidant activity of serum albumin and 5′ genomic sequences and multilevel regulation of the mouse
autocrine catalase in chronic lymphocytic leukaemia. Br. J. Hae- catalase gene. Genomics 21, 325–336.
matol. 116, 316–328. Rodon, J., Dienstmann, R., Serra, V., and Tabernero, J. (2013). Devel-
Mueller, S., Riedel, H.D., and Stremmel, W. (1997). Direct evidence opment of PI3K inhibitors: lessons learned from early clinical
for catalase as the predominant H2O2 -removing enzyme in trials. Nat. Rev. Clin. Oncol. 10, 143–153.
human erythrocytes. Blood 90, 4973–4978. Rohrdanz, E. and Kahl, R. (1998). Alterations of antioxidant enzyme
Murakami, K., Ichinohe, Y., Koike, M., Sasaoka, N., Iemura, S., Nat- expression in response to hydrogen peroxide. Free Radic. Biol.
sume, T., and Kakizuka, A. (2013). VCP is an integral component Med. 24, 27–38.
of a novel feedback mechanism that controls intracellular Rohrdanz, E., Schmuck, G., Ohler, S., and Kahl, R. (2001). The influ-
localization of catalase and H2O2 Levels. PLoS One 8, e56012. ence of oxidative stress on catalase and MnSOD gene transcrip-
Nagem, R.A., Martins, E.A., Goncalves, V.M., Aparicio, R., and tion in astrocytes. Brain Res. 900, 128–136.
Polikarpov, I. (1999). Crystallization and preliminary X-ray Rovira, C. (2005). Structure, protonation state and dynamics of cata-
diffraction studies of human catalase. Acta Crystallogr. D Biol. lase compound II. Chemphyschem. 6, 1820–1826.
Crystallogr. 55, 1614–1615. Riethmüller, M., Burger. N., and Bauer, G. (2015). Singlet oxygen
Nakashima, H., Yamamoto, M., Goto, K., Osumi, T., Hashimoto, T., treatment of tumor cells triggers extracellular singlet oxygen
and Endo, H. (1990). Isolation and characterization of the rat generation, catalase inactivation and reactivation of intercellu-
catalase-encoding gene. Gene 89, 295. lar apoptosis-inducing signaling. Redox Biol. 6, 157–168.
Nakatani, M. (1961). Studies on histidine residues in hemeproteins Sander, C.S., Hamm, F., Elsner, P., and Thiele, J.J. (2003). Oxidative
related to their activities. V. Photooxidation of catalase in the stress in malignant melanoma and non-melanoma skin cancer.
presence of methylene blue. J. Biochem. 49, 98–102. Br. J. Dermatol. 148, 913–922.
Nenoi, N., Ichimura, S., Mita, K., Yukawa,O., and Cartwright, I.L. Sandstrom, P.A. and Buttke, P.M. (1993). Autocrine production of
(2001). Regulation of the catalase gene promoter by Sp1, extracellular catalase prevents apoptosis of the human CEM
CCAAT-recognizing factors, and a WT1/Egr-related factor in T-cell line in serum-free medium. Proc. Natl. Acad. Sci. USA 90,
hydrogen peroxide-resistant HP100 cells. Cancer Res. 61, 4708–4712.
5885–5894. Scheit, K. and Bauer, G. (2015). Direct and indirect inactivation of
Oshino, N., Oshino, R., and Chance, B. (1973). The characteristics tumor cell protective catalase by salicylic acid and anthocya-
of the ‘peroxidatic’ reaction of catalase in ethanol oxidation. nidins reactivates intercellular ROS signaling and allows for
Biochem J. 131, 555–563. synergistic effects. Carcinogenesis 36, 400–411.
Otera, H. and Fujiki, Y. (2012). Pex5p imports folded tetrameric cata- Schönbein, C.F. (1863). Chemische Mittheilungen. Med. Chir. Trans.
lase by interaction with Pex13p. Traffic 13, 1364–1377. 89, 1–38.
Park, Y.S., You, S.Y., Cho, S., Jeon, H.J., Lee, S., Cho, D.H., Kim, Schriner, S.E., Linford, N.J., Martin, G.M., Treuting, P., Ogburn, C.E.,
J.S., and Oh, J.S. (2016). Eccentric localization of catalase to Emond, M., Coskun, P.E., Ladiges, W., Wolf, N., Van Remmen,
protect chromosomes from oxidative damages during meiotic H., et al. (2005). Extension of murine life span by overex-
maturation in mouse oocytes. Histochem. Cell Biol. 146, pression of catalase targeted to mitochondria. Science 308,
281–288. 1909–1911.
Pelicano, H., Feng, L., Zhou, Y., Carew, J.S., Hileman, E.O., Sen, P., Mukherjee, S., Bhaumik, G., Das, P., Ganguly, S., Choud-
Plunkett, W., Keating, M.J., and Huang, P. (2003). Inhibition hury, N., and Raha, S. (2003). Enhancement of catalase activity
of mitochondrial respiration: a novel strategy to enhance by repetitive low-grade H2O2 exposures protects fibroblasts
drug-induced apoptosis in human leukemia cells by a reactive from subsequent stress-induced apoptosis. Mutat. Res. 529,
oxygen species-mediated mechanism. J. Biol. Chem. 278, 87–94.
37832–37839. Shimozawa, N., Zhang, Z., Suzuki, Y., Imamura, A., Tsukamoto, T.,
Purdue, P.E. and Lazarow, P.B. (1996). Targeting of human catalase Osumi, T., Fujiki, Y., Orii, T., Barth, P.G., Wanders, R.J., et al.
to peroxisomes is dependent upon a novel COOH-terminal (1999). Functional heterogeneity of C-terminal peroxisome tar-
peroxisomal targeting sequence. J. Cell Biol. 134, 849–862. geting signal 1 in PEX5-defective patients. Biochem. Biophys.
Putnam, C.D., Arvai, A.S., Bourne, Y., and Tainer, J.A. (2000). Active Res. Commun. 262, 504–508.
and inhibited human catalase structures: ligand and NADPH Shipman, R., Schraml, P., Colombi, M., and Ludwig, C.U. (1998).
binding and catalytic mechanism. J. Mol. Biol. 296, 295–309. Allelic deletion at chromosome 11p13 defines a tumour sup-
Quan, F., Korneluk, R.G., Tropak, M.B., and Gravel, R.A. (1986). pressor region between the catalase gene and D11S935 in
Isolation and characterization of the human catalase gene. human non-small cell lung carcinoma. Int. J. Oncol. 12, 107–111.
Nucleic Acids Res. 14, 5321–5335. Sies, H. (1974). Biochemistry of the peroxisome in the liver cell.
Radi, R., Turrens, J.F., Chang, L.Y., Bush, K.M., Crapo, J.D., and Free- Angew. Chem. Int. Ed. Engl. 13, 706–718.
man, B.A. (1991). Detection of catalase in rat heart mitochon- Sies, H. (2017). Hydrogen peroxide as a central redox signaling
dria. J. Biol. Chem. 266, 22028–22034. molecule in physiological oxidative stress: oxidative eustress.
Rainis, T., Maor, I., Lanir, A., Shnizer, S., and Lavy, A. (2007). Redox Biol. 11, 613–619.
Enhanced oxidative stress and leucocyte activation in neoplas- Sies, H. and Chance, B. (1970). The steady state level of catalase
tic tissues of the colon. Dig. Dis. Sci. 52, 526–530. compound I in isolated hemoglobin-free perfused rat liver.
Ramu, A., Cohen, L., and Glaubiger, D. (1984). Oxygen radical FEBS Lett. 11, 172–176.
detoxification enzymes in doxorubicin-sensitive and -resistant Stern, K.G. (1937). Spectroscopy of catalase. J. Gen. Physiol. 20,
P388 murine leukemia cells. Cancer Res. 44, 1976–1980. 631–648.

Brought to you by | Bibliothèques de l'UCL


Authenticated
Download Date | 9/8/17 1:29 PM
C. Glorieux and P. B. Calderon: Targeting catalase as new cancer treatment approach      1107

Sumner, J.B. and Dounce, A.L. (1937). Crystalline catalase. Science Verrax, J., Stockis, J., Tison, A., Taper, H.S., and Buc Calderon, P.
85, 366–367. (2006). Oxidative stress by ascorbate/menadione association
Takahara, S. (1952). Progressive oral gangrene probably due to lack kills K562 human chronic myelogenous leukaemia cells and
of catalase in the blood (acatalasaemia); report of nine cases. inhibits its tumour growth in nude mice. Biochem. Pharmacol.
Lancet 2, 1101–1104. 72, 671–680.
Taper, H.S., de Gerlache, J., Lans, M., and Roberfroid, M. (1987). Verrax, J., Stockis, J., Vanbever, S., Taper, H., and Buc Calderon,
Non-toxic potentiation of cancer chemotherapy by combined C P. (2007). Role of glycolysis inhibition and Poly(ADP-ribose)
and K3 vitamin pre-treatment. Int. J. Cancer 40, 575–579. polymerase activation in necrotic-like cell death caused by
Thayer, W.S. (1977). Adriamycin stimulated superoxide formation in ascorbate/menadione-induced oxidative stress in K562 human
submitochondrial particles. Chem. Biol. Interact. 19, 265–278. chronic myelogenous leukemic cells. Int. J. Cancer 120,
Thénard, L.J. (1811). Observations sur des nouvelles combinaisons 1192–1197.
entre l’oxigène et divers acides. Ann. Chim. Phys. 8, 306–312. Verrax, J., Pedrosa, R.C., Beck, R., Dejeans, N., Taper, H., and Buc
Thurman, R.G., Ley, H.G., and Scholz, R. (1972). Hepatic microsomal Calderon, P. (2009). In situ modulation of oxidative stress:
ethanol oxidation. Hydrogen peroxide formation and the role of a novel and efficient strategy to kill cancer cells. Curr. Med.
catalase. Eur. J. Biochem. 25, 420–430. Chem. 16, 1821–1830.
Trachootham, D., Alexandre, J., and Huang, P. (2009). Targeting can- Verrax, J., Dejeans, N., Sid, B., Glorieux, C., and Buc Calderon, P.
cer cells by ROS-mediated mechanisms: a radical therapeutic (2011a). Intracellular ATP levels determine cell death fate of
approach? Nat. Rev. Drug Discov. 8, 579–591. cancer cells exposed to both standard and redox chemothera-
Turdi, S., Li, Q., Lopez, F.L., and Ren, J. (2007). Catalase alleviates peutic agents. Biochem. Pharmacol. 82, 1540–1548.
cardiomyocyte dysfunction in diabetes: role of Akt, Forkhead Verrax, J., Beck, R., Dejeans, N., Glorieux, C., Sid, B., Pedrosa, R.C.,
transcriptional factor and silent information regulator 2. Life Benites, J., Vasquez, D., Valderrama, J.A., and Buc Calderon, P.
Sci. 81, 895–905. (2011b). Redox-active quinones and ascorbate: an innovative
Valderrama, J.A., Rios, D., Muccioli, G.G., Buc Calderon, P., Brito, cancer therapy that exploits the vulnerability of cancer cells to
I., and Benites, J. (2015). Hetero-annulation reaction between oxidative stress. Anticancer Agents Med. Chem. 11, 213–221.
2-acylnaphthoquinones and 2-aminobenzothiazoles. A new Vetrano, A.M., Heck, D.E., Mariano, T.M., Mishin, V., Laskin, D.L.,
synthetic route to antiproliferative benzo[g]benzothiazolo[2,3- and Laskin, J.D. (2005). Characterization of the oxidase activity
b]quinazoline-7,12-quinones. Tetrahedron Lett. 56, 5103–5105. in mammalian catalase. J. Biol. Chem. 280, 35372–35381.
Valenzuela, M., Glorieux, C., Stokis, J., Sid, B., Sandoval, M., Felipe, Vlasits, J., Jakopitsch, C., Schwanninger, M., Holubar, P., and
K.B., Kviecinski, M., Verrax, J., and Buc Calderon, P. (2014). Obinger, C. (2007). Hydrogen peroxide oxidation by catalase-
Retinoic acid synergizes ATO-mediated cytotoxicity by preclud- peroxidase follows a non-scrambling mechanism. FEBS Lett.
ing Nrf2 activity in AML cells. Br. J. Cancer 111, 874–882. 581, 320–324.
van Heyningen, V., Boyd, P.A., Seawright, A., Fletcher, J.M., Fantes, Wanders, R.J., Kos, M., Roest, B., Meijer, A.J., Schrakamp, G.,
J.A., Buckton, K.E., Spowart, G., Porteous, D.J., Hill, R.E., and Heymans, H.S., Tegelaers, W.H., van den Bosch, H., Schutgens,
Newton. M.S. (1985). Molecular analysis of chromosome 11 R.B., and Tager, J.M. (1984). Activity of peroxisomal enzymes
deletions in aniridia-Wilms tumor syndrome. Proc. Natl. Acad. and intracellular distribution of catalase in Zellweger syn-
Sci. USA 82, 8592–8596. drome. Biochem. Biophys. Res. Commun. 123, 1054–1061.
Venkatesan, B., Mahimainathan, L., Das, F., Ghosh-Choudhury, N., Wang, C.D., Sun, Y., Chen, N., Huang, L., Huang, J.W., Zhu, M., Wang,
and Ghosh, C.G. (2007). Downregulation of catalase by reactive T., and Ji, Y.L. (2016). The role of catalase C262T gene polymor-
oxygen species via PI 3 kinase/Akt signaling in mesangial cells. phism in the susceptibility and survival of cancers. Sci. Rep. 6,
J. Cell Physiol. 211, 457–467. 26973.
Verrax, J. and Buc Calderon, P. (2009). Pharmacologic concentra- Warburg, O. (1923). Versuche an uberlebendem karzinomgewebe.
tions of ascorbate are achieved by parenteral administration Biochem. Z. 142, 317–333.
and exhibit antitumoral effects. Free Radic. Biol. Med. 47, Welsh, S.J., Williams, R.R., Birmingham, A., Newman, D.J., Kirkpat-
32–40. rick, D.L., and Powis, G. (2003). The thioredoxin redox inhibi-
Verrax, J., Cadrobbi, J., Delvaux, M., Jamison, J.M., Gilloteaux, J., tors 1-methylpropyl 2-imidazolyl disulfide and pleurotin inhibit
Summers, J.L., Taper, H.S., and Buc Calderon, P. (2003). The hypoxia-induced factor 1alpha and vascular endothelial growth
association of vitamins C and K3 kills cancer cells mainly by factor formation. Mol. Cancer Ther. 2, 235–243.
autoschizis, a novel form of cell death. Basis for their potential Wieacker, P., Mueller, C.R., Mayerova, A., Grzeschik, K.H., and
use as coadjuvants in anticancer therapy. Eur. J. Med. Chem. Ropers, H.H. (1980). Assignment of the gene coding for human
38, 451–457. catalase to the short arm of chromosome 11. Ann. Genet. 23,
Verrax, J., Cadrobbi, J., Marques, C., Taper, H., Habraken, Y., Piette, 73–77.
J., and Buc Calderon, P. (2004). Ascorbate potentiates the cyto- Wink, D.A. and Mitchell, J.B. (1998). Chemical biology of nitric oxide:
toxicity of menadione leading to an oxidative stress that kills insights into regulatory, cytotoxic, and cytoprotective mecha-
cancer cells by a non-apoptotic caspase-3 independent form of nisms of nitric oxide. Free Radic. Biol. Med. 25, 434–456.
cell death. Apoptosis 9, 223–233. Winternitz, M.C. and Meloy, C.R. (1908). On the occurrence of
Verrax, J., Delvaux, M., Beghein, N., Taper, H., Gallez, B., and Buc catalase in human tissues and its variations in diseases. J. Exp.
Calderon, P. (2005). Enhancement of quinone redox cycling Med. 10, 759–781.
by ascorbate induces a caspase-3 independent cell death in Yano, S., Arroyo, N., and Yano, N. (2004a). Catalase binds Grb2 in
human leukaemia cells. An in vitro comparative study. Free tumor cells when stimulated with serum or ligands for integrin
Radic. Res. 39, 649–657. receptors. Free Radic. Biol. Med. 36, 1542–1554.

Brought to you by | Bibliothèques de l'UCL


Authenticated
Download Date | 9/8/17 1:29 PM
1108      C. Glorieux and P. B. Calderon: Targeting catalase as new cancer treatment approach

Yano, S., Arroyo, N., and Yano, N. (2004b). SHP2 binds catalase and Zamocky, M., Furtmuller, P.G., and Obinger, C. (2008). Evolution of
acquires a hydrogen peroxide-resistant phosphatase activity catalases from bacteria to humans. Antioxid. Redox. Signal. 10,
via integrin-signaling. FEBS Lett. 577, 327–332. 1527–1548.
Zamocky, M. and Koller, F. (1999). Understanding the structure and Zamocky, M., Furtmuller, P.G., and Obinger, C. (2010). Evolution of
function of catalases: clues from molecular evolution and in structure and function of Class I peroxidases. Arch. Biochem.
vitro mutagenesis. Prog. Biophys. Mol. Biol. 72, 19–66. Biophys. 500, 45–57.

Brought to you by | Bibliothèques de l'UCL


Authenticated
Download Date | 9/8/17 1:29 PM

Potrebbero piacerti anche