You are on page 1of 16

Biochimica et Biophysica Acta 1603 (2002) 31 – 46

www.bba-direct.com

Review
The role of iron in cell cycle progression and the proliferation
of neoplastic cells
Nghia T.V. Le, Des R. Richardson *
The Iron Metabolism and Chelation Group, The Heart Research Institute, 145 Missenden Rd, Camperdown, Sydney, New South Wales, 2050 Australia
Received 3 May 2002; accepted 6 June 2002

Abstract

Iron (Fe) is an obligate requirement for life and it is well known that Fe depletion leads to G1/S arrest and apoptosis. These facts, together
with studies showing that Fe chelators can inhibit the growth of aggressive tumours such as neuroblastoma, suggest that Fe-deprivation may be
an important therapeutic strategy. To optimise the anti-proliferative effects of Fe chelators, the role of Fe in cell cycle control requires intense
investigation. For many years, Fe chelators were known to prevent the activity of the R2 subunit of ribonucleotide reductase (RR) that catalyzes
the conversion of ribonucleotides into deoxyribonucleotides (dNTPs) for DNA synthesis. In addition, Fe depletion may also inhibit the newly
identified p53-inducible form of this molecule called p53R2. This protein has the same Fe-binding sites as found in R2, and its activity is
thought to supply dNTPs for the critical process of DNA repair. Iron chelation also causes hypophosphorylation of the retinoblastoma protein
(pRb) and decreases the expression of cyclins A, B and D, which are vital for cell cycle progression. Other regulatory molecules whose
expression is affected by Fe depletion include p53 and hypoxia inducible factor-1a (HIF-1a). The levels of p53 increase following Fe chelation
via the ability of HIF-1a to bind and stabilize p53. The activity of HIF-1a is controlled by an Fe-dependent enzyme known as HIF-a prolyl
hydroxylase (PH). Chelation of Fe from this enzyme inhibits its activity, leading to stabilization of HIF-1a and the subsequent effects on
downstream targets critical for angiogenesis and tumour growth. The levels of p53 may also increase after Fe chelation by phosphorylation of
this protein at serine-15 and -37. This prevents the interaction of p53 with murine double minute-2 (mdm-2) and its degradation. Iron chelation
also markedly increases the mRNA levels of the p53-inducible cyclin-dependent kinase (cdk) inhibitor, p21WAF1/CIP1. Surprisingly, the increase
in p21WAF1/CIP1 mRNA was not reciprocated at the protein level, and this may result in cell cycle dysregulation. This review will focus on the
molecular mechanisms induced following Fe chelation and the role of Fe in cell cycle progression.
D 2002 Elsevier Science B.V. All rights reserved.

Keywords: Iron; Iron chelation; Iron chelator; Iron metabolism; Cancer; p53; p21; Ribonucleotide reductase

1. Introduction: iron as a molecular target for anti- key reactions involved in oxygen sensing [4,5], energy
cancer agents metabolism, respiration, folate metabolism (e.g. serine
hydroxymethyltransferase expression; [6]) and DNA syn-
A major challenge facing researchers is the development thesis (e.g. ribonucleotide reductase (RR) [7]). In fact,
of effective anti-cancer drugs that show high selectivity without Fe, cells are unable to proceed from the G1 to the
against tumour cells compared to normal cells. Compound- S phase of the cell cycle [7].
ing this problem is the emergence of tumours that are Before describing the effects of Fe depletion on the
unresponsive to radiation and chemotherapeutic treatments. molecular control of the cell cycle, we will discuss the
As a consequence, novel strategies for cancer therapy must mechanisms involved in the uptake and intracellular metab-
be sought. olism of Fe.
One such approach involves the targeting of intracellular
iron (Fe) to induce cell cycle arrest and apoptosis (for 1.1. Iron transport, uptake and metabolism by cells
reviews see Refs. [1 –3]). Indeed, Fe is an absolute require-
ment for proliferation, as Fe-containing proteins catalyze Iron is transported in the serum bound to the protein
transferrin (Tf) [8,9]. The uptake of Fe into cells involves
binding of Tf to the transferrin receptor 1 (TfR1). This
*
Corresponding author. Tel.: +61-2-9550-3560; fax: +61-2-9550-3302. complex is then internalised by receptor-mediated endocy-
E-mail address: d.richardson@hri.org.au (D.R. Richardson). tosis and the Fe released from Tf by endosomal acidification

0304-419X/02/$ - see front matter D 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 3 0 4 - 4 1 9 X ( 0 2 ) 0 0 0 6 8 - 9
32 N.T.V. Le, D.R. Richardson / Biochimica et Biophysica Acta 1603 (2002) 31–46

(for reviews see Refs. [8– 11]). Recently, another TfR has subsequent inhibition of energy metabolism and DNA syn-
been identified and is known as TfR2 [12,13]. However, Tf thesis [33,34].
has a lower affinity for TfR2 than TfR1, and the exact
function of TfR2 remains unclear [12,13]. 1.2. The iron-containing enzyme, RR, is an important
Iron released from Tf in the endosome is transported molecular target for iron chelators
through the endosomal membrane by Nramp2 (also known
as the divalent metal ion transporter 1; DMT1; [14,15]; for The fact that Fe is critical for both cellular growth and the
review see Ref. [9]). Once Fe has entered the cell, it is activity of the Fe-containing enzyme RR means that it is an
incorporated into vital Fe-containing molecules (e.g. haem important target for anti-tumour drugs including Fe chelators
and non-haem Fe-containing proteins) and/or stored in the [1]. RR is the rate-limiting enzyme involved in the conver-
Fe-storage protein, ferritin (for review see Ref. [16]). sion of ribonucleotides into deoxyribonucleotides (dNTPs)
Ferritin is a polymeric protein consisting of 24 subunits of for DNA synthesis. This enzyme consists of two protein
two types, a heavy (H) subunit of 21 kDa and a light (L) subunits, designated as R1 and R2 [41,42]. The R1 subunit
subunit of 19 kDa [10,16]. At present, the intermediate(s) contains a catalytic site that is involved in the binding of
involved in intracellular Fe transport from Nramp2 to ribonucleotides for reduction to dNTPs [41,42]. In the
ferritin or to the haem synthesis pathway [17] remains absence of a constant supply of Fe to R2, the R1 subunit is
unknown. Likely molecular candidates for intracellular Fe inactive and thus, RR cannot function [41,42]. The activity
transport include low-Mr Fe complexes [18] or specific Fe- of RR is dependent on Fe since the R2 subunit contains a
binding chaperones [19]. tyrosyl radical that requires Fe for stabilization [42].
The regulation of intracellular Fe homeostasis is largely The potential of RR as a therapeutic target for the
controlled by two mRNA-binding molecules known as iron- treatment of cancer is illustrated by the cytotoxic drug
regulatory protein-1 and -2 (IRP1 and IRP2; [10,20]). The hydroxyurea (HU) that acts to scavenge the tyrosyl radical
mRNAs of certain molecules involved in Fe metabolism of this enzyme [43,44]. However, HU has limited potency
including ferritin, TfR1 and Nramp2 [9,15] contain hairpin- due to its short half-life, low affinity for RR, and the fact
loop structures in their 5V- or 3V-untranslated regions that resistance develops against this drug [45 – 47]. In
(UTRs), called iron-responsive elements (IREs). The IRPs comparison to several key enzymes, RR shows the greatest
can bind to IREs and either stabilize the mRNA against increase in activity in tumour compared to normal cells
degradation or inhibit translation [10,20]. For instance, in [48,49]. This may partly explain the greater sensitivity of
ferritin mRNA, the IRE is located in the 5V UTR, and IRP – neoplastic cells to Fe chelators and RR inhibitors, e.g. HU
RNA-binding inhibits translation, thereby decreasing Fe and TriapineR [50 –52]. Therefore, Fe chelation may pro-
storage. However, in TfR1 mRNA the IREs (five hairpin vide an alternative mechanism to inhibit RR activity in HU-
loops) are located in the 3V UTR, and IRP – RNA-binding resistant tumours. In fact, studies in vitro have shown that
confers stability to the mRNA, enhancing translation and Fe some chelators that are RR inhibitors can overcome HU-
uptake via the TfR1 [10,20]. Therefore, the IRP –RNA- resistance via their ability to bind Fe [53].
binding mechanism is crucial in regulating cellular Fe In contrast to the R2 subunit, R1 is constantly expressed
homeostasis. throughout the cell cycle [54]. Synthesis of R2 is initiated
Neoplastic cells, as compared to normal cells, have during S-phase where the protein is rapidly produced for
significantly higher levels of TfR1 [21] and take up Fe DNA synthesis [54]. Following S-phase, R2 is then quickly
rapidly from Tf [7,22 –24]. This is exploited clinically by degraded to barely detectable levels following M-phase
67
Ga radioimaging, where 67Ga binds to Tf and is taken up [41]. Thus, the regulation of RR involves the control of
by tumour cells that express much higher TfR1 levels than R2 expression and degradation. The effect of Fe chelators
normal cells [21,25,26]. on R2 expression has not been studied in detail, particularly
The marked Fe requirement of tumour cells results in in terms of the differences in expression between normal
their greater sensitivity to Fe chelators relative to normal and neoplastic cells.
cells [7,27,28]. In addition, antibodies that inhibit Fe uptake
by preventing the binding of Tf to the TfR1 can halt tumour 1.3. A novel p53-inducible form of RR (p53R2): another
cell growth in vitro [29] and in vivo [30]. Significantly, molecular target for iron chelators?
these antibodies can stop cellular proliferation in the
absence [29] and presence of Fe chelators [31,32]. While the function of RR is to ensure that the cellular
The importance of Fe as a likely molecular target for levels of dNTPs are adequate for DNA synthesis during S-
anti-cancer agents is exemplified by the fact that nitric oxide phase [41,42], little is known about the supply of dNTPs to
(NO) produced by activated macrophages acts as a cytotoxic cells that are undergoing DNA repair. The recent discovery
effector molecule via its ability to bind Fe [33,34]. More- of a stress-inducible RR subunit called p53-inducible R2
over, NO acts similarly to chelators [35] to induce Fe release (p53R2) has provided insight into this process [55,56]. In
from cells [36 –38] and to inhibit Fe uptake from Tf [39,40]. contrast to R2, expression of p53R2 can be induced via p53
These effects of NO on Fe metabolism may contribute to the following DNA damage by UV-light, gamma-irradiation and
N.T.V. Le, D.R. Richardson / Biochimica et Biophysica Acta 1603 (2002) 31–46 33

adriamycin treatment [55,56]. The RR activity of p53R2 may 2. Effect of iron chelators on tumour growth
not be important for the maintenance of a dNTP pool, but for
the urgent supply of dNTPs for DNA repair [56,57]. It Many studies have shown that the anti-proliferative
remains unknown whether Fe chelation can induce p53R2 effects of Fe deprivation on cancer cells in vitro [60,61],
to compensate for the decreased R2 activity after Fe deple- in vivo [60 – 63] and in a number of clinical trials [64 – 69].
tion. However, intriguingly, it is known that p53R2 can Indeed, the aggressive RR inhibitor, Triapine (3-aminopyr-
substitute for R2 to form an active complex with R1 [57]. idine-carboxaldehyde thiosemicarbazone) [51,52], is a tri-
The mouse R2 subunit has been shown to bind Fe(III) dentate Fe chelator that is currently in Phase I combination
through Glu-170, Glu-233, Glu-267 (bidentate) and His-270 studies and Phase II single agent trials (Vion Pharmaceut-
[58]. These residues are highly conserved in a wide variety icals Ltd—www.vionpharm.com/). The role of Fe chelators
of organisms, indicating their key roles in RR activity [58]. as anti-proliferative agents has previously been reviewed
By sequence comparison, we identified that the same Fe- [1 –3,70] and the reader is referred to these publications for
coordination sites in human R2 are also found in human a more detailed analysis.
p53R2 (Fig. 1). This is an interesting observation, as it may Currently, Fe chelators that show much greater anti-
suggest that the enzymatic activity of this molecule will be proliferative activity than DFO in vitro are in development.
affected by Fe chelators. Hence, because p53R2 activity Some of these include aroylhydrazones such as 2-hydroxy-
could be important for the production of dNTPs for DNA 1-naphthylaldehyde isonicotinoyl hydrazone and its ana-
repair [56,57], Fe chelators may effectively inhibit this logues [28,71 – 74], Triapine [51,52], and tachpyridine
process. Therefore, this could partly explain why DNA- (N,NV,NW-tris(2-pyridylmethyl)-cis,cis-1,3,5-triaminocyclo-
damaging agents such as doxorubicin are more effective at hexane) [75,76]. Comprehensive studies of these chelators
inhibiting cancer cell proliferation in the presence of Fe in vivo are underway, and are essential to judge the efficacy
chelators [59]. and toxicity of these drugs.

Fig. 1. Sequence alignment of the human ribonucleotide reductase subunit, R2, and the p53-inducible R2 subunit. Both human R2 and p53R2 contain the amino
acid residues (denoted by *) critical for binding Fe(III), namely Glu-169, Glu-232, Glu-266 and His-269 [58]. Therefore, p53R2 activity may be susceptible to
Fe chelators and could inhibit the production of deoxyribonucleotides that are essential for DNA repair (see text for details). Note that mouse R2 has one more
amino acid in its sequence compared to human R2. Hence, in the mouse, the R2 Fe-binding sites are at Glu-170, Glu-233, Glu-267 and His-270 [58].
34 N.T.V. Le, D.R. Richardson / Biochimica et Biophysica Acta 1603 (2002) 31–46

Considering the great potential of Fe chelators at inhibit- These are the inhibitors of cdk4 (INK4) or the kinase
ing the proliferation of tumour cells, further studies are inhibitor proteins (CIP/KIP) (for a detailed review see
required to understand the molecular basis of their effects on Ref. [84]). The CIP/KIP family of inhibitors consists of
proliferation. In the present review, we summarize the three proteins: namely p21WAF1/CIP1, p27KIP1 and p57KIP2.
molecular mechanisms of regulating cell cycle progression These molecules prevent cell cycle progression and are
and then discuss the effects of Fe chelators on these commonly regarded as negative regulators of the cycle.
processes. The CIP/KIP inhibitors exert their influence during most
periods of the cell cycle by binding directly to the cdk/cyclin
complex to inhibit their activity [84].
3. The cell cycle There are numerous mechanisms involved in the regu-
lation of CIP/KIP inhibitors. For instance, p21WAF1/CIP1 may
While traditionally it was thought that the anti-prolifer- be maintained at low levels by proteasomal protein degra-
ative effect of Fe chelation was solely related to the dation [85]. However, during periods of stress, the expres-
inhibition of RR, there is growing evidence that this may sion of p21WAF1/CIP1 can be induced by p53 to cause G1/S
not be the only molecular target. Several studies have shown arrest [86]. In contrast to p21WAF1/CIP1, the inhibitory effects
that Fe chelation affects the expression of proteins critical of the KIP proteins involve the modulation of their cdk/
for cell cycle progression [73,74,77,78]. In addition, Fe cyclin-binding affinities via phosphorylation [87,88]. Phos-
chelation can also induce the tumour suppressor protein p53 phorylated KIP inhibitors have lower affinities for their
[79 – 82] that transactivates genes involved in cell cycle respective cdks and readily dissociate for degradation
arrest and apoptosis. Before describing the effects of chela- [87,88].
tors on the cell cycle, we will briefly describe the major The notion that KIP/CIP inhibitors are solely negative
molecules involved in regulating this critical process. regulators of the cell cycle has recently been challenged.
Several studies have shown that the activity of cdk4 or -6
3.1. Cyclins and cyclin-dependent kinases (cdks) during mid-G1 phase not only requires cyclin D, but also
p21WAF1/CIP1 or p27KIP1 [89 – 92]. These studies demon-
The cell cycle consists of five distinct phases, G0, G1, S, strate that KIP/CIP inhibitors are required for the assembly,
G2 and M (Fig. 2A). In the last 15 years, significant advances nuclear accumulation and stabilization of the cdk/cyclin
have been made in understanding the molecular mechanisms complex. In addition, the sequestration of p21WAF1/CIP1 or
involved in cell cycle control (for a detailed review see Ref. p27KIP1 during mid-G1 may allow or enhance the hyper-
[83]). Cell cycle progression depends on the activity of the phosphorylation of pRb by cdk2– cyclin E [89– 91]. At
cdks. The active forms of these kinases occur as hetero- present, the mechanisms that govern whether a KIP/CIP
dimers that are composed of a regulatory subunit called a CKI becomes a positive or negative cell cycle regulator are
cyclin, and its catalytic counterpart, the cdk (Fig. 2A) [83]. unclear.
The cyclin – cdk complexes are activated by phosphor- In contrast to the KIP/CIP family of CKIs, the inhibitory
ylation via cyclin-activating kinases (CAKs) [84]. Cyclins activities of INK4 are restricted to cdk4 and cdk6 [93]. As a
facilitate cell cycle progression and are regarded as positive consequence, the INK4 family is thought to play a major
regulators of the cycle [84]. Indeed, progression through G1 role in G1/S arrest. This family of CKI includes p15INK4B,
is mediated by the expression of cyclin D, E and A [83], p16INK4A, p18INK4C and p19INK4D [93].
while S and M phase progression are characterised by the
appearance of cyclin A and B, respectively [83] (Fig. 2A). 3.3. The p53 tumour suppressor protein
An important regulatory mechanism performed by cdk
molecules involves the phosphorylation of the retinoblas- The p53 tumour suppressor protein plays a pivotal role in
toma protein (pRb) [83]. This molecule mediates progres- preventing cancer development by acting as a critical tran-
sion of cells from G1 to the S phase of the cell cycle. The scription factor to induce cellular arrest and/or apoptosis
hyperphosphorylation of pRb occurs following the forma- [94 – 98]. There are many stress factors that can initiate the
tion of cdk4 – or cdk6 –cyclin D complexes during mid to stabilization, accumulation and activation of p53. These
late G1 phase (Fig. 2A) [83]. This phosphorylation event include DNA damage, decreased dNTP levels, hypoxia,
results in the release of transcription factor, E2F1, from pRb, loss of a cell survival signal, oncogene activation, telomere
which then transcribes genes critical for cell cycle progres- attrition, abnormal cell growth and, more recently, Fe
sion (e.g. dihydrofolate reductase, cyclin A and cyclin E) chelation [79,82,99,100]. Once activated, p53 can initiate
(Fig. 2B) [83]. the transcription and subsequent expression of various
downstream genes that commit the cell to differentiation,
3.2. cdk inhibitors senescence, DNA repair, cellular arrest and/or apoptosis
(Table 1) (for review see Ref. [100]).
The activity of cyclin– cdk complexes are affected by As a consequence of its function, p53 transcription,
cdk inhibitors (CKI) that are classed into two families. translation, protein stabilization, subcellular localization
N.T.V. Le, D.R. Richardson / Biochimica et Biophysica Acta 1603 (2002) 31–46 35

Fig. 2. (A) The cell cycle. (B) Some of the critical molecular events involved in G1/S progression. Cells enter the cycle by expressing cyclin D at G1 which
binds cdk4 or -6 to form a catalytic complex involved in the phosphorylation of the retinoblastoma protein (pRb). This latter molecule acts as a suppressor of
the E2F1 family of transcription factors. Phosphorylation of pRb leads to the dissociation of E2F that promotes the expression of a variety of genes important
for S-phase entry. One of the genes upregulated by E2F1 is cyclin E. This protein when associated with cdk2 further hyperphosphorylates pRb to ‘‘drive’’ cells
into S-phase.

and activation are tightly regulated. A molecule that regu- ity, but also targets the protein [102] and itself for cytoplas-
lates the level of p53 is the murine double minute-2 (mdm- mic export, followed by degradation via the ubiquitin
2) protein that acts as a ubiquitin ligase to mediate p53 pathway [101,103].
degradation [101]. The expression of this protein is con- There are several pathways that can activate and stabilize
trolled by p53 to form an auto-regulatory feedback loop p53 (Fig. 3). A p53 response can be initiated after DNA
(Fig. 3). Due to this loop, any increase in p53 results in damage, for instance when cells are exposed to ionising
increased mdm-2 expression [101]. This control of p53 radiation. As a consequence of DNA damage, the ataxia
activity ensures that correct activation and stabilization of telangiectasia mutated (ATM) protein and check-point kin-
this protein only occurs as a result of cellular stress. In ase 2 (CHK2) phosphokinase are expressed to stabilize p53
addition, mdm-2 not only inhibits p53 transcriptional activ- by phosphorylation [104]. In a similar manner to the ATM
36 N.T.V. Le, D.R. Richardson / Biochimica et Biophysica Acta 1603 (2002) 31–46

Table 1
p53-inducible proteins involved in apoptosis, cell arrest and DNA repair
p53-inducible molecules Comments References
BAX Well-characterised [160]
pro-apoptotic protein.

BAX is inactivated by bcl-2.

Overexpression of BAX
induces mitochondrial
apoptosis.

NOXA Member of the bcl [161]


(for damage) family of pro- and
anti-apoptotic proteins.

Cells exposed to X-ray


irradiation express
NOXA to induce
mitochondrial apoptosis.

PUMA Consists of an alpha [162,163]


(p53 up-regulated and beta form with
modulator of similar apoptotic functions.
apoptosis)
Localizes to mitochondria
to induce apoptosis.

p53AIP1 Pro-apoptotic protein. [164]


(p53 mediated
apoptosis inducing Localizes to mitochondria
protein 1) to induce apoptosis.

Requires p53 to be
phosphorylated at serine 46.

p53 DINP1 Apoptosis induced after [165]


(p53-dependent double-stranded DNA
damage-inducible breaks via p53AIP1
nuclear protein 1) expression.

This protein is associated


with the phosphorylation
of p53 at serine 46 to
initiate apoptosis.

p21WAF1/CIP1 Cell cycle inhibitor of [86]


the CIP/KIP family.

Can arrest cell during all


stages of the cell cycle.

GADD45 Protein causes cellular arrest [166]


and DNA excision repair.

p53R2 Shares 80% homology to R2. [56,57]


(p53-inducible R2)
This protein is probably
required for ribonucleotide
reductase activity during
DNA repair.

MDM-2 Involved in the targeting [101,167]


of p53 via ubiquitination for
proteasomal degradation.
N.T.V. Le, D.R. Richardson / Biochimica et Biophysica Acta 1603 (2002) 31–46 37

pathway, other forms of DNA damage (e.g. chemotherapeu- Normally, during the G1-S phases, cyclin D and cyclin E are
tic drugs, ultraviolet light or protein kinase inhibitors) can bound to cdk4 and cdk2, respectively (Fig. 2A), and act to
stabilize p53 by phosphorylation via the ataxia telangiecta- phosphorylate pRb (for review see Ref. [83]). Reduction in
sia-related (ATR) phosphokinase (Fig. 3) [105]. Alterna- the levels of these cyclins would result in G1/S arrest and
tively, several oncogenes (e.g. c-myc and ras) can increase hypophosphorylation of pRb. Indeed, several investigators
the levels of p53 via the expression of the alternative were able to demonstrate hypophosphorylation of pRb after
reading frame of the INK4A locus (ARF) protein Fe chelation [74,77,114].
([106,107]; for detailed review see Ref. [108]). Apart from the effects on pRb, the marked decrease in
As a consequence of cellular damage, either p53 or cdk2 levels after incubation with Fe chelators [74] would
mdm-2 can be post-translationally modified to stabilize the prevent the formation of the cyclin A – cdk2 complex that is
former for nuclear accumulation. For example, p53 is often essential for G1/S progression (Fig. 2A) [115]. It has been
phosphorylated at serine-15 and/or -37 to prevent its shown that DFO reduces cdk1 expression that is vital for the
association with mdm-2, thus inhibiting its degradation in G2/M transition [116]. Hence, the decrease in the levels of
cells [105]. Another reported mechanism that inhibits cdk1 and cdk2 may explain why a G1/S arrest, and more
mdm-2 activity involves its rapid phosphorylation by rarely, a G2/M arrest can be observed after exposure to
ATM following exposure to radiation [109]. The phosphor- chelators [72,117]. Other cell cycle molecules that showed
ylation of mdm-2 may prevent its association with p53, decreased expression after exposure to Fe chelators include
thus allowing nuclear accumulation of this latter protein cyclin A and B [74], which are required for cdk1 and cdk2
[109]. In contrast, the ARF protein (p14ARF) binds to activity during S, G2 and M progression (Fig. 2A).
mdm-2, resulting in nucleolar sequestration that inhibits The observed decrease in the D-type cyclin family and
p53 degradation [110,111] (Fig. 3). Moreover, the ARF increase in cyclin E after Fe chelation [74] deserve
protein in conjunction with mdm-2 can bind to E2F1 to discussion. The difference in the response of these latter
inhibit its transactivation role and results in cellular arrest cyclins to Fe chelators was not unexpected, as both
[112]. regulate different aspects of G1 progression (Fig. 2A)
[118]. After combination with the relevant cdks, cyclin E
is thought to act on pRb, but only after the D-type cyclins
4. Effect of iron chelation on the expression of cell cycle [115,119,120]. In fact, cdk4– cyclin D complexes phos-
control molecules phorylate pRb, releasing E2F transcription factors that act
to increase cyclin E transcription (Fig. 2B) [115,119,120].
4.1. Effect of iron deprivation on cyclins and cdks Kung et al. [122] have shown that inhibition of DNA
synthesis using the S phase inhibitor aphidicolin [121]
Several investigators examining a variety of tumours, results in cyclin B accumulation. These authors propose
including neuroepithelioma cells, breast cancer cells, leu- that inhibition of cell cycle progression can dissociate
kemia cells and Kaposi’s sarcoma cells, have examined normally coupled cell cycle events, which has implications
the effects of Fe chelation on the expression of cyclin A, for cytotoxicity [121,122]. Similarly, the increase in cyclin
B, E, and D, cdk2 and cdk4 [74,77,78,113]. Iron chela- E protein levels may possibly reflect dysregulation of the
tion caused a marked decrease in the levels of cyclin D1, cell cycle [74]. Alternatively, and more likely, the Fe
D2 and D3, while the expression of cyclin A and B were chelators may inhibit progression through G1 at approx-
also reduced but to a lesser extent [74,77,78]. Importantly, imately the G1/S transition [74], when cyclin E protein
in contrast to the chelators, their Fe complexes had no levels are at their maximum and cyclin D levels have fallen
effect on the expression of these molecules, suggesting [119].
that Fe depletion was critical [74]. Using gene array
technology, Alcantara et al. [113] showed that incubation 4.2. Effect of iron deprivation on cdk inhibitors
of HL-60 leukaemia cells with DFO resulted in a marked
decrease in the mRNA levels of cyclin A, while the The effect of Fe chelators on cell cycle inhibitory
expression of other cyclins such as cyclin D3 and E1 molecules has only been superficially studied. Previous
were only slightly decreased. Hence, there appears to be investigations in this field have focussed on the effects of
some differences in cyclin expression in response to Fe Fe chelators on the more well-characterised cdk inhibitors
chelation, which may reflect variations in cell cycle (e.g. p16INK4A, p21WAF1/CIP1, p27KIP1) [73,74,80,81,123].
control between cell types. Investigations using neuroepithelioma, neuroblastoma, leu-
The expressions of cdk2 mRNA [113] and protein [74] kemia and breast cancer cells demonstrated a large increase
are also decreased after Fe chelation in leukemia and in p21WAF1/CIP1 mRNA after Fe chelation, which was not
neuroepithelioma cells, respectively, while cdk4 protein reciprocated at the protein level [73,74,80,81,123].
levels were not affected in this latter cell type [74]. In Studies with the proteasomal inhibitor, lactacystin, dem-
addition to this, there was a marked increase in cyclin E onstrated restoration of p21WAF1/CIP1 protein expression
expression in neuroepithelioma cells after Fe depletion [74]. after chelation using DFO in human myelocytic leukaemia
38 N.T.V. Le, D.R. Richardson / Biochimica et Biophysica Acta 1603 (2002) 31–46

Fig. 3. Schematic illustration of the stabilization of p53 after DNA damage or oncogene expression. DNA damage stabilizes p53 by the expression of ATM or
ATR phosphokinases. These enzymes phosphorylate p53 at serine-15 and -37 to prevent mdm-2 from sequestering p53 for exportation and degradation.
Alternatively, the detection of abnormal proliferation or expression of various oncogenes leads to expression of the alternative reading frame of the INK4A
locus (ARF) protein. This molecule directly binds to mdm-2 to inhibit the targeted degradation of p53. Once stabilized, p53 is subjected to phosphorylation,
acetylation and sumolyation to enhance or regulate its transactivational abilities. p53 can transactivate a variety of molecules involved in cellular differentiation,
apoptosis, DNA repair and cell cycle arrest (see Table 1).

(ML-1) cells [80]. In addition to this, following Fe chelation 4.3. Effect of iron deprivation on p53 and its downstream
and lactacystin treatment, p21WAF1/CIP1 was ubiquitinated targets
[81]. This suggested that after Fe chelation the levels of
p21WAF1/CIP1 could be controlled via ubiquitination for Several studies have examined the effects of Fe chelation
proteasomal degradation [80,81]. However, the fact that on p53 and its downstream targets p21CIP1/WAF1 and
the p21WAF1/CIP1 mRNA level was markedly increased GADD45 [73,74,76,79,82,123]. Data from these studies
while its protein level was reduced, is totally unexpected. showed increased p53 protein expression after chelation
This observation was paradoxical, as it could be predicted [74,76,79,123], although p53-DNA-binding activity was
that an increase in p21WAF1/CIP1 protein levels may play an not directly assessed.
important role in the G1/S arrest observed after exposure to In a cell-free system, p53 appears to be subject to redox
chelators. regulation [127]. Oxidants or metal chelators disrupt wild-
In contrast to the studies described above [73,74, type p53 conformation and decrease its DNA-binding
80,81,123], incubation of HL-60 leukemia cells with DFO activity in lysates [128 –131]. In marked contrast, using
and the differentiation agent, phorbol myristate acetate whole cell systems, it has been shown that the Fe chelator,
(PMA), results in decreased p21WAF1/CIP1 mRNA levels 1,10-phenanthroline, can induce p53 transactivational activ-
relative to PMA alone [124]. The different effect of DFO ity as well as sequence-specific DNA-binding in a dose-
on p21WAF1/CIP1 expression in this cell line may relate to the dependent manner [131]. In addition, this chelator induced
use of PMA that acts to induce protein kinase C-h activity the expression of several known p53-target genes including
[125]. In fact, it has been shown that the transcription of p21WAF1/CIP1 and mdm-2, but not Bax, GADD45 or the
protein kinase C-h is inhibited by Fe chelators in HL-60 proliferating cell nuclear antigen (PCNA) [131]. Surpris-
cells and can be restored using an Fe source [126]. Hence, ingly, the increase in p53 activity induced by 1,10-phenan-
the results in PMA-treated cells could represent a unique throline was not due to elevated p53 mRNA or protein
effect of Fe-deprivation on p21WAF1/CIP1 expression that levels [131]. Presumably, the increase in p53 activity in
cannot be extrapolated to other cell types not treated with these latter experiments may be due to an increase in the
this agent. conversion of latent p53 to its active DNA-binding form.
N.T.V. Le, D.R. Richardson / Biochimica et Biophysica Acta 1603 (2002) 31–46 39

Recent studies have suggested that several mechanisms may [133,134] and increased ARF expression [108]. The results
be involved in increasing p53-transcriptional activity after of Ashcroft et al. [123] suggest that DFO may stabilize p53
Fe chelation (see Sections 4.3.1 –4.3.3 below). by this former mechanism.
Although treatment of normal and neoplastic cells with
4.3.1. Phosphorylation of p53 at serine-15 may prevent DFO, ActD and Campt caused a robust increase in p53
mdm-2-mediated degradation after iron chelation: the protein levels, only ActD and Campt resulted in an increase
effects of chelators at increasing p53 levels is different in the mRNA and protein levels of p21WAF1/CIP1 [123].
from that found after DNA damage Further studies examining other molecules transactivated by
Ashcroft et al. [123] have demonstrated that p53 was p53 showed that DFO only induced mdm-2 in fibroblasts.
phosphorylated at serine-15 but not serine-20 after incuba- This result was in contrast to ActD, which markedly
tion of DFO with normal MRC-5 fibroblasts and neoplastic induced mdm-2 protein levels in normal and tumour cells
cells (RKO colon cancer cells, MCF7 breast cancer cells and [123]. Since p53 transcriptional activity depends on both
U2OS osteosarcoma cells). These results suggested that Fe- elevation of protein levels and a shift from latent to the
chelator-induced stabilization of p53 may be related to the DNA-binding form of p53 [135], these data suggest that
up-regulation of the ATM and/or ATR activity, both of DFO increased the expression of latent p53 [100].
which phosphorylate p53 at serine-15 (Fig. 4) [104]. Con-
sidering this, later studies showed that DFO activated the 4.3.2. Iron chelation increases the levels of hypoxia-
ATR-signalling pathway that phosphorylates p53 at serine- inducible factor-a (HIF-1a) that stabilizes p53
15 and -37 [132]. It has been shown that hypoxia, cobalt chloride or
Interestingly, it has been shown that the DNA-damaging chelation of Fe from cells by DFO resulted in the rapid
agents, actinomycin D (ActD) and camptothecin (Campt), accumulation of HIF-1a that then stabilized p53 [82] (Figs.
have different effects to DFO on p53 phosphorylation [123]. 4 and 5). The mechanism of stabilization of p53 by HIF-1a
The Campt treatment induced phosphorylation of p53 at involves the direct interaction of these two molecules [82].
serine-15 and -20, while no phosphorylation was detected Increased p53 stability may then initiate G1/S arrest and
with ActD [123]. Since ActD, Campt and DFO all increased apoptosis via downstream effectors [136].
p53 protein levels, these studies suggested the agents act via HIF is a heterodimer composed of a and h subunits that
different mechanisms to stabilize the protein. Indeed, there plays a central role in oxygen homeostasis (for review see
are at least two mechanisms for disrupting mdm-2 function, Ref. [137]). Its transcriptional targets include genes with
namely phosphorylation of p53 to prevent mdm-2 binding critical roles in erythropoiesis, energy metabolism, vaso-

Fig. 4. Schematic illustration demonstrating the effect of Fe chelation on the induction of p53-dependent and -independent pathways and their failure to result
in an increase in p21WAF1/CIP1 protein levels. Iron chelation can result in elevated p53 protein levels via the hypoxia inducible factor-1a (HIF-1a) that binds to
p53 to stabilize it. In addition to this mechanism, Fe chelation may induce the phosphorylation of p53 at serine-15 and -37. This prevents association with
mdm-2 that sequesters p53 for proteasomal degradation. Furthermore, Fe depletion can inhibit ribonucleotide reductase (RR) activity that results in reduced
deoxyribonucleotide (dNTP) levels that may act to increase p53 transcriptional activity. The increase in p53 results in transactivation of its downstream target,
p21WAF1/CIP1. The increase in p21WAF1/CIP1 mRNA does not result in an increase in its protein levels, perhaps due to increased proteasomal degradation. The
nature of the p53-independent pathway involved in increasing p21WAF1/CIP1 expression remains unknown. Possible transactivating candidate molecules that can
bind to the p21WAF1/CIP1 promoter include AP-2, Sp1, Sp3, BRACA1 and E2F1 (see text for details).
40 N.T.V. Le, D.R. Richardson / Biochimica et Biophysica Acta 1603 (2002) 31–46

motor function and angiogenesis [137]. Under conditions of indicates that Fe is critical for both the hypoxic response
hypoxia, HIF-1a binds to the aryl hydrocarbon receptor [4,5] and cell cycle control.
nuclear translocator (ARNT; also known as HIF-1h) to The stabilization of p53 following HIF-1a accumulation
activate expression of genes important in cell survival. may also be a response to limit the transactivational activity
Alternatively, HIF-1a can bind to p53 and promote p53- of HIF-1a [139] (Fig. 5). However, it is unclear whether this
dependent activities, e.g. apoptosis [138]. Whether HIF-1a repression of HIF-1a transactivational activity is relevant
could be involved in either function may be controlled by its after Fe deprivation. This is because the downstream targets
phosphorylation state [138]. Phosphorylation of HIF-1a of HIF-1a such as vascular endothelial growth factor-1
results in the binding to ARNT, while the dephosphorylated (VEGF1) and erythropoietin (EPO) increase following Fe
form of HIF-1a becomes bound to p53 and mediates chelation [140,141] (Fig. 5).
apoptosis [138].
Recently, the mechanism of how HIF-1a becomes 4.3.3. Can iron chelation increase p53 transcriptional
stabilized under hypoxia or Fe-depletion has become clear. activity by depletion of deoxyribonucleotide pools?
Both Ivan et al. [4] and Jaakkola et al. [5] have charac- Apart from the mechanisms described above, decreased
terized the activity of an Fe-dependent HIF-a prolyl dNTP levels are known to increase the transcriptional
hydroxylase (PH) that hydroxylates HIF-1a at proline- activity of p53 (Fig. 4). In fact, previous studies have
564 (Fig. 5). This then mediates the ubiquitination of suggested that p53 can act as a metabolic sensor [99]. When
HIF-1a for proteasomal degradation. When cells are Fe- nucleotide pools are disturbed by a variety of inhibitors, this
replete and under normal oxygen tension, the hydroxyla- results in a reversible Go/G1 cell cycle arrest associated with
tion of proline-564 of HIF-1a alters its conformation to induction of p53 and p21WAF1/CIP1 [99]. Since Fe chelators
allow binding to the von Hippel –Lindau protein (pVHL) decrease RR activity and subsequently dNTP concentrations
[4,5]. Subsequently, the pVHL molecule organizes the [53,142,143], this may be important in terms of the signal-
assembly of a complex that activates the ubiquitin E3 ling involved in increasing p53 levels (Fig. 4). It is interest-
ligase. This enzyme then ubiquitinates the bound HIF-1a ing to note that the phosphorylation state of p53 is the same
to target it for proteasomal degradation [4,5]. Thus, Fe after Fe chelation as found in cells treated with the potent
depletion or hypoxia prevents hydroxylation of proline-564 RR inhibitor, HU, and hypoxia [132]. This could indicate
to stabilize the HIF-1a protein that subsequently leads to that the p53-signalling pathways induced due dNTP deple-
increased p53 levels (Fig. 5). This important discovery tion, hypoxia and Fe deprivation may be similar.

Fig. 5. Schematic illustration of the downstream effects of the transcription factors, HIF-1a and PLAGL2, following Fe chelation. Cellular Fe depletion
prevents hydroxylation of HIF-1a by an Fe-dependent enzyme called HIF-a prolyl-hydroxylase. Hydroxylation of HIF-1a by this protein occurs in Fe-replete
and oxygenated cells. This results in binding to the pVHL for ubiquitination and proteasomal degradation. The inhibition of HIF-a prolyl hydroxylase by Fe-
chelation allows the stabilization and accumulation of HIF-1a. Once expressed, HIF-1a can initiate the transcription of genes involved in neovascularization,
oxygen transport and Fe uptake, namely, VEGF1, EPO and TfR1, respectively. Iron depletion can also increase TfR1 expression via the IRP – IRE mechanism.
Up-regulation of HIF-1a expression following Fe chelation stabilizes p53 (see Section 1.1). The expression of p53 may then allow transactivation of various
genes involved in cellular arrest, DNA repair or apoptosis (see Table 1). The p53 protein may inhibit the transcriptional role of HIF-1a. Iron chelation can also
induce the expression of the transcription factor, PLAGL2, which then transactivates the pro-apoptotic protein Nip3 (see text for details).
N.T.V. Le, D.R. Richardson / Biochimica et Biophysica Acta 1603 (2002) 31–46 41

4.4. Iron deprivation induces a p53-independent increase in Recent studies by Fan et al. [154] have added further
p21WAF1/CIP1 and GADD45 mRNA levels evidence to support a link between N-myc and Fe metab-
olism. These authors showed that Fe chelation using DFO
Recently, growing evidence has indicated that Fe chela- resulted in down-regulation of N-myc protein levels [154].
tion can mediate a p53-independent response to cause cell In contrast, the expression of other oncogenes such as c-fos
cycle arrest and apoptosis in cells lacking functional p53 was increased, while housekeeping genes remained
[53,73,74,76]. The p53-independent responses induced by unchanged [154]. Nuclear run-on experiments and transient
Fe-depletion resulted in the up-regulation of GADD45 and reporter gene expression studies suggested that the decrease
p21WAF1/CIP1 mRNA levels in SK-N-MC neuroepithelioma in N-myc expression occurred at the level of transcription
and K562 erythroleukemia cells that have mutant p53 initiation and by inhibiting N-myc promoter activity [154].
[73,74] (Fig. 4). Surprisingly, despite a relatively large
increase in mRNA levels there was no change in both 5.2. c-myc
GADD45 or p21WAF1/CIP1 protein expression [74] (Fig. 4).
The transcription factors responsible for activating the Studies using peripheral blood mononuclear cells have
p53-independent increase in p21WAF1/CIP1 expression fol- shown that chelators decrease c-myc expression and this can
lowing Fe chelation [73,74] are unknown at present. How- be prevented by the addition of Fe [155]. At the mRNA
ever, numerous transcription factors such as specificity level, c-myc expression in HL-60 cells has been shown to
protein (Sp) 1 and 3, activating protein 2 (AP2), breast increase and then subsequently decrease after Fe chelation
and ovarian cancer susceptibility gene product (BRACA1) using DFO [156]. More recently, studies by Wu et al. [157]
and E2F1, have DNA-binding sites within the p21WAF1/CIP1 have demonstrated that c-myc coordinately regulates genes
promoter (for a detailed review, see Ref. [144]). It is already controlling intracellular Fe levels, namely the ferritin H-
known that the Sp1 transcription factor can induce p21WAF1/ subunit and IRP2. In fact, c-myc was shown to repress
CIP1
mRNA expression following exposure to okadaic acid ferritin-H chain expression and stimulate the transcription of
and PMA [145]. In contrast to Sp1 that has its own trans- IRP2. Interestingly, the down-regulation of ferritin-H chain
activation sequence within the p21WAF1/CIP1 promoter, p53 expression appeared to be required for cell transformation
homologues like p73 can transactivate the p21WAF1/CIP1 [157]. These changes in ferritin-H and IRP2 expression may
gene using the p53 protein consensus sequence [146,147]. be necessary for maintaining intracellular Fe levels that are
Contrary to the p21WAF1/CIP1 gene, GADD45 has an AP1 required for critical molecules such as RR. However, it is of
site, and AP1 is the only transcription factor-binding site interest that transfection of c-myc into other cell types
other than that of p53 which initiates its transcription [148]. resulted in increased ferritin-H chain expression [153].
At present, the mechanism(s) by which Fe chelation induces Collectively, these results suggest that c-myc plays a mod-
cell cycle arrest in cells without wild-type p53 remains ulatory role in controlling the expression of ferritin-H.
unclear and warrants further investigation. In summary, these studies examining the roles of N-myc
and c-myc suggest some linkage between the metabolism of
Fe and the role of proto-oncogenes in controlling cellular
5. Iron-regulated expression of N-myc and c-myc proto- proliferation. However, these results are preliminary, and
oncogenes require further investigation.

5.1. N-myc
6. PLAGL2: an iron-regulated zinc finger protein is
Apart from the effect of Fe on p53, the influence of involved in apoptosis
intracellular Fe levels on N-myc and c-myc expression have
also been investigated. The N-myc gene has some sequence Furukawa et al. [158] have recently identified the pleo-
homology to c-myc, and is known to play an important role morphic adenoma gene like 2 (PLAGL2) after screening Fe-
in the pathogenesis of the aggressive childhood cancer, deficient inducible cDNA from the mouse macrophage cell
neuroblastoma [149 – 151]. The amplification of N-myc line RAW264.7. This molecule contained seven C2H2 zinc
correlates with rapid NB progression [149 – 151]. Further- finger motifs and was induced when cells were incubated
more, it is worthy to note that secretion of high levels of under hypoxic conditions or with DFO [158]. Moreover,
serum ferritin exist in NB patients with advanced disease, PLAGL2 appeared to act as a transcription factor through
and this is actually used as a prognostic indicator [151,152]. the HIF-1 responsive element [158,159]. Treatment of cells
These facts are of interest, since transfection of cells with with DFO induced apoptosis, nuclear accumulation of
copies of the related c-myc gene resulted in overexpression PLAGL2 and an increase in the expression of the pro-
of ferritin-H chain in the SW-613-S human carcinoma cell apoptotic factor Nip3 [159] (Fig. 5). While the Nip3
type [153]. Hence, these observations could indicate a promoter contained an hypoxia-responsive element, the
possible role of N-myc in neuroblastoma cell Fe metabo- transcription of this gene was activated independent of
lism. HIF-1a. In addition, it was suggested that PLAGL2 acted
42 N.T.V. Le, D.R. Richardson / Biochimica et Biophysica Acta 1603 (2002) 31–46

as a tumour suppressor in association with HIF-1a [159]. [7] H.M. Lederman, A. Cohen, J.W. Lee, M.H. Freedman, E.W. Gelfand,
Deferoxamine: a reversible S-phase inhibitor of human lymphocyte
Hence, in combination with the induction of p53 and its
proliferation, Blood 64 (1984) 748 – 753.
possible role in cell cycle arrest and apoptosis, PLAGL2 [8] E.H. Morgan, Transferrin: biochemistry, physiology and clinical sig-
represents a novel apoptotic pathway that is activated after nificance, Mol. Aspects Med. 4 (1981) 1 – 123.
Fe chelation (Fig. 5). [9] N.C. Andrews, Disorders of iron metabolism, N. Engl. J. Med. 341
(1999) 1986 – 1995.
[10] D.R. Richardson, P. Ponka, The molecular mechanisms of the me-
tabolism and transport of iron in normal and neoplastic cells, Bio-
7. Summary chim. Biophys. Acta 1331 (1997) 1 – 40.
[11] P. Ponka, C. Beaumont, D.R. Richardson, Function and regulation of
The emergence of tumours that are resistant to conven- transferrin and ferritin, Semin. Hematol. 35 (1998) 35 – 54.
tional therapies is of great concern. Since cancer cells have a [12] H. Kawabata, R. Yang, T. Hirama, P.T. Vuong, S. Kawano, A.F.
higher Fe requirement than their normal counterparts, they Gombart, H.P. Koeffler, Molecular cloning of transferrin receptor
2: a new member of the transferrin receptor-like family, J. Biol.
are susceptible to the effects of Fe chelation. At present, the Chem. 274 (1999) 20826 – 20832.
exact molecular mechanisms that result in cell cycle arrest [13] H. Kawabata, R.S. Germain, P.T. Vuong, T. Nakamaki, J.W. Said,
after Fe chelation have only been superficially assessed. H.P. Koeffler, Transferrin receptor 2-a supports cell growth both in
However, the studies performed have revealed a glimpse of iron-chelated cultured cells and in vivo, J. Biol. Chem. 275 (2000)
the role of Fe in complex biological processes such as 16618 – 16625.
[14] M.D. Fleming, C.C. Trenor, M.A. Su, D. Foernzler, D.R. Beier, W.F.
oxygen sensing and DNA synthesis. Indeed, we now know Dietrich, N.C. Andrews, Microcytic anaemia mice have a mutation in
that the expression of critical molecules involved in cell Nramp2, a candidate iron transporter gene, Nat. Genet. 16 (1997)
cycle progression and proliferation such as p53, HIF-1a, 383 – 386.
p21WAF1/CIP1, N-myc and c-myc are regulated by intra- [15] H. Gunshin, B. MacKenzie, U.V. Berger, Y. Gunshin, M.F. Romero,
cellular Fe levels. Further examination of the role that Fe W.F. Boron, S. Nussberger, J.L. Gollan, M.A. Hediger, Cloning and
characterization of a mammalian proton-coupled metal-ion transport-
plays in cell cycle progression and proliferation may be vital er, Nature 388 (1997) 482 – 488.
in terms of designing and developing Fe chelators as [16] P.M. Harrison, P. Arosio, The ferritins: molecular properties, iron
effective anti-tumour agents. storage function and cellular regulation, Biochim. Biophys. Acta
1275 (1996) 161 – 203.
[17] P. Ponka, Tissue specific regulation of iron metabolism and heme
synthesis: distinct control mechanisms in erythoid cells, Blood 89
Acknowledgements (1997) 1 – 25.
[18] A. Jacobs, Low molecular weight intracellular iron transport com-
The authors are grateful to Mr. David Lovejoy and Ms. pounds, Blood 50 (1977) 433 – 439.
[19] M.D. Harrison, C.E. Jones, C.T. Dameron, Copper chaperones: func-
Juliana Kwok for helpful comments on the manuscript prior
tion, structure and copper-binding properties, J. Biol. Inorg. Chem. 4
to submission. D.R.R. thanks the National Health and (1999) 145 – 153.
Medical Research Council and Australian Research Council [20] M.W. Hentze, C. Kuhn, Molecular control of vertebrate iron metab-
for grant and fellowship funding. The Heart Research olism: mRNA-based regulatory circuits operated by iron, nitric oxide
Institute is also thanked for financial support. and oxidative stress, Proc. Natl. Acad. Sci. U. S. A. 93 (1996) 8175 –
8182.
[21] J.W. Larrick, P. Cresswell, Modulation of cell surface iron transferrin
receptors by cellular density and state of activation, J. Supramol.
References Struct. 11 (1979) 579 – 586.
[22] D.R. Richardson, E. Baker, The uptake of iron and transferrin by the
[1] C. Hershko, Control of disease by selective iron depletion: a novel human malignant melanoma cell, Biochim. Biophys. Acta 1053
therapeutic strategy utilizing iron chelators, Bailliere’s Clin. Haema- (1990) 1 – 12.
tol. 7 (1994) 965 – 1000. [23] D.R. Richardson, E. Baker, The effect of desferrioxamine and ferric
[2] D.R. Richardson, Potential of iron chelators as effective anti-prolif- ammonium citrate on the uptake of iron by the membrane iron-bind-
erative agents, Can. J. Physiol. Pharm. 75 (1997) 1164 – 1180. ing component of human melanoma cells, Biochim. Biophys. Acta
[3] D.R. Richardson, Analogues of pyridoxal isonicotinoyl hydrazone 1103 (1992) 275 – 280.
(PIH) as potential Fe chelators for the treatment of neoplasia, Leuk. [24] D. Trinder, O. Zak, P. Aisen, Transferrin receptor-independent uptake
Lymphoma 31 (1998) 47 – 60. of differic transferrin by human hepatoma cells with antisense inhib-
[4] M. Ivan, K. Kondo, H. Yang, W. Kim, J. Valiando, M. Ohh, A. Salic, ition of receptor expression, Hepatology 23 (1996) 1512 – 1520.
J.M. Asara, W.S. Lane , W.G. Kaelin Jr., HIFalpha targeted for VHL- [25] C.R. Chitamber, P.A. Seligman, Effects of different transferrin forms
mediated destruction by proline hydroxylation: implications for O2 on transferrin receptor expression, iron uptake, and cellular prolifer-
sensing, Science 292 (2001) 464 – 468. ation of human leukemic HL60 cells: mechanisms responsible for the
[5] P. Jaakkola, D.R. Mole, Y.M. Tian, M.I. Wilson, J. Gielbert, S.J. specific cytotoxicity of transferrin-gallium, J. Clin. Invest. 78 (1986)
Gaskell, A. Kriegsheim, H.F. Hebestreit, M. Mukherji, C.J. Scho- 1538 – 1546.
field, P.H. Maxwell, C.W. Pugh, P.J. Ratcliffe, Targeting of HIF- [26] S.M. Chan, P.B. Hoffer, N. Maric, P. Duray, Inhibition of gallium-67
alpha to the von Hippel – Lindau ubiquitylation complex by O2-regu- uptake in melanoma by an anti-human transferrin receptor monoclo-
lated prolyl hydroxylation, Science 292 (2001) 468 – 472. nal antibody, J. Nucl. Med. 28 (1987) 1303 – 1307.
[6] E.W. Oppenheim, T.M. Nasrallah, M.G. Mastri, P.J. Stover, Mimo- [27] D. Hileti, P. Panayiotidis, A.V. Hoffbrand, Iron chelators induce
sine is a cell-specific antagonist of folate metabolism, J. Biol. Chem. apoptosis in proliferating cells, Br. J. Haematol. 89 (1995) 181 – 187.
275 (2000) 19268 – 19274. [28] D. Lovejoy, D.R. Richardson, Novel ‘‘hybrid’’ iron chelators derived
N.T.V. Le, D.R. Richardson / Biochimica et Biophysica Acta 1603 (2002) 31–46 43

from aroylhydrazones and thiosemicarbazones demonstrate high anti- side Diphosphate Reductase Activity, Pergamon, Oxford, 1989, pp.
proliferative activity that is selective for tumor cells, Blood 100 165 – 201.
(2002) 666 – 676. [47] P.R. Gwilt, W.G. Tracewell, Pharmacokinetics and pharmacodynam-
[29] I.S. Trowbridge, F. Lopez, Monoclonal antibody to transferrin recep- ics of hydroxyurea, Clin. Pharmacokinet. 34 (1998) 347 – 358.
tor blocks transferrin binding and inhibits tumor cell growth in vitro, [48] E. Takeda, G. Weber, Role of ribonucleotide reductase in expression
Proc. Natl. Acad. Sci. U. S. A. 79 (1982) 1175 – 1179. in the neoplastic program, Life Sci. 28 (1981) 1007 – 1014.
[30] J.D. Kemp, J.A. Thorson, B.C. Stewart, P.W. Nauman, Inhibition of [49] L. Witt, T. Yap, R.L. Blakley, Regulation of ribonucleotide reductase
hematopoietic tumour growth by combined treatment with deferox- activity and its possible exploitation in chemotherapy, Adv. Enzyme
amine and an IgG monoclonal antibody against the transferrin recep- Regul. 17 (1978) 157 – 171.
tor: evidence for a theshold model of iron deprivation toxicity, [50] S. Salmon, Chemotherapeutic agents—cancer chemotherapy, in: F.
Cancer Res. 52 (1992) 4144 – 4148. Meyers, E. Jawetz, E. Goldfien, A. Goldfien (Eds.), Review of
[31] J.D. Kemp, T. Cardillo, B.C. Stewart, E. Kehrberg, G. Weiner, B. Medical Pharmacology, Lange Medical Publications, Los Altos,
Hedlund, P.W. Naumann, Inhibition of lymphoma growth in vivo by 1980, pp. 77 – 513.
combined treatment with hydroxyethyl starch deferoxamine conju- [51] R.A. Finch, M.C. Liu, A.H. Cory, J.G. Cory, A.C. Sartorelli, Triapine
gate and IgG monoclonal antibodies against the transferrin receptor, (3-aminopyridine-2-carboxaldehyde thiosemicarbazone; 3-AP): an
Cancer Res. 55 (1995) 3817 – 3824. inhibitor of ribonucleotide reductase with antineoplastic activity,
[32] J.D. Kemp, Iron deprivation and cancer: a view beginning with stud- Adv. Enzyme Regul. 39 (1999) 3 – 12.
ies of monoclonal antibodies against the transferrin receptor, Histol. [52] R.A. Finch, M. Liu, S.P. Grill, W.C. Rose, R. Loomis, K.M. Vasquez,
Histopathol. 12 (1997) 291 – 296. Y. Cheng, A.C. Sartorelli, Triapine (3-aminopyridine-2-carboxalde-
[33] J.B. Hibbs Jr., R.R. Taintor, Z. Vavrin, Iron deprivation: possible hyde-thiosemicarbazone): a potent inhibitor of ribonucleotide reduc-
cause of tumour cell cytoxicity induced by activated macrophages, tase activity with broad spectrum antitumor activity, Biochem.
Biochem. Biophys. Res. Commun. 123 (1984) 716 – 723. Pharmacol. 59 (2000) 983 – 991.
[34] J.B. Hibbs Jr., R.R. Taintor, Z. Vavrin, E.M. Rachlin, Nitric oxide: a [53] D.A. Green, W.E. Antholine, S.J. Wong, D.R. Richardson, C.R. Chi-
cytotoxic activated macrophage effector molecule, Biochem. Bio- tambar, Ribonucleotide reductase as a preferential target for the in-
phys. Res. Commun. 157 (1988) 87 – 94. hibition of leukemic cell growth by 311, a novel iron chelator of the
[35] R.N. Watts, D.R. Richardson, Nitrogen monoxide (NO) and glucose: pyridoxal isonicotinoyl hydrazone class, Clin. Cancer Res. 7 (2001)
unexpected links that affect intracellular iron metabolism and iron 3574 – 3579.
mobilisation, J. Biol. Chem. 276 (2001) 4724 – 4732. [54] A. Chabes, L. Thelander, Controlled protein degradation regulates
[36] S.L. Wardrop, R. Watts, D.R. Richardson, Nitrogen monoxide (NO) ribonucleotide reductase activity in proliferating mammalian cells
activates IRP-RNA binding by two possible mechanisms—an effect during the normal cell cycle and in response to DNA damage and
on the Fe – S cluster and iron release from cells, Biochemistry 39 replication blocks, J. Biol. Chem. 275 (2000) 17747 – 17753.
(2000) 2748 – 2758. [55] K. Nakano, E. Balint, M. Ashcroft, K.H. Vousden, A ribonucleotide
[37] F.P. Nestle, R.N. Greene, K. Kichian, P. Ponka, W.S. Lapp, Activa- reductase gene is a transcriptional target of p53 and p73, Oncogene
tion of macrophage cytostatic effector mechanisms during acute 19 (2000) 4283 – 4289.
graft-versus-host disease: release of intracellular iron and nitric ox- [56] H. Tanaka, H. Arakawa, T. Yamaguchi, K. Shiraishi, S. Fukuda, K.
ide-mediated cytostasis, Blood 96 (2000) 1836 – 1843. Matsui, Y. Takei, Y. Nakamura, A ribonucleotide reductase gene
[38] F. Feger, H. Ferry-Dumazet, M. Mamani Matsuda, J. Bordenave, M. involved in a p53-dependent cell-cycle checkpoint for DNA damage,
Dupouy, A.K. Nussler, M. Arock, L. Devevey, J. Nafziger, J.J. Nature 404 (2000) 42 – 49.
Guillosson, J. Reiffers, M.D. Mossalayi, Role of iron in tumor cell [57] O. Guittet, P. Hakansson, N. Voevodskaya, A. Graslund, H. Arakawa,
protection from the pro-apoptotic effect of nitric oxide, Cancer Res. Y. Nakamura, L. Thelander, Mammalian p53R2 protein forms an
61 (2001) 5289 – 5294. active ribonucleotide reductase in vitro with the R1 protein, which
[39] D.R. Richardson, V. Neumannova, P. Ponka, Nitrogen monoxide is expressed both in resting cells in response to DNA damage and in
decreases iron uptake from transferrin but does not mobilise iron proliferating cells, J. Biol. Chem. 276 (2001) 40647 – 40651.
from prelabelled neoplastic cells, Biochim. Biophys. Acta 1266 [58] B. Kauppi, B.B. Nielsen, S. Ramaswamy, I.K. Larson, M. Thelander,
(1995) 250 – 260. L. Thelander, H. Eklund, The three-dimensional structure of mam-
[40] R.N. Watts, D.R. Richardson, Examination of the mechanism of malian ribonucleotide reductase protein R2 reveals a more-accessible
action of nitrogen monoxide on iron uptake from transferrin, J. iron-radical site than Escherichia coli R2, J. Mol. Biol. 262 (1996)
Lab. Clin. Med. 136 (2000) 149 – 156. 706 – 720.
[41] L. Thelander, P. Reichard, Reduction of ribonucleotides, Annu. Rev. [59] J. Blatt, D. Huntley, Enhancement of in vitro activity against neuro-
Biochem. 48 (1979) 133 – 158. blastoma by doxorubicin and deferoxamine, J. Natl. Cancer Inst. 81
[42] L. Thelander, A. Gräslund, M. Thelander, Continual presence of (1989) 866 – 870.
oxygen and iron are required for mammalian ribonucleotide reduc- [60] H.J. Thompson, K. Kennedy, M. Witt, J. Juzefyk, Effect of dietary
tion: possible regulation mechanism, Biochem. Biophys. Res. Com- iron deficiency or excess on the induction of mammary carcinogen-
mun. 110 (1983) 859 – 865. esis by 1-methyl-1-nitrosourea, Carcinogenesis 12 (1991) 111 – 114.
[43] S.E. Salmon, Part VII. Chemotherapeutic agents: cancer chemother- [61] F. Wang, R.L. Elliott, J.F. Head, Inhibitory effect of deferoxamine
apy, in: F.H. Meyers, E. Jawetz, E. Goldfien (Eds.), Review of Med- mesylate and low iron diet on the 13762NF rat mammary adenocar-
ical Pharmacology, 7th edn., Lange Medical Publications, California, cinoma, Anticancer Res. 19 (1999) 445 – 450.
1980, Chapter 45. [62] L. Pickart, W.H. Goodwin, W. Burgua, T.B. Murphy, D.K. Johnson,
[44] S. Nyholm, L. Thelander, A. Gräslund, Reduction and loss of the iron Inhibition of the growth of cultured cells and an implanted fibrosar-
centre in the reaction of the small subunit of mouse ribonucleotide coma by aroylhydrazone analogs of the Gly – His – Lys – Cu(II) com-
reductase with hydroxyurea, Biochemistry 32 (1993) 11569 – 11574. plex, Biochem. Pharmacol. 32 (1983) 3868 – 3871.
[45] G.L. Beckloff, H.J. Lerner, D. Frost, F.M. Russo-Alesi, S. Gitomer, [63] H.W. Hann, M.W. Stahlhut, R. Rubin, W.C. Maddrey, Antitumor
Hydroxyurea (NSC-32065) in biological fluids: dose – concentration effect of deferoxamine on human hepatocellular carcinoma growing
relationship, Cancer Chemother. Rep. 48 (1965) 57 – 58. in athymic nude mice, Cancer 70 (1992) 2051 – 2056.
[46] E.C. Moore, R.B. Hurlbert, The inhibition of ribonucleoside diphos- [64] Z. Estrov, A. Tawa, X.H. Wang, I.D. Dube, H. Sulh, A. Cohen, E.W.
phate reductase by hydroxyurea, guanazole and pyrazoloimidazole Gelfand, M.H. Freedman, In vivo and in vitro effects of desferriox-
(IMPY), in: J.G. Cory, A.H. Cory (Eds.), Inhibitors of Ribonucleo- amine in neonatal acute leukaemia, Blood 69 (1987) 757 – 761.
44 N.T.V. Le, D.R. Richardson / Biochimica et Biophysica Acta 1603 (2002) 31–46

[65] L. Dezza, M. Cazzola, M. Danova, C. Carlo-Stella, G. Bergamaschi, [83] C.J. Sherr, The Pezcoller lecture: cancer cell cycles revisited, Cancer
S. Brugnatelli, R. Invernizzi, G. Mazzini, A. Riccardi, E. Ascari, Res. 60 (2000) 3689 – 3695.
Effects of desferrioxamine on normal and leukemic human hemato- [84] A. Vidal, A. Koff, Cell-cycle inhibitors: three families united by a
poietic cell growth: in vitro and in vivo studies, Leukemia 3 (1989) common cause, Gene 247 (2000) 1 – 15.
104 – 107. [85] C.G. Maki, P.M. Howley, Ubiquitination of p53 and p21 is differ-
[66] A. Donfrancesco, G. Deb, C. Dominici, D. Pileggi, M.A. Castello, L. entially affected by ionizing and UV radiation, Mol. Cell. Biol. 17
Helson, Effects of a single course of deferoxamine in neuroblastoma (1997) 355 – 363.
patients, Cancer Res. 50 (1990) 4929 – 4930. [86] W.S. el-Deiry, T. Tokino, V.E. Velculescu, D.B. Levy, R. Parsons,
[67] A. Donfrancesco, G. Deb, C. Dominici, A. Angioni, M. Caniglia, L. J.M. Trent, D. Lin, W.E. Mercer, K.W. Kinzler, B. Vogelstein,
De Sio, P. Fidani, A. Amici, L. Helson, Deferoxamine, cyclophos- WAF1, a potential mediator of p53 tumor suppression, Cell 75
phamide, etoposide, carboplatin, and thiotepa (D-CECat): a new cy- (1993) 817 – 825.
toreductive chelation-chemotherapy regimen in patients with [87] M. Pagano, S.W. Tam, A.M. Theodoras, P. Beer-Romero, G. Del Sal,
advanced neuroblastoma, Am. J. Clin. Oncol. 15 (1992) 319 – 322. V. Chau, P.R. Yew, G.F. Draetta, M. Rolfe, Role of the ubiquitin –
[68] A. Donfrancesco, B. De Bernardi, M. Carli, A. Mancini, M. Nigro, L. proteasome pathway in regulating abundance of the cyclin-dependent
De Sio, F. Casale, S. Bagnulo, L. Helson, G. Deb, Deferoxamine (D) kinase inhibitor p27, Science 269 (1995) 682 – 685.
followed by cytoxan (C), etoposide (E), carboplatin (Ca), thio-TEPA [88] L.M. Tsvetkov, K.H. Yeh, S.J. Lee, H. Sun, H. Zhang, p27(Kip1)
(T), induction regimen in advanced neuroblastoma, Eur. J. Cancer ubiquitination and degradation is regulated by the SCF(Skp2) com-
31A (1995) 612 – 615. plex through phosphorylated Thr187 in p27, Curr. Biol. 9 (1999)
[69] A. Donfrancesco, G. Deb, L. De Sio, R. Cozza, A. Castellano, Role of 661 – 664.
desferrioxamine in tumor therapy, Acta Haematol. 95 (1996) 66 – 69. [89] S.W. Blain, E. Montalvo, J. Massague, Differential interaction of the
[70] D.R. Richardson, Iron chelators as therapeutic agents for the treat- cyclin-dependent kinase (Cdk) inhibitor p27Kip1 with cyclin A –
ment of cancer, Crit. Rev. Oncol./Hematol. 42 (2002) 267 – 281. Cdk2 and cyclin D2 – Cdk4, J. Biol. Chem. 272 (1997) 25863 –
[71] D.R. Richardson, E.H. Tran, P. Ponka, The potential of iron chelators 25872.
of the pyridoxal isonicotinoyl hydrazone class as antiproliferative [90] M. Cheng, P. Olivier, J.A. Diehl, M. Fero, M.F. Roussel, J.M. Rob-
agents, Blood 86 (1995) 4295 – 4306. erts, C.J. Sherr, The p21(Cip1) and p27(Kip1) CDK ‘inhibitors’ are
[72] D.R. Richardson, K. Milnes, The potential of iron chelators of the essential activators of cyclin D-dependent kinases in murine fibro-
pyridoxal isonicotinyl hydrazone class as effective antiproliferative blasts, EMBO J. 18 (1999) 1571 – 1583.
agents II: the mechanism of action of ligands derived from salicy- [91] J. LaBaer, M.D. Garrett, L.F. Stevenson, J.M. Slingerland, C. Sand-
laldehyde benzoyl hydrazone and 2-hydroxy-1-naphthylaldehyde hu, H.S. Chou, A. Fattaey, E. Harlow, New functional activities for
benzoyl hydrazone, Blood 89 (1997) 3025 – 3038. the p21 family of CDK inhibitors, Genes Dev. 11 (1997) 847 – 862.
[73] G. Darnell, D.R. Richardson, The potential of iron chelators of the [92] J.R. Alt, A.B. Gladde, J.A. Diehl, p21Cip1 promotes cyclin D1
pyridoxal isonicotinoyl hydrazone class as effective antiproliferative nuclear accumulation via direct inhibition of nuclear export, J. Biol.
agents: III. The effect of the ligands on molecular targets involved in Chem. 277 (2001) 8517 – 8523.
proliferation, Blood 94 (1999) 781 – 792. [93] M. Ruas, G. Peters, The p16INK4a/CDKN2A tumor suppressor and
[74] J. Gao, D.R. Richardson, The potential of iron chelators of the pyr- its relatives, Biochim. Biophys. Acta 1378 (1998) F115 – F177.
idoxal isonicotinoyl hydrazone class as effective antiproliferative [94] M. Hollstein, D. Sidransky, B. Vogelstein, C.C. Harris, p53 mutations
agents: IV. The mechanisms involved in inhibiting cell-cycle progres- in human cancers, Science 253 (1991) 49 – 53.
sion, Blood 98 (2001) 842 – 850. [95] A.J. Levine, J. Momand, C.A. Finlay, The p53 tumour suppressor
[75] S.V. Torti, F.M. Torti, S.P. Whitman, M.W. Brechbiel, G. Park, R.P. gene, Nature 351 (1991) 453 – 456.
Planalp, Tumor cell cytotoxicity of a novel metal chelator, Blood 92 [96] P. Hainaut, M. Hollstein, p53 and human cancer: the first ten thou-
(1998) 1384 – 1389. sand mutations, Adv. Cancer Res. 77 (2000) 81 – 137.
[76] R.D. Abeysinghe, B.T. Greene, R. Haynes, M.C. Willingham, J. [97] G.S. Jimenez, M. Nister, J.M. Stommel, M. Beeche, E.A. Barcarse,
Turner, R.P. Planalp, M.W. Brechbiel, F.M. Torti, S.V. Torti, p53- X.Q. Zhang, S. O’Gorman, G.M. Wahl, A transactivation-deficient
independent apoptosis mediated by tachpyridine, an anti-cancer iron mouse model provides insights into Trp53 regulation and function,
chelator, Carcinogenesis 22 (2001) 1607 – 1614. Nat. Genet. 26 (2000) 37 – 43.
[77] K.S. Kulp, S.L. Green, P.R. Vulliet, Iron deprivation inhibits cyclin- [98] B. Vogelstein, D. Lane, A.J. Levine, Surfing the p53 network, Nature
dependent kinase activity and decreases cyclin D/CDK4 protein lev- 408 (2000) 307 – 310.
els in asynchronous MDA-MB-453 human breast cancer cells, Exp. [99] S.P. Linke, K.C. Clarkin, A. Di Leonardo, A. Tsou, G.M. Wahl, A
Cell Res. 229 (1996) 60 – 68. reversible, p53-dependent G0/G1 cell cycle arrest induced by ribo-
[78] T. Simonart, C. Degraef, G. Andrei, R. Mosselmans, P. Hermans, J.P. nucleotide depletion in the absence of detectable DNA damage,
Van Vooren, J.C. Noel, J.R. Boelaert, R. Snoeck, M. Heenen, Iron Genes Dev. 10 (1996) 934 – 947.
chelators inhibit the growth and induce the apoptosis of Kaposi’s [100] K.H. Vousden, G.F. Woude, The ins and outs of p53, Nat. Cell Biol. 2
sarcoma cells and of their putative endothelial precursors, J. Invest. (2000) E178 – E180.
Dermatol. 115 (2000) 893 – 900. [101] R. Honda, H. Tanaka, H. Yasuda, Oncoprotein MDM2 is a ubiquitin
[79] K. Fukuchi, S. Tomoyasu, H. Watanabe, S. Kaetsu, N. Tsuruoka, K. ligase E3 for tumor suppressor p53, FEBS Lett. 420 (1997) 25 – 27.
Gomi, Iron deprivation results in an increase in p53 expression, Biol. [102] J. Momand, G.P. Zambetti, D.C. Olson, D. George, A.J. Levine, The
Chem. Hoppe-Seyler 376 (1995) 627 – 630. mdm-2 oncogene product forms a complex with the p53 protein and
[80] K. Fukuchi, S. Tomoyasu, H. Watanabe, N. Tsuruoka, K. Gomi, G1 inhibits p53-mediated transactivation, Cell 69 (1992) 1237 – 1245.
accumulation caused by iron deprivation with deferoxamine does not [103] Y. Haupt, R. Maya, A. Kazaz, M. Oren, Mdm2 promotes the rapid
accompany change of pRB status in ML-1 cells, Biochim. Biophys. degradation of p53, Nature 387 (1997) 296 – 299.
Acta 1357 (1997) 297 – 305. [104] A.M. Carr, Cell cycle. Piecing together the p53 puzzle, Science 287
[81] K. Fukuchi, S. Tomoyasu, T. Nakamaki, N. Tsuruoka, K. Gomi, (2000) 1765 – 1766.
DNA damage induces p21 protein expression by inhibiting ubiquiti- [105] R.S. Tibbetts, K.M. Brumbaugh, J.M. Williams, J.N. Sarkaria, W.A.
nation in ML-1 cells, Biochim. Biophys. Acta 1404 (1998) 405 – 411. Cliby, S.Y. Shieh, Y. Taya, C. Prives, R.T. Abraham, A role for ATR
[82] W.G. An, M. Kanekal, M.C. Simon, E. Maltepe, M.V. Blagosklonny, in the DNA damage-induced phosphorylation of p53, Genes Dev. 13
L.M. Neckers, Stabilization of wild-type p53 by hypoxia-inducible (1999) 152 – 157.
factor 1alpha, Nature 392 (1998) 405 – 408. [106] M.J. Elliott, Y.B. Dong, H. Yang, K.M. McMasters, E2F-1 up-regu-
N.T.V. Le, D.R. Richardson / Biochimica et Biophysica Acta 1603 (2002) 31–46 45

lates c-Myc and p14(ARF) and induces apoptosis in colon cancer [127] L.A. Donehower, A. Bradley, The tumor suppressor p53, Biochim.
cells, Clin. Cancer Res. 7 (2001) 3590 – 3597. Biophys. Acta 1155 (1993) 181 – 205.
[107] A.W. Lin, S.W. Lowe, Oncogenic ras activates the ARF-p53 path- [128] T. Russo, N. Zambrano, F. Esposito, R. Ammendola, F. Cimino, M.
way to suppress epithelial cell transformation, Proc. Natl. Acad. Sci. Fiscella, J. Jackman, P.M. O’Connor, C.W. Anderson, E. Appella, A
U. S. A. 98 (2001) 5025 – 5030. p53-independent pathway for activation of WAF1/CIP1 expression
[108] C.J. Sherr, J.D. Weber, The ARF/p53 pathway, Curr. Opin. Genet. following oxidative stress, J. Biol. Chem. 270 (1995) 29386 – 29391.
Dev. 10 (2000) 94 – 99. [129] P. Hainaut, J. Milner, A structural role for metal ions in the ‘‘wild-
[109] R. Khosravi, R. Maya, T. Gottlieb, M. Oren, Y. Shiloh, D. Shkedy, type’’ conformation of the tumor suppressor protein p53, Cancer Res.
Rapid ATM-dependent phosphorylation of MDM2 precedes p53 53 (1993) 1739 – 1742.
accumulation in response to DNA damage, Proc. Natl. Acad. Sci. [130] P. Hainaut, S. Butcher, J. Milner, Temperature sensitivity for confor-
U. S. A. 96 (1999) 14973 – 14977. mation is an intrinsic property of wild-type p53, Br. J. Cancer 71
[110] L. Chin, J. Pomerantz, R.A. DePinho, The INK4a/ARF tumour sup- (1995) 227 – 231.
pressor: one gene-two products-two pathways, Trends Biochem. Sci. [131] Y. Sun, J. Bian, Y. Wang, C. Jacobs, Activation of p53 transcriptional
23 (1998) 291 – 296. activity by 1,10-phenanthroline, a metal chelator and redox sensitive
[111] D.E. Quelle, F. Zindy, R.A. Ashmun, C.J. Sherr, Alternative reading compound, Oncogene 14 (1997) 385 – 393.
frames of the INK4a tumor suppressor gene encode two unrelated [132] E.M. Hammond, N.C. Denko, M.J. Dorie, R.T. Abraham, A.J. Giac-
proteins capable of inducing cell cycle arrest, Cell 83 (1995) 993 – cia, Hypoxia links ATR and p53 through replication arrest, Mol. Cell.
1000. Biol. 22 (2002) 1834 – 1843.
[112] B. Eymin, L. Karayan, P. Seite, C. Brambilla, E. Brambilla, C.J. [133] S.Y. Shieh, M. Ikeda, Y. Taya, C. Prives, DNA damage-induced
Larsen, S. Gazzeri, Human ARF binds E2F1 and inhibits its tran- phosphorylation of p53 alleviates inhibition by MDM2, Cell 91
scriptional activity, Oncogene 10 (2001) 1033 – 1041. (1997) 325 – 334.
[113] O. Alcantara, M. Kalidas, I. Baltathakis, D.H. Boldt, Expression of [134] T. Unger, T. Juven-Gershon, E. Moallem, M. Berger, R. Vogt Sionov,
multiple genes regulating cell cycle and apoptosis in differentiating G. Lozano, M. Oren, Y. Haupt, Critical role for Ser20 of human p53
hematopoietic cells is dependent on iron, Exp. Hematol. 29 (2001) in the negative regulation of p53 by Mdm2, EMBO J. 18 (1999)
1060 – 1069. 1805 – 1814.
[114] N. Terada, J.J. Lucas, E.W. Gelfand, Differential regulation of the [135] T.R. Hupp, D.P. Lane, Allosteric activation of latent p53 tetramers,
tumor suppressor molecules, retinoblastoma susceptibility gene prod- Curr. Biol. 4 (1994) 865 – 875.
uct (Rb) and p53, during cell cycle progression of normal human T [136] P. Carmeliet, Y. Dor, J.M. Herbert, D. Fukumura, K. Brusselmans, M.
cells, J. Immunol. 147 (1991) 698 – 704. Dewerchin, M. Neeman, F. Bono, R. Abramovitch, P. Maxwell, C.J.
[115] C.J. Sherr, G1 phase progression: cycling on cue, Cell 79 (1994) Koch, P. Ratcliffe, L. Moons, R.K. Jain, D. Collen, E. Keshert, Role
551 – 555. of HIF-1alpha in hypoxia-mediated apoptosis, cell proliferation and
[116] C. Brodie, G. Siriwardana, J. Lucas, R. Schleicher, N. Terada, A. tumour angiogenesis, Nature 394 (1998) 485 – 490.
Szepesi, E. Gelfand, P. Seligman, Neuroblastoma sensitivity to [137] G.L. Semenza, HIF-1 and human disease: one highly involved factor,
growth inhibition by deferoxamine: evidence for a block in the G1 Genes Dev. 14 (2000) 1983 – 1991.
phase of the cell cycle, Cancer Res. 53 (1993) 3968 – 3975. [138] H. Suzuki, A. Tomida, T. Tsuruo, Dephosphorylated hypoxia-indu-
[117] F.J. Renton, T.M. Jeitner, Cell cycle-dependent inhibition of the pro- cible factor-1alpha as a mediator of p53-dependent apoptosis during
liferation of human neural tumor cell lines by iron chelators, Bio- hypoxia, Oncogene 20 (2001) 5779 – 5788.
chem. Pharmacol. 51 (1996) 1553 – 1561. [139] M.V. Blagosklonny, W.G. An, L.Y. Romanova, J. Trepel, T. Fojo, L.
[118] D. Resnitzky, S.I. Reed, Different roles for cyclins D1 and E in Neckers, p53 inhibits hypoxia-inducible factor-stimulated transcrip-
regulation of the G1-to-S transition, Mol. Cell. Biol. 15 (1995) tion, J. Biol. Chem. 273 (1998) 11995 – 11998.
3463 – 3469. [140] L.V. Beerepoot, D.T. Shima, M. Kuroki, K.T. Yeo, E.E. Voest, Up-
[119] A. Zetterberg, O. Larsson, K.G. Wiman, What is the restriction point? regulation of vascular endothelial growth factor production by iron
Curr. Opin. Cell Biol. 7 (1995) 835 – 842. chelators, Cancer Res. 56 (1996) 3747 – 3751.
[120] J. DeGregori, T. Kowalik, J.R. Nevins, Cellular targets for activation [141] N.A. Daghman, G.E. Elder, G.A. Savage, P.C. Winter, A.P. Maxwell,
by the E2F1 transcription factor include DNA synthesis- and G1/S- T.R. Lappin, Erythropoietin production: evidence for multiple oxy-
regulatory genes, Mol. Cell. Biol. 15 (1995) 4215 – 4224. gen sensing pathways, Ann. Hematol. 78 (1999) 275 – 278.
[121] R.T. Schimke, A.L. Kung, D.F. Rush, S.W. Sherwood, Differences in [142] S. Nyholm, G.J. Mann, A.G. Johansson, R.J. Bergeron, A. Gräslund,
mitotic control among mammalian cells, Cold Spring Harbor Symp. L. Thelander, Role of ribonucleotide reductase in inhibition of mam-
Quant. Biol. 56 (1991) 417 – 423. malian cell growth by potent iron chelators, J. Biol. Chem. 268
[122] A.L. Kung, S.L. Sherwood, R.T. Schimke, Cell line-specific differ- (1993) 26200 – 26205.
ences in the control of cell cycle progression in the absence of mi- [143] C.E. Cooper, G.R. Lynagh, K.P. Hoyes, R.C. Hider, R. Cammack,
tosis, Proc. Natl. Acad. Sci. U. S. A. 87 (1990) 9553 – 9557. J.B. Porter, The relationship of intracellular iron chelation to the
[123] M. Ashcroft, Y. Taya, K.H. Vousden, Stress signals utilize multiple inhibition and regeneration of human ribonucleotide reductase, J.
pathways to stabilize p53, Mol. Cell. Biol. 20 (2000) 3224 – 3233. Biol. Chem. 271 (1996) 20291 – 20299.
[124] Y. Gazitt, S.V. Reddy, O. Alcantara, J. Yang, D.H. Boldt, A new [144] A.L. Gartel, A.L. Tyner, Transcriptional regulation of the p21(WAF1/
molecular role for iron in regulation of cell cycling and differentiation CIP1) gene, Exp. Cell Res. 246 (1999) 280 – 289.
of HL-60 human leukemia cells: iron is required for transcription of [145] J.R. Biggs, J.E. Kudlow, A.S. Kraft, The role of the transcription
p21(WAF1/CIP1) in cells induced by phorbol myrisate acetate, J. factor Sp1 in regulating the expression of the WAF1/CIP1 gene in
Cell. Physiol. 187 (2001) 124 – 135. U937 leukemic cells, J. Biol. Chem. 271 (1996) 901 – 906.
[125] D.A. Tonetti, C. Henning-Chubb, D.T. Yamanishi, E. Huberman, [146] L. Fang, S.W. Lee, S.A. Aaronson, Comparative analysis of p73 and
Protein kinase C-h is required for macrophage differentiation of hu- p53 regulation and effector functions, J. Cell Biol. 147 (1999) 823 –
man HL-60 leukemia cells, J. Biol. Chem. 269 (1994) 23230 – 23235. 830.
[126] O. Alcantara, L. Obeid, Y. Hannun, P. Ponka, D.H. Boldt, Regulation [147] C.A. Jost, M.C. Marin, W.G. Kaelin Jr., p73 is a simian [correction of
of protein kinase C (PKC) expression by iron: effect of different iron human] p53-related protein that can induce apoptosis, Nature 389
compounds on PKC-h and PKC-a gene expression and role of the (1997) 191 – 194.
5V-flanking region of the PKC-h gene in the response to ferric trans- [148] D.M. Graunke, A.J. Fornace Jr., R.O. Pieper, Presetting of chromatin
ferrin, Blood 84 (1994) 3510 – 3517. structure and transcription factor binding poise the human GADD45
46 N.T.V. Le, D.R. Richardson / Biochimica et Biophysica Acta 1603 (2002) 31–46

gene for rapid transcriptional up-regulation, Nucleic Acids Res. 27 [158] T. Furukawa, Y. Adachi, J. Fujisawa, T. Kambe, Y. Yamaguchi-Iwai,
(1999) 3881 – 3890. R. Sasaki, J. Kuwahara, S. Ikehara, R. Tokunaga, S. Taketani, In-
[149] M. Schwab, K. Alitalo, K.H. Klempnauer, H.E. Varmus, J.M. volvement of PLAGL2 in activation of iron deficient- and hypoxia-
Bishop, F. Gilbert, G. Brodeur, M. Goldstein, J. Trent, Amplified induced gene expression in mouse cell lines, Oncogene 20 (2001)
DNA with limited homology to myc cellular oncogene is shared by 4718 – 4727.
human neuroblastoma cell lines and a neuroblastoma tumor, Nature [159] A. Mizutani, T. Furukawa, Y. Adachi, S. Ikehara, S. Taketani, A zinc-
305 (1983) 245 – 248. finger protein, PLAGL2, induces the expression of a proapoptotic
[150] M. Schwab, H.E. Varmus, J.M. Bishop, K.H. Grzeschik, S.L. Naylor, protein Nip3, leading to cellular apoptosis, J. Biol. Chem. 277 (2002)
A.Y. Sakaguchi, G. Brodeur, J. Trent, Chromosomal localization in 15851 – 15858.
normal human cells and neuroblastomas of a gene related to c-myc, [160] Z.N. Oltvai, C.L. Milliman, S.J. Korsmeyer, Bcl-2 heterodimerizes in
Nature 308 (1984) 288 – 291. vivo with a conserved homolog, Bax, that accelerates programmed
[151] G.M. Brodeur, R.C. Seeger, M. Schwab, H.E. Varmus, J.M. Bishop, cell death, Cell 74 (1993) 609 – 619.
Amplification of N-myc in untreated human neuroblastomas corre- [161] E. Oda, R. Ohki, H. Murasawa, J. Nemoto, T. Shibue, T. Yamashita,
lates with advanced disease stage, Science 224 (1984) 1121 – 1124. T. Tokino, T. Taniguchi, N. Tanaka, Noxa, a BH3-only member of
[152] H.W. Hann, M.W. Stahlhut, A.E. Evans, Serum ferritin as a prognos- the Bcl-2 family and candidate mediator of p53-induced apoptosis,
tic indicator in neuroblastoma: biological effects of isoferritins, Prog. Science 288 (2000) 1053 – 1058.
Clin. Biol. Res. 175 (1985) 331 – 345. [162] K. Nakano, K.H. Vousden, PUMA, a novel proapoptotic gene, is
[153] N. Modjtahedi, T. Frebourg, N. Fossar, C. Lavialle, C. Cremisi, O. induced by p53, Mol. Cell 7 (2001) 683 – 694.
Brison, Increased expression of cytokeratin and ferritin-H genes in [163] J. Yu, L. Zhang, P.M. Hwang, K.W. Kinzler, B. Vogelstein, PUMA
tumorigenic clones of the SW 613-S human colon carcinoma cell induces the rapid apoptosis of colorectal cancer cells, Mol. Cell 7
line, Exp. Cell Res. 201 (1992) 74 – 82. (2001) 673 – 682.
[154] L. Fan, J. Iyer, S. Zhu, K.K. Frick, R.K. Wada, A.E. Eskenazi, P.E. [164] K. Oda, H. Arakawa, T. Tanaka, K. Matsuda, C. Tanikawa, T. Mori,
Berg, N. Ikegani, R.H. Kennett, C.N. Frantz, Inhibition of N-myc H. Nishimori, K. Tamai, T. Tokino, Y. Nakamura, Y. Taya, p53AIP1,
expression and induction of apoptosis by iron chelation in human a potential mediator of p53-dependent apoptosis, and its regulation
neuroblastoma cells, Cancer Res. 61 (2001) 1073 – 1079. by Ser-46-phosphorylated p53, Cell 102 (2000) 849 – 862.
[155] D. Kyriakou, A.G. Eliopoulos, A. Papadakis, M. Alexandrakis, G.D. [165] S. Okamura, H. Arakawa, T. Tanaka, H. Nakanishi, C.C. Ng, Y. Taya,
Eliopoulos, Decreased expression of c-myc oncoprotein by periph- M. Monden, Y. Nakamura, p53DINP1, a p53-inducible gene, regu-
eral blood mononuclear cells in thalassaemia patients receiving des- lates p53-dependent apoptosis, Mol. Cell 8 (2001) 85 – 94.
ferrioxamine, Eur. J. Haematol. 60 (1998) 21 – 27. [166] M.B. Kastan, Q. Zhan, W.S. el-Deiry, F. Carrier, T. Jacks, W.V.
[156] S. Tomoyasui, K. Fukuchi, K. Yajima, K. Watanabe, H. Suzuki, K. Walsh, B.S. Plunkett, B. Vogelstein , A.J. Fornace Jr., A mammalian
Kawakami, K. Gomi, N. Tsuruoka, Suppression of HL-60 cell pro- cell cycle checkpoint pathway utilizing p53 and GADD45 is defec-
liferation by deferoxamine: changes in c-myc expression, Anticancer tive in ataxia – telangiectasia, Cell 71 (1992) 587 – 597.
Res. 13 (1993) 407 – 410. [167] S. de Rozieres, R. Maya, M. Oren, G. Lozano, The loss of mdm2
[157] K.J. Wu, A. Polack, R. Dalla-Favera, Coordinated regulation of iron- induces p53-mediated apoptosis, Oncogene 19 (2000) 1691 – 1697.
controlling genes, H-ferritin and IRP2, by c-myc, Science 283 (1999)
676 – 679.

You might also like