You are on page 1of 11

Curr. Med. Chem.

- Anti-Cancer Agents, 2005, 5, 363-372 363

Etoposide, Topoisomerase II and Cancer

E.L. Baldwin1 and N. Osheroff1,2,*

Departments of 1Biochemistry and 2Medicine (Hematology/Oncology), Vanderbilt University School of Medicine,


Nashville TN 37232-0146, USA

Abstract: Etoposide is an important chemotherapeutic agent that is used to treat a wide spectrum of human cancers. It has
been in clinical use for more than two decades and remains one of the most highly prescribed anticancer drugs in the
world. The primary cytotoxic target for etoposide is topoisomerase II. This ubiquitous enzyme regulates DNA under- and
overwinding, and removes knots and tangles from the genome by generating transient double-stranded breaks in the
double helix. Etoposide kills cells by stabilizing a covalent enzyme-cleaved DNA complex (known as the cleavage
complex) that is a transient intermediate in the catalytic cycle of topoisomerase II. The accumulation of cleavage
complexes in treated cells leads to the generation of permanent DNA strand breaks, which trigger recombination/repair
pathways, mutagenesis, and chromosomal translocations. If these breaks overwhelm the cell, they can initiate death
pathways. Thus, etoposide converts topoisomerase II from an essential enzyme to a potent cellular toxin that fragments
the genome. Although the topoisomerase II-DNA cleavage complex is an important target for cancer chemotherapy, there
also is evidence that topoisomerase II-mediated DNA strand breaks induced by etoposide and other agents can trigger
chromosomal translocations that lead to specific types of leukemia. Given the central role of topoisomerase II in both the
cure and initiation of human cancers, it is imperative to further understand the mechanism by which the enzyme cleaves
and rejoins the double helix and the process by which etoposide and other anticancer drugs alter topoisomerase II
function.
Key Words: Etoposide, topoisomerase II, anticancer drugs, DNA cleavage, DNA ligation, DNA cleavage complex, cancer
chemotherapy, leukemia.

A. INTRODUCTION It also will discuss therapy-related leukemias that are


associated with the use of etoposide.
Etoposide is one of the most widely prescribed anticancer
drugs in the world. It is used to treat a variety of cancers, B. HISTORY OF ETOPOSIDE
including small cell lung cancer, germ-line malignancies,
sarcomas, leukemias, and lymphomas. The primary cellular Etoposide is derived from podophyllotoxin (Fig. (1)) [1,
target for etoposide is topoisomerase II. This essential 2]. This natural product is found in the plant Podophyllum
enzyme removes knots and tangles from the genome by peltatum, more commonly known as mandrake or May apple.
introducing transient double-stranded breaks in the DNA Podophyllotoxin has been used as a folk remedy for a variety
backbone. of illnesses for over a thousand years [2]. It has well estab-
lished antimitotic properties, which are related to its ability
Etoposide kills cells in an insidious fashion. Rather than to inhibit tubulin polymerization. Although podophyllotoxin
robbing the cell of the critical functions of topoisomerase II, demonstrated potential as an antineoplastic agent, clinical
the drug stabilizes a covalent enzyme-cleaved DNA complex development was prevented by its high toxicity [2-4].
(known as the cleavage complex) that is a requisite In the 1950’s, investigators at Sandoz Pharmaceuticals
intermediate in the catalytic cycle of topoisomerase II. The initiated a program in the hopes of overcoming the prohibi-
accumulation of cleavage complexes in treated cells leads to tive toxicity of the parent compound [1, 2]. As a result, a
the generation of permanent breaks in the genetic material, series of semi-synthetic podophyllotoxin derivatives was
which ultimately trigger cell death pathways. Thus, etopo- generated. Ultimately, two analogs with increased antineo-
side converts this essential enzyme into a potent cellular plastic activity and decreased toxicity were identified in the
toxin that fragments the genome. mid-1960’s, etoposide and teniposide (Fig. (1)) [1, 2].
Following pre-clinical development and a series of clinical
This review will focus on etoposide, its interactions with
trials, the United States Food and Drug Administration
topoisomerase II, and the cellular pathways that are involved
approved the use of etoposide for cancer chemotherapy in
in processing the DNA breaks that are induced by the drug. 1983. Remarkably, at the time of this approval, the mech-
anism of action of the drug was unknown. In contrast to
podophyllotoxin, etoposide displayed no ability to inhibit
*Address correspondence to this author at the Department of Biochemistry,
654 Robinson Research Building, Vanderbilt University School of
tubulin polymerization [5]. It was not until the mid-1980’s
Medicine, Nashville, TN 37232-0146, USA; Tel: (615)322-4338; Fax: that it was determined that the primary cellular target of
(615)343-1166; E-mail: neil.osheroff@vanderbilt.edu etoposide was topoisomerase II [6-8].

1568-0118/05 $50.00+.00 © 2005 Bentham Science Publishers Ltd.


364 Curr. Med. Chem. – Anti-Cancer Agents, 2005, Vol. 5, No. 4 Baldwin and Osheroff

g8
OH g7
H3 C
g6
O O g
4
O
O g5
O
g2
O HO g3
O
O g1
OH
11
O 5 4
6 3
A B C D O12
7 8 2
CH3O OCH3 1 13
O
OCH3 O
1' 6'
2'
3' E
Podophyllotoxin Etoposide 4' 5'
H3CO OCH3
OH

H3 C CH3
H N
O
O
S O
HO
H3 C N
OH
O

O O
O O
O O
O O

CH3O OCH3 CH3O OCH3


OH OH

Teniposide TOP-53

Fig. (1). Structures of podophyllotoxin, etoposide, teniposide and TOP-53. Etoposide can be separated into three distinct structural domains:
a glycosidic group substituted at the C-4 position, a polycyclic core (A-D rings), and a pendant ring (E-ring). Teniposide contains of a
thiophene moiety in place of the methyl group at the g-8 position, whereas TOP-53 contains an amino-alkyl side chain at the C-4 position.

C. TOPOISOMERASE II Topoisomerase IIα is the isoform originally described in


vertebrate species and in general, represents the enzyme that
1. Isoforms
was characterized as “topoisomerase II” in early studies in
Topoisomerase II is a ubiquitous enzyme that regulates these organisms. Since the enzymological properties of all
DNA topology by passing an intact double helix through a eukaryotic type II enzymes appear to be similar and distinc-
transient double-stranded break that it generates in the back- tions between isoforms are not always apparent, unless
bone of a separate DNA segment [9-15]. In contrast to lower specifically stated otherwise, the term topoisomerase II will
eukaryotes such as yeast and Drosophila that encode a single be used to refer to all members of this enzyme family.
type II enzyme [16, 17], vertebrates have two isoforms of
topoisomerase II, α and β [18, 19]. These two isoforms share 2. Catalytic Cycle
~70% amino acid sequence identity and display similar
enzymatic activities. However, they are encoded by separate One round of catalysis mediated by topoisomerase II is
genes, and differ in their protomer molecular masses (~170 depicted in Fig. (2) (see Refs. [10-13, 15] for comprehensive
kDa and ~180 kDa, respectively), physiological regulation, reviews of the topoisomerase II catalytic cycle). A brief
and cellular functions [18-25]. Topoisomerase IIα levels description follows: 1) Topoisomerase II binds two segments
increase dramatically during periods of cell growth, and this of DNA to initiate a round of catalysis [35, 36]. 2) In the
isoform appears to be primarily responsible for the required presence of a divalent cation [37], the two subunits of the
roles of the enzyme during DNA replication and mitosis [26- enzyme each cleave one strand of the double helix [38-40].
31]. Topoisomerase IIα is an essential enzyme [23, 32, 33]. The two scissile bonds are located four base pairs apart
Vertebrate cells can survive in the absence of topoisomerase directly across the major groove from one another. Thus,
IIβ, but this isoform is required for proper neurological cleavage generates DNA molecules with 4-base single-
development in mouse embryos [34]. stranded cohesive ends on their 5'-termini [38, 39]. It is
Etoposide, Topoisomerase II and Cancer Curr. Med. Chem. – Anti-Cancer Agents, 2005, Vol. 5, No. 4 365

Fig. (2). Catalytic cycle of topoisomerase II. One round of catalysis is shown as a series of six discrete steps. 1) Topoisomerase II-DNA
binding; 2) pre-strand passage DNA cleavage/religation equilibrium; 3) ATP binding and DNA strand passage; 4) post-strand passage DNA
cleavage-religation equilibrium; 5) ATP hydrolysis and gate opening; 6) DNA release and enzyme turnover. Adapted from [13].

notable that this reaction is reversible and that the enzyme II, it is the primary link between the enzyme and cancer. The
establishes a DNA cleavage-religation equilibrium. 3) topoisomerase II-DNA cleavage complex is the cytotoxic
Topoisomerase II binds two ATP molecules and undergoes a target for etoposide and other anticancer drugs and also
conformational change that causes the passage of the intact appears to play a central role in the initiation of leukemic
DNA segment through the cleaved double helix [41, 42]. chromosomal translocations.
DNA strand passage appears to be faster if one of the two Normally, cleavage complexes are fleeting intermediates
ATP molecules is hydrolyzed [43]. The transported DNA is in the catalytic cycle of topoisomerase II. Consequently, the
passed into a central cavity in topoisomerase II through an enzyme-linked DNA breaks are present at low steady-state
N-terminal gate [44, 45]. The gate is closed and the enzyme levels and are tolerated by cells [10, 13, 15, 49]. However,
forms a protein clamp on the DNA [44, 45]. 4) Following increases in the physiological concentration or lifetime of
DNA translocation, topoisomerase II reseals the original cleavage complexes can initiate a number of mutagenic
nucleic acid break, and the enzyme once again establishes a events (Fig. (3)) [10, 13, 15, 50-53]. For example, when
DNA cleavage-religation equilibrium [44]. 5) Topoiso- DNA tracking systems such as replication forks collide with
merase II hydrolyzes the second ATP molecule, which opens a topoisomerase II-DNA cleavage complex, the transient
a C-terminal gate on the protein and allows the transported enzyme-associated DNA break is converted to a permanent
DNA segment to be released [41, 44, 45]. 6) Finally, the double-stranded break in the genome [54-57]. Once the
enzyme returns to its original conformation and regains the DNA termini are no longer bridged by topoisomerase II, they
ability to begin a new round of catalysis [41, 44].
become targets for recombination and repair pathways.
As a result of its catalytic mechanism, topoisomerase II Depending on the pathway employed, repair of topoisomer-
can relax under- or overwound (i.e., negatively or positively ase II-mediated DNA strand breaks can generate chromo-
supercoiled) DNA molecules and can remove knots and somal insertions, deletions, translocations, or other aberra-
tangles from the genome [10-13, 15]. It is these latter func- tions [10, 50, 52, 58]. If these strand breaks are present in
tions that make topoisomerase II essential for cell survival. sufficient numbers, they initiate programmed cell death
In the absence of the type II enzyme, cells cannot segregate pathways (Fig. (3)) [51].
daughter chromosomes and die as a result of mitotic failure. Agents such as etoposide that increase the concentration
of cleavage complexes are known as topoisomerase II
3. Topoisomerase II-DNA Cleavage Complexes
poisons because they convert this essential enzyme to a
In order to maintain the integrity of the genetic material potent cellular toxin [59]. This is in contrast to catalytic
throughout the double-stranded DNA passage reaction, inhibitors, which block the overall catalytic activity of the
topoisomerase II forms covalent bonds between active site enzyme. Due to the mechanism of action of topoisomerase II
tyrosyl residues (one per each subunit of the homodimeric poisons, the higher the cellular concentration of the type II
enzyme) and the newly generated termini of the cleaved enzyme, the more potent these drugs become [10, 13, 15,
DNA [38-40, 46-48]. This covalent enzyme-cleaved DNA 60]. The high levels of topoisomerase II often expressed in
complex, which is referred to as the cleavage complex, is a cancer cells provide part of the therapeutic window for
hallmark of all type II enzymes. Beyond the essential role of etoposide and other chemotherapeutic agents that target this
the cleavage complex in the catalytic cycle of topoisomerase enzyme.
366 Curr. Med. Chem. – Anti-Cancer Agents, 2005, Vol. 5, No. 4 Baldwin and Osheroff

Fig. (3). Effects of topoisomerase II-DNA cleavage complexes in the cell. Topoisomerase II-DNA cleavage complexes normally are
transient intermediates in the catalytic cycle of the enzyme and are present at low cellular concentrations. However, when levels of cleavage
complexes rise, they can be converted to permanent DNA strand breaks that trigger DNA recombination/repair pathways or cell death
pathways. Alternatively, chromosomal translocations can be generated that lead to specific types of leukemia.

D. ACTIONS OF ETOPOSIDE AGAINST TOPOISO- The final conclusive evidence that etoposide killed cells
MERASE II by acting as a topoisomerase II poison came from studies
that used the budding yeast, Saccharomyces cerevisiae. This
1. Cellular Target
organism allowed, for the first time, the cellular effects of
As discussed earlier, etoposide kills cells by increasing etoposide to be studied in a series of isogenic strains. The
levels of topoisomerase II-DNA cleavage complexes [2, 10, initial work of Nitiss and colleagues determined that overex-
13, 15, 60, 61]. This drug was among the earliest anticancer pression of topoisomerase II rendered yeast cells hypersen-
agents that were identified as targeting the type II enzyme. sitive to etoposide [72]. Subsequent experiments demon-
Several lines of evidence led to this identification. strated that yeast expressing a temperature-sensitive mutant
of the enzyme were refractory to drug treatment at semiper-
By the early 1980’s, it was well established that treat- missive temperatures (at which topoisomerase II activity is
ment of mammalian cells with etoposide induced the forma- reduced to less than 10% that of the wild-type enzyme) [73].
tion of protein-associated single-and double-stranded DNA
These two studies were seminal to our understanding of
breaks [62-64]. It was not until the mid-1980’s, however, etoposide action. Together they provided the first direct
that the role of topoisomerase II in mediating the toxicity of evidence that topoisomerase II is the cytotoxic target of
the drug became apparent. The first direct evidence that the etoposide and that the drug kills cells by converting the
enzyme was the target of etoposide came from in vitro
enzyme into a physiological toxin. As further evidence that
studies in the Liu laboratory [6, 7]. Inclusion of etoposide in topoisomerase II is the primary cellular target of etoposide,
topoisomerase II-DNA cleavage assays significantly raised yeast strains that expressed mutant type II enzymes that were
levels of enzyme-mediated strand breaks. Subsequent studies either resistant or hypersensitive to etoposide in vitro
demonstrated that DNA breaks generated following treat-
displayed the corresponding drug phenotype in vivo [74-78].
ment of cells with the drug were attached to topoisomerase
II. The initial work utilized SV40 as a model system [8], but The relative contributions of topoisomerase IIα and β to
later experiments extended this result to chromosomal DNA the actions of etoposide in mammalian cells are still under
[65-67]. debate. The drug has similar effects on DNA cleavage
mediated by the two isoforms in vitro [19, 67, 79] and
Mutagenesis studies also played an important role in the appears to target both enzymes in cultured human cells [65,
identification of topoisomerase II as the primary cellular 67]. Thus, it is likely that topoisomerase IIα and β both play
target of etoposide. Mammalian cell lines selected for a role in mediating the cellular efficacy of etoposide, and that
resistance to etoposide or teniposide were demonstrated to the relative contributions of the two isoforms are dependent
express a type II enzyme that displayed resistance to these on the tumor and the individual patient.
drugs in vitro [68-71].
Etoposide, Topoisomerase II and Cancer Curr. Med. Chem. – Anti-Cancer Agents, 2005, Vol. 5, No. 4 367

2. Topoisomerase II-DNA-Etoposide Ternary Complex Irrespective of the precise mechanism, evidence suggests
Formation that etoposide acts in close proximity to the scissile bonds
[85-88, 92]. Furthermore, it appears that the interaction of
The topoisomerase II-DNA-etoposide ternary complex
etoposide with one scissile bond is independent from its
can form by one of three different pathways: etoposide can
interaction with the other, and that each drug molecule
enter the complex through interactions with the enzyme,
with DNA, or with the binary topoisomerase II-DNA com- stabilizes only a single-stranded break [92]. Thus, two drug
plex [10, 13, 80]. Originally, it was believed that anticancer molecules are required in order to increase levels of topo-
drugs entered the complex through interactions with either isomerase II-mediated double-stranded DNA breaks [92].
the DNA or the enzyme-DNA binary complex. However, all This finding explains the fact that etoposide produces high
levels of enzyme-associated single-stranded DNA breaks at
currently available evidence suggests that etoposide enters
physiologically relevant drug concentrations [62-64, 92]. In
the complex through interactions with topoisomerase II.
addition, it implies that the cytotoxic effects of etoposide
First, in contrast to most other topoisomerase II poisons,
may be more closely associated with the generation of topo-
etoposide displays little ability to bind free DNA [81].
isomerase II-mediated single-, rather than double-stranded
Second, the drug binds to yeast topoisomerase II and human
topoisomerase IIα in a stoichiometric fashion in the absence DNA breaks. These single-stranded DNA breaks subse-
of DNA [82, 83]. Furthermore, the concentrations of etopo- quently are converted to permanent double-stranded DNA
side required for binding approximate the levels required to breaks by the actions of replication forks. Consistent with
this suggestion, etoposide activates the ATR- rather than
increase levels of topoisomerase II-DNA cleavage com-
plexes [82, 83]. Third, the binding affinity of etoposide for a ATM-dependent DNA damage checkpoint in Xenopus egg
extracts which were not undergoing active DNA replication
mutant drug-resistant enzyme is decreased as compared to
[96].
wild-type topoisomerase II [77, 82]. Finally, rates of topo-
isomerase II-mediated DNA cleavage are higher when etopo-
4. Structure-Activity Relationships
side is incubated with the enzyme prior to the addition of
DNA, as compared to those generated when the drug is incu- Etoposide can be separated into three distinct structural
bated with the DNA prior to the addition of topoisomerase II domains: a glycosidic group (at the C-4 position), a
[84]. polycyclic core (A-D rings), and a pendant ring (E-ring) (see
Fig. (1)) [97]. The glycosidic group contributes little to the
The above notwithstanding, it is likely that etoposide
activity of the drug against topoisomerase II [98]. However,
contacts DNA in the ternary complex. For example, the drug
alters DNA cleavage site utilization by topoisomerase II and it plays important roles in the physiology of the compound.
If the glycosidic group is removed, the remaining demethyl-
preferentially induces scission at sites that have a cytosine
epipodophyllotoxin displays strong interactions with tubulin
one base 5’ to the scissile bonds [85-89].
[1, 5, 99]. Furthermore, exchanging the g-8 methyl group of
etoposide for a thiophene moiety in teniposide increases the
3. Mechanism of Action physiological potency of the drug by ~10–fold [1, 100-102].
Topoisomerase II poisons increase levels of cleavage If the glycosidic group at the C-4 position is replaced by an
complexes by two non-mutually exclusive mechanisms [10, amino-alkyl side chain, such as that found in TOP-53 (Fig.
13, 15]. Some drugs act by inhibiting the ability of topoiso- (1)), activity against topoisomerase II rises significantly [67].
merase II to ligate cleaved DNA. Agents in this class not Other modifications at the C-4 position also alter the activity
only raise the concentration of cleavage complexes, but of the drug against topoisomerase II both in cells and in vitro
increase their lifetime as well. Other drugs have little effect [103-105]. Thus, while the C-4 glycosidic group in etoposide
on the ligation activity of topoisomerase II and are presumed does not appear to promote drug interactions with topoiso-
to act primarily by increasing the forward rate of enzyme- merase II, modifications at this position can significantly
catalyzed DNA cleavage. affect drug activity.
On the basis of assays that monitor the ability of topoiso- The polycyclic core of etoposide plays an important role
merase II to religate DNA that it has cleaved or to ligate two in promoting drug-enzyme interactions [98]. There is little
separate fragments of DNA, etoposide acts primarily by evidence that etoposide binds free nucleic acids [81].
inhibiting the ligation activity of the type II enzyme [90-92]. However, it is possible that portions of the polycyclic core
Although the mechanism of drug action against topoiso- interact with DNA in the active site of topoisomerase II [15,
merase II had long been debated, etoposide was the first 95].
compound to be analyzed experimentally [90]. Substituents on the pendant E-ring appear to be essential
Unfortunately, none of the available crystal structures for for the activity of etoposide. The 4’-OH group is very
the DNA cleavage-ligation domain of topoisomerase II important for increasing levels of topoisomerase II-DNA
include either DNA or anticancer drugs [93, 94]. Therefore, cleavage complexes. If this group is removed or converted to
it is not known precisely how etoposide impairs the ability of a methoxy group, the activity of etoposide against topoiso-
topoisomerase II to seal DNA strand breaks. The drug may merase II is reduced significantly [62, 98-100, 106]. Removal
interfere with noncovalent enzyme-DNA interactions. Alter- of both methoxy groups (at the 3’- and 5’-positions) decre-
natively, it may alter the position of the DNA termini to be ases drug activity, however, the loss of only one of these
ligated or insert between the DNA termini and act as a groups appears to have little effect [98, 107]. In addition,
physical block to ligation [15, 95]. replacement of the 3’- and 5’-methoxy groups by hydroxyl
groups increases drug activity [102].
368 Curr. Med. Chem. – Anti-Cancer Agents, 2005, Vol. 5, No. 4 Baldwin and Osheroff

Finally, one of the methoxy groups on the E-ring can be This pathway is dependent on the presence of direct repeats
converted to a hydroxyl group in vivo by the actions of (or closely related sequences) proximal to and flanking the
cytochrome P450 3A4 (CYP 3A4) [108-111]. The resulting initial break site. It relies on Rad52p and the Rad1p/Rad10p
etoposide catechol, which has been found in patients that endonuclease. Single-strand annealing is not an error-free
were treated with etoposide [112, 113], can be further pathway and deletes one of the repeated sequences, as well
oxidized to form etoposide quinone [108-111]. Both meta- as the genetic information that is located between them.
bolites display higher activity against topoisomerase II than Finally, the break can be rejoined by the nonhomologous
the parent compound [89]. end-joining pathway [114, 115, 118, 119]. This pathway
utilizes Ku70p/Ku80p and Lig4p and results in the loss of
E. CELLULAR REPAIR OF ETOPOSIDE-INDUCED sequences proximal to the original DNA break. If multiple
DNA DAMAGE breaks are present in the genome, nonhomologous end-join-
ing can lead to the formation of chromosomal rearrange-
Double-stranded breaks in the genetic material are ments or translocations. In general, homologous recombina-
repaired primarily by DNA recombination pathways. S. tion pathways are considerably more active than nonho-
cerevisiae has been used as a model system to investigate the mologous end-joining in S. cerevisiae [114-119].
pathways by which etoposide-induced DNA damage is
repaired. The most common pathways used by budding yeast Two lines of evidence strongly suggest that the single-
are depicted in Fig. (4) [114-116]. The initial processing of strand invasion pathway of homologous recombination is the
double-stranded DNA breaks utilizes the Rad50p/Mre11p/ primary pathway used to repair topoisomerase II-associated
Xrs2p complex to generate single-stranded ends at the site of DNA breaks that are induced by etoposide. First, deletion of
the break [114-117]. Following this processing, the DNA can RAD54 significantly decreases cell survival and the induc-
be shuttled into three distinct recombination pathways. The tion of recombination/repair in the presence of etoposide. In
break can be repaired by the single-strand invasion pathway contrast, deletion of RAD1 or KU70 does not sensitize yeast
of homologous recombination [114-116]. This pathway, cells to the drug [120]. Second, a genome-wide screen for
which utilizes Rad51p/52p/54p/55p/57p as well as the hypersensitivity to etoposide identified every member of the
replication machinery, is capable of repairing the initial single-strand invasion pathway, but did not identify genes
double-stranded DNA break in an error-free manner. Alter- that were exclusive to either the single-strand annealing
natively, the break can be repaired by the single-strand pathway or nonhomologous end-joining [121]. In addition to
annealing pathway of homologous recombination [114-116]. the single-strand invasion pathway, recent evidence suggests

Fig. (4). Pathways that repair double-stranded DNA breaks in S. cerevisiae. Components of the pathways that play integral roles in
homologous recombination (single-strand invasion or single-strand annealing) and nonhomologous end-joining are shown. Adapted from
[120].
Etoposide, Topoisomerase II and Cancer Curr. Med. Chem. – Anti-Cancer Agents, 2005, Vol. 5, No. 4 369

an important role for MMS22 in the repair of etoposide- exons 5-11 of the mixed lineage leukemia (MLL) gene at
induced DNA damage in yeast [121]. This gene is not part of chromosome band 11q23 [52, 58, 135, 142-147]. Transloca-
the single-strand invasion pathway and is thought to be tions involving chromosomal band 11q23 also are observed
involved in the resolution of aberrant DNA structures that in primary human CD34+ cells following treatment with
accumulate at replication forks in response to genomic etoposide [148].
damage [121, 122]. MLL translocations involve partner genes on a variety of
Studies using the fission yeast, Schizosaccharomyces different chromosomes [52, 58, 144, 145, 147, 149, 150].
pombe, also suggest that homologous recombination path- However, the der(11) chromosomes in all of these trans-
ways play a major role in repairing topoisomerase II- locations express a chimeric protein consisting of the N-
mediated DNA breaks that are induced by the etoposide terminus of MLL and the C-terminus of the partner protein.
derivative TOP-53 [123]. Furthermore, they implicate The mechanism through which the chimeric fusion protein
nucleotide excision repair proteins in the processing of these leads to leukemia is not known. It has been suggested that
DNA breaks [123]. the fusion gene encodes an inactive MLL protein that acts as
a dominant negative to block the actions of the normal
Nonhomologous end-joining appears to play a more protein [151]. It also has been proposed that the translocation
significant role in the repair of double-stranded DNA breaks partner protein provides an active functional role [152].
in mammalian species [124]. A number of studies indicate
that homologous recombination and nonhomologous end- Several lines of evidence point to a direct role for
joining pathways both are involved in the repair of topoiso- topoisomerase II-mediated DNA cleavage in the initiation of
merase II-mediated DNA damage induced by etoposide. at least some MLL translocations. Treatment-related leuke-
Evidence for the importance of homologous recombination mias with 11q23 chromosomal rearrangements are observed
comes largely from studies on the Rad51 protein (which is primarily in patients that have been treated with regimens
homologous to S. cerevisiae Rad51p) [125-129]. There is a that include etoposide [52, 58, 141]. Similar chromosomal
strong correlation between levels of this protein and rearrangements also are seen in patients that receive other
sensitivity to etoposide in human small cell lung cancer cells. topoisomerase II-targeted drugs as part of their therapy [52,
The lower the level of Rad51 protein, the more sensitive 58, 141]. In addition, all of the DNA breakpoints mapped in
cells are to the drug and the greater the physiological patient samples are located in close proximity to topoiso-
concentration of DNA breaks [129]. In addition, treatment of merase II-DNA cleavage sites [89, 153-157]. Furthermore,
mammalian cell lines with etoposide induces the formation these chromosomal translocations often involve precise or
of Rad51 nuclear foci [125, 126]. near-precise rearrangements between topoisomerase II
cleavage sites on chromosome 11 and its partners [157, 158].
Evidence for the importance of nonhomologous end- Finally, as discussed above, etoposide is converted to its
joining comes from studies on the Ku70/80 and Lig4 more active catechol (and subsequent quinone) metabolite by
proteins. In a variety of mammalian cell lines, deficiencies in CYP3A4 [108-111]. Individuals who carry a polymorphism
any of these proteins result in hypersensitivity to etoposide in the promoter region of the CYP3A4 gene display a decre-
[130]. Deletion of DNA-PKcs (the catalytic subunit, which ased probability of developing treatment-related leukemias
together with the Ku70/80 proteins, comprises the DNA-PK [159].
complex [131]) shows a less severe phenotype [130]. In this
regard, it is notable that the Ku70 protein does not recognize It should be noted that recent studies with human cell
the DNA in a topoisomerase II cleavage complex unless the lines indicate that DNA breaks within the MLL gene and
enzyme is first proteolyzed [132]. Similarly, the enzyme has tranlocations involving this locus are observed in cells that
to be processed from the DNA termini in order for DNA-PK survive apoptotic events [160-162]. Therefore, it is possible
to be activated [132]. that the etoposide-induced breakpoints observed in deriva-
tive chromosomes do not arise directly from the drug-
Finally, studies using mutant chicken cell lines suggest stabilized DNA breaks mediated by topoisomerase II, but
that nonhomologous end-joining plays a more important role rather from DNA breaks generated by apoptotic processes. In
than homologous recombination in the repair of etoposide- this latter case, increased cellular levels of cleavage com-
induced topoisomerase II-mediated DNA breaks [133, 134]. plexes would serve primarily to initiate apoptosis. However,
Whereas RAD54 null cells are 10- to 100-fold hypersensitive since topoisomerase II has been implicated in excising DNA
to etoposide, KU70 or LIG4 null cells are 1,000- to 10,000- loop domains during apoptosis [163, 164], topoisomerase II
fold hypersensitive the drug [133]. could still be generating the apoptotic DNA breaks that
eventually recombine to form the chromosomal transloca-
F. INITIATION OF LEUKEMIA BY ETOPOSIDE tions.
Despite the importance of etoposide in the chemothera- In addition to treatment-related leukemias, translocations
peutic treatment of human cancers, considerable evidence that involve the MLL gene are associated with infant acute
suggests that the drug also triggers chromosomal trans- myeloid leukemias (AMLs) [147, 165-167]. The chromo-
locations that lead to specific types of leukemia (see Fig. (3)) somal translocations associated with these cancers are
[52, 58, 135-141]. It is estimated that 2-3% of patients who observed in utero [167-169]. Epidemiological studies indi-
receive drug regimens that include etoposide eventually cate that maternal consumption of foods during pregnancy
develop treatment-related leukemias [140]. The majority of that are high in naturally occurring topoisomerase II poisons
etoposide-related leukemias display translocations with increase the risk of developing infant AMLs ~10-fold [170,
breakpoints that cluster within an 8.3 kb region between 171].
370 Curr. Med. Chem. – Anti-Cancer Agents, 2005, Vol. 5, No. 4 Baldwin and Osheroff

The ability of etoposide and other topoisomerase II [20] Tsai-Pflugfelder, M.; Liu, L. F.; Liu, A. A.; Tewey, K. M.; Whang-
poisons to cure rather than cause cancer may be related to Peng, J.; Knutsen, T.; Huebner, K.; Croce, C. M.; Wang, J. C.
Proc. Natl. Acad. Sci. USA, 1988, 85, 7177-7181.
cellular levels of topoisomerase II-mediated DNA cleavage [21] Jenkins, J. R.; Ayton, P.; Jones, T.; Davies, S. L.; Simmons, D. L.;
and the activity of cell death pathways. If the concentration Harris, A. L.; Sheer, D.; Hickson, I. D. Nucleic Acids Res., 1992,
of enzyme-associated DNA breaks is too low to overwhelm 20, 5587-5592.
recombination/repair pathways or to trigger cell death, the [22] Tan, K. B.; Dorman, T. E.; Falls, K. M.; Chung, T. D.; Mirabelli,
C. K.; Crooke, S. T.; Mao, J. Cancer Res., 1992, 52, 231-234.
pathways that promote cell survival can lead to the formation [23] Austin, C. A.; Marsh, K. L. BioEssays, 1998, 20, 215-226.
of the chromosomal translocations that ultimately lead to [24] Nitiss, J. L. Biochim. Biophys. Acta, 1998, 1400, 63-81.
cancerous growth. [25] Wang, J. C. Nat. Rev. Mol. Cell. Biol., 2002, 3, 430-440.
[26] Woessner, R. D.; Mattern, M. R.; Mirabelli, C. K.; Johnson, R. K.;
G. CONCLUSIONS AND PERSPECTIVES Drake, F. H. Cell Growth Differ., 1991, 2, 209-214.
[27] Sullivan, D. M.; Latham, M. D.; Ross, W. E. Cancer Res. , 1987,
Etoposide is an important chemotherapeutic agent that is 47, 3973-3979.
[28] Heck, M. M.; Hittelman, W. N.; Earnshaw, W. C. Proc. Natl. Acad.
used to treat a wide spectrum of human cancers. However, Sci. USA, 1988, 85, 1086-1090.
there is also evidence that the drug has the capacity to trigger [29] Hsiang, Y. H.; Wu, H. Y.; Liu, L. F. Cancer Res., 1988, 48, 3230-
distinct leukemias that are associated with chromosomal 3235.
translocations involving the MLL gene. Both of these conse- [30] Isaacs, R. J.; Davies, S. L.; Sandri, M. I.; Redwood, C.; Wells, N.
quences of etoposide treatment appear arise from the ability J.; Hickson, I. D. Biochim. Biophys. Acta, 1998, 1400, 121-137.
[31] Carpenter, A. J.; Porter, A. C. Mol. Biol. Cell, 2004, 15, 5700-
of the drug to stabilize topoisomerase II-DNA cleavage 5711.
complexes. Given the clinical prevalence of etoposide, it is [32] Dereuddre, S.; Delaporte, C.; Jacquemin-Sablon, A. Cancer Res.,
critical to fully understand the mechanism by which this 1997, 57, 4301-4308.
drug affects the catalytic functions of topoisomerase II and [33] Grue, P.; Grasser, A.; Sehested, M.; Jensen, P. B.; Uhse, A.;
Straub, T.; Ness, W.; Boege, F. J. Biol. Chem., 1998, 273, 33660-
the cellular processes that repair topoisomerase II-mediated 33666.
DNA breaks. With this knowledge, it may be possible to [34] Yang, X.; Li, W.; Prescott, E. D.; Burden, S. J.; Wang, J. C.
improve the anticancer properties of etoposide while Science, 2000, 287, 131-134.
decreasing the risk of treatment-related leukemias. [35] Zechiedrich, E. L.; Osheroff, N. EMBO J., 1990, 9, 4555-4562.
[36] Roca, J.; Berger, J. M.; Wang, J. C. J. Biol. Chem., 1993, 268,
14250-14255.
ACKNOWLEDGEMENTS [37] Osheroff, N. Biochemistry, 1987, 26, 6402-6406.
[38] Liu, L. F.; Rowe, T. C.; Yang, L.; Tewey, K. M.; Chen, G. L. J.
The authors are grateful to Jennifer S. Dickey and Renier Biol. Chem., 1983, 258, 15365-15370.
Vélez-Cruz for the critical reading of this manuscript. The [39] Sander, M.; Hsieh, T. J. Biol. Chem., 1983, 258, 8421-8428.
senior author’s laboratory is supported by National Institutes [40] Zechiedrich, E. L.; Christiansen, K.; Andersen, A. H.; Westergaard,
of Health research grants GM33944 and GM53960. E.L.B. O.; Osheroff, N. Biochemistry, 1989, 28, 6229-6236.
was a trainee under National Institutes of Health grant 5 T32 [41] Osheroff, N.; Shelton, E. R.; Brutlag, D. L. J. Biol. Chem., 1983,
258, 9536-9543.
CA09582. [42] Lindsley, J. E.; Wang, J. C. Proc. Natl. Acad. Sci. USA, 1991, 88,
10485-10489.
REFERENCES [43] Baird, C. L.; Harkins, T. T.; Morris, S. K.; Lindsley, J. E. Proc.
Natl. Acad. Sci. USA, 1999, 96, 13685-13690.
[1] Stahelin, H. F.; von Wartburg, A. Cancer Res., 1991, 51, 5-15. [44] Osheroff, N. J. Biol. Chem., 1986, 261, 9944-9950.
[2] Hande, K. R. Eur. J. Cancer, 1998, 34, 1514-1521. [45] Roca, J.; Wang, J. C. Cell, 1992, 71, 833-840.
[3] Vogelzang, N. J.; Raghavan, D.; Kennedy, B. J. Am. J. Med., 1982, [46] Rowe, T. C.; Chen, G. L.; Hsiang, Y. H.; Liu, L. F. Cancer Res.,
72, 136-144. 1986, 46, 2021-2026.
[4] Sinkule, J. A. Pharmacotherapy, 1984, 4, 61-73. [47] Horowitz, D. S.; Wang, J. C. J. Biol. Chem., 1987, 262, 5339-5344.
[5] Loike, J. D.; Horwitz, S. B. Biochemistry, 1976, 15, 5435-5443. [48] Worland, S. T.; Wang, J. C. J. Biol. Chem., 1989, 264, 4412-4416.
[6] Ross, W.; Rowe, T.; Glisson, B.; Yalowich, J.; Liu, L. Cancer Res., [49] Wang, J. C. Annu. Rev. Biochem., 1996, 65, 635-692.
1984, 44, 5857-5860. [50] Baguley, B. C.; Ferguson, L. R. Biochim. Biophys. Acta, 1998,
[7] Chen, G. L.; Yang, L.; Rowe, T. C.; Halligan, B. D.; Tewey, K. M.; 1400, 213-222.
Liu, L. F. J. Biol. Chem., 1984, 259, 13560-13566. [51] Kaufmann, S. H. Biochim. Biophys. Acta, 1998, 1400, 195-211.
[8] Yang, L.; Rowe, T. C.; Liu, L. F. Cancer Res., 1985, 45, 5872- [52] Felix, C. A. Biochim. Biophys. Acta, 1998, 1400, 233-255.
5876. [53] Holden, J. A. Curr. Med. Chem. Anti-Canc. Agents, 2001, 1, 1-25.
[9] Wang, J. C. Quart. Rev. Biophys., 1998, 31, 107-144. [54] Holm, C.; Covey, J. M.; Kerrigan, D.; Pommier, Y. Cancer Res. ,
[10] Burden, D. A.; Osheroff, N. Biochim. Biophys. Acta, 1998, 1400, 1989, 49, 6365-6368.
139-154. [55] D'Arpa, P.; Beardmore, C.; Liu, L. F. Cancer Res., 1990, 50, 6919-
[11] Berger, J. M. Biochim. Biophys. Acta, 1998, 1400, 3-18. 6924.
[12] Berger, J. M. Curr. Opin. Struct. Biol., 1998, 8, 26-32. [56] Corbett, A. H.; Osheroff, N. Chem. Res. Toxicol., 1993, 6, 585-597.
[13] Fortune, J. M.; Osheroff, N. Prog. Nucleic Acid Res. Mol. Biol., [57] Howard, M. T.; Neece, S. H.; Matson, S. W.; Kreuzer, K. N. Proc.
2000, 64, 221-253. Natl. Acad. Sci. USA, 1994, 91, 12031-12035.
[14] Champoux, J. J. Annu. Rev. Biochem., 2001, 70, 369-413. [58] Felix, C. A. Med. Pediatr. Oncol., 2001, 36, 525-535.
[15] Wilstermann, A. M.; Osheroff, N. Curr. Top. Med. Chem., 2003, 3, [59] Kreuzer, K. N.; Cozzarelli, N. R. J. Bacteriol., 1979, 140, 424-435.
321-338. [60] Walker, J. V.; Nitiss, J. L. Cancer Invest., 2002, 20, 570-589.
[16] Wyckoff, E.; Natalie, D.; Nolan, J. M.; Lee, M.; Hsieh, T. J. Mol. [61] Hande, K. R. Biochim. Biophys. Acta, 1998, 1400, 173-184.
Biol., 1989, 205, 1-13. [62] Loike, J. D.; Horwitz, S. B. Biochemistry, 1976, 15, 5443-5448.
[17] Goto, T.; Wang, J. C. Cell, 1984, 36, 1073-1080. [63] Wozniak, A. J.; Ross, W. E. Cancer Res., 1983, 43, 120-124.
[18] Drake, F. H.; Zimmerman, J. P.; McCabe, F. L.; Bartus, H. F.; Per, [64] Glisson, B. S.; Smallwood, S. E.; Ross, W. E. Biochim. Biophys.
S. R.; Sullivan, D. M.; Ross, W. E.; Mattern, M. R.; Johnson, R. Acta, 1984, 783, 74-79.
K.; Crooke, S. T. J. Biol. Chem., 1987, 262, 16739-16747. [65] Willmore, E.; Frank, A. J.; Padget, K.; Tilby, M. J.; Austin, C. A.
[19] Drake, F. H.; Hofmann, G. A.; Bartus, H. F.; Mattern, M. R.; Mol. Pharmacol., 1998, 54, 78-85.
Crooke, S. T.; Mirabelli, C. K. Biochemistry, 1989, 28, 8154-8160. [66] Hsiang, Y. H.; Liu, L. F. J. Biol. Chem., 1989, 264, 9713-9715.
Etoposide, Topoisomerase II and Cancer Curr. Med. Chem. – Anti-Cancer Agents, 2005, Vol. 5, No. 4 371

[67] Byl, J. A.; Cline, S. D.; Utsugi, T.; Kobunai, T.; Yamada, Y.; [104] Arimondo, P.; Boukarim, C.; Bailly, C.; Dauzonne, D.; Monneret,
Osheroff, N. Biochemistry, 2001, 40, 712-718. C. Anticancer Drug Des., 2000, 15, 413-421.
[68] Glisson, B. S.; Sullivan, D. M.; Gupta, R.; Ross, W. E. Natl. [105] Guianvarc'h, D.; Duca, M.; Boukarim, C.; Kraus-Berthier, L.;
Cancer Inst. Monogr., 1987, 4, 89-93. Leonce, S.; Pierre, A.; Pfeiffer, B.; Renard, P.; Arimondo, P. B.;
[69] Danks, M. K.; Yalowich, J. C.; Beck, W. T. Cancer Res., 1987, 47, Monneret, C.; Dauzonne, D. J. Med. Chem., 2004, 47, 2365-2374.
1297-1301. [106] Sinha, B. K.; Politi, P. M.; Eliot, H. M.; Kerrigan, D.; Pommier, Y.
[70] Danks, M. K.; Schmidt, C. A.; Cirtain, M. C.; Suttle, D. P.; Beck, Eur. J. Cancer, 1990, 26, 590-593.
W. T. Biochemistry, 1988, 27, 8861-8869. [107] Saulnier, M. G.; Vyas, D. M.; Langley, D. R.; Doyle, T. W.; Rose,
[71] Sullivan, D. M.; Latham, M. D.; Rowe, T. C.; Ross, W. E. W. C.; Crosswell, A. R.; Long, B. H. J. Med. Chem., 1989, 32,
Biochemistry, 1989, 28, 5680-5687. 1418-1420.
[72] Nitiss, J. L.; Liu, Y. X.; Harbury, P.; Jannatipour, M.; Wasserman, [108] van Maanen, J. M.; Retel, J.; de Vries, J.; Pinedo, H. M. J. Natl.
R.; Wang, J. C. Cancer Res., 1992, 52, 4467-4472. Cancer Inst., 1988, 80, 1526-1533.
[73] Nitiss, J. L.; Liu, Y. X.; Hsiung, Y. Cancer Res., 1993, 53, 89-93. [109] Mans, D. R.; Retel, J.; van Maanen, J. M.; Lafleur, M. V.; van
[74] Jannatipour, M.; Liu, Y. X.; Nitiss, J. L. J. Biol. Chem., 1993, 268, Schaik, M. A.; Pinedo, H. M.; Lankelma, J. Br. J. Cancer, 1990,
18586-18592. 62, 54-60.
[75] Liu, Y.-X.; Hsiung, Y.; Jannatipour, M.; Yeh, Y.; Nitiss, J. L. [110] Relling, M. V.; Nemec, J.; Schuetz, E. G.; Schuetz, J. D.;
Cancer Res., 1994, 54, 2943-2951. Gonzalez, F. J.; Korzekwa, K. R. Mol. Pharmacol., 1994, 45, 352-
[76] Nitiss, J. L.; Vilalta, P. M.; Wu, H.; McMahon, J. Mol. Pharmacol., 358.
1994, 46, 773-777. [111] Zhuo, X.; Zheng, N.; Felix, C. A.; Blair, I. A. Drug Metab. Dispos.,
[77] Elsea, S. H.; Hsiung, Y.; Nitiss, J. L.; Osheroff, N. J. Biol. Chem. , 2004, 32, 993-1000.
1995, 270, 1913-1920. [112] Stremetzne, S.; Jaehde, U.; Kasper, R.; Beyer, J.; Siegert, W.;
[78] Hsiung, Y.; Elsea, S. H.; Osheroff, N.; Nitiss, J. L. J. Biol. Chem. , Schunack, W. Eur. J. Cancer, 1997, 33, 978-979.
1995, 270, 20359-20364. [113] Relling, M. V.; Yanishevski, Y.; Nemec, J.; Evans, W. E.; Boyett,
[79] Cornarotti, M.; Tinelli, S.; Willmore, E.; Zunino, F.; Fisher, L. M.; J. M.; Behm, F. G.; Pui, C. H. Leukemia, 1998, 12, 346-352.
Austin, C. A.; Capranico, G. Mol. Pharmacol., 1996, 50, 1463- [114] Paques, F.; Haber, J. E. Microbiol. Mol. Biol. Rev., 1999, 63, 349-
1471. 404.
[80] Froelich-Ammon, S. J.; Osheroff, N. J. Biol. Chem., 1995, 270, [115] Haber, J. E. Trends Genet., 2000, 16, 259-264.
21429-21432. [116] Dudas, A.; Chovanec, M. Mutat. Res., 2004, 566, 131-167.
[81] Chow, K. C.; Macdonald, T. L.; Ross, W. E. Mol. Pharmacol., [117] Hopfner, K. P.; Putnam, C. D.; Tainer, J. A. Curr. Opin. Struct.
1988, 34, 467-473. Biol., 2002, 12, 115-122.
[82] Kingma, P. S.; Burden, D. A.; Osheroff, N. Biochemistry, 1999, 38, [118] Moore, J. K.; Haber, J. E. Mol. Cell. Biol., 1996, 16, 2164-2173.
3457-3461. [119] Lewis, L. K.; Resnick, M. A. Mutat. Res., 2000, 451, 71-89.
[83] Leroy, D.; Kajava, A. V.; Frei, C.; Gasser, S. M. Biochemistry, [120] Sabourin, M.; Nitiss, J. L.; Nitiss, K. C.; Tatebayashi, K.; Ikeda,
2001, 40, 1624-1634. H.; Osheroff, N. Nucleic Acids Res., 2003, 31, 4373-4384.
[84] Burden, D. A.; Kingma, P. S.; Froelich-Ammon, S. J.; Bjornsti, M.- [121] Baldwin, E. L.; Berger, A. C.; Corbett, A. H.; Osheroff, N. Nucleic
A.; Patchan, M. W.; Thompson, R. B.; Osheroff, N. J. Biol. Chem., Acids Res., 2005, 33, 1021-1030.
1996, 271, 29238-29244. [122] Araki, Y.; Kawasaki, Y.; Sasanuma, H.; Tye, B. K.; Sugino, A.
[85] Pommier, Y.; Capranico, G.; Orr, A.; Kohn, K. W. Nucleic Acids Genes Cells, 2003, 8, 465-480.
Res., 1991, 19, 5973-5980. [123] Malik, M.; Nitiss, J. L. Eukaryot. Cell, 2004, 3, 82-90.
[86] Capranico, G.; Zunino, F. Eur. J. Cancer, 1992, 12, 2055-2060. [124] Valerie, K.; Povirk, L. F. Oncogene, 2003, 22, 5792-5812.
[87] Pommier, Y.; Fesen, M. R.; Goldwasser, F. In Cancer [125] Haaf, T.; Raderschall, E.; Reddy, G.; Ward, D. C.; Radding, C. M.;
Chemotherapy and Biotherapy: Principles and Practice., B. A. Golub, E. I. J. Cell Biol., 1999, 144, 11-20.
Chabner; D. L. Longo, eds.; Lippincott-Raven Publishers: [126] Raderschall, E.; Golub, E. I.; Haaf, T. Proc. Natl. Acad. Sci. USA,
Philadelphia, 1996, pp. 435-461. 1999, 96, 1921-1926.
[88] Capranico, G.; Binaschi, M. Biochim. Biophys. Acta, 1998, 1400, [127] Raderschall, E.; Bazarov, A.; Cao, J.; Lurz, R.; Smith, A.; Mann,
185-194. W.; Ropers, H. H.; Sedivy, J. M.; Golub, E. I.; Fritz, E.; Haaf, T. J.
[89] Lovett, B. D.; Strumberg, D.; Blair, I. A.; Pang, S.; Burden, D. A.; Cell Sci., 2002, 115, 153-164.
Megonigal, M. D.; Rappaport, E. F.; Rebbeck, T. R.; Osheroff, N.; [128] Lundin, C.; Schultz, N.; Arnaudeau, C.; Mohindra, A.; Hansen, L.
Pommier, Y. G.; Felix, C. A. Biochemistry, 2001, 40, 1159-1170. T.; Helleday, T. J. Mol. Biol., 2003, 328, 521-535.
[90] Osheroff, N. Biochemistry, 1989, 28, 6157-6160. [129] Hansen, L. T.; Lundin, C.; Spang-Thomsen, M.; Petersen, L. N.;
[91] Robinson, M. J.; Osheroff, N. Biochemistry, 1991, 30, 1807-1813. Helleday, T. Int. J. Cancer, 2003, 105, 472-479.
[92] Bromberg, K. D.; Burgin, A. B.; Osheroff, N. J. Biol. Chem., 2003, [130] Jin, S.; Inoue, S.; Weaver, D. T. Carcinogenesis, 1998, 19, 965-
278, 7406-7412. 971.
[93] Berger, J. M.; Gamblin, S. J.; Harrison, S. C.; Wang, J. C. Nature, [131] Jackson, S. P. Int. J. Biochem. Cell. Biol., 1997, 29, 935-938.
1996, 379, 225-232. [132] Martensson, S.; Nygren, J.; Osheroff, N.; Hammarsten, O. Radiat.
[94] Fass, D.; Bogden, C. E.; Berger, J. M. Nat. Struct. Biol., 1999, 6, Res., 2003, 160, 291-301.
322-326. [133] Adachi, N.; Suzuki, H.; Iiizumi, S.; Koyama, H. J. Biol. Chem.,
[95] Wilstermann, A. M.; Osheroff, N. J. Biol. Chem., 2001, 276, 2003, 278, 35897-35902.
17727-17731. [134] Adachi, N.; Iiizumi, S.; So, S.; Koyama, H. Biochem. Biophys. Res.
[96] Costanzo, V.; Shechter, D.; Lupardus, P. J.; Cimprich, K. A.; Commun., 2004, 318, 856-861.
Gottesman, M.; Gautier, J. Mol. Cell, 2003, 11, 203-213. [135] DeVore, R.; Whitlock, J.; Hainsworth, T.; Johnson, D. Ann. Intern.
[97] Macdonald, T. L.; Legnert, E. K.; Loper, J. T.; Chow, K.-C.; Ross, Med., 1989, 110, 740-742.
W. E. In DNA Topoisomerases in Cancer, M. Potmesil; K.W. [136] Ratain, M. J.; Kaminer, L. S.; Bitran, J. D.; Lawson, R. A.; Le
Kohn, eds.; Oxford University Press: New York, 1991, pp. 199- Beau, M. M.; Skosey, C.; Purl, S.; Hoffman, P. C.; Wade, J.;
214. Vardiman, J. W.; Daly, K.; Rowley, J. D.; Golomb, H. M. Blood,
[98] Wilstermann, A. M. G., M.; Anklin, C.; Berkowitz, D. B.; Choi, S.; 1987, 70, 1412-1417.
Osheroff, N.; Graves, D. E. Proc. Am. Assoc. Cancer Res., 2003, [137] Pui, C.-H.; Ribeiro, R. C.; Hancock, M. L.; Rivera, G. K.; Evans,
135. W. E.; Raimondi, S. C.; Head, D. R.; Behm, F. G.; Mahmoud, M.
[99] Loike, J. D. Cancer Chemother. Pharmacol., 1982, 7, 103-111. H.; Sandlund, J. T.; Crist, W. M. N. Engl. J. Med., 1991, 325,
[100] Long, B. H.; Musial, S. T.; Brattain, M. G. Biochemistry, 1984, 23, 1682-1687.
1183-1188. [138] Winick, N. J.; McKenna, R. W.; Shuster, J. J.; Schneider, N. R.;
[101] Long, B. H. Sem. Oncol., 1992, 19, 3-19. Borowitz, M. J.; Bowman, W. P.; Jacaruso, D.; Kamen, B. A.;
[102] Long, B. H.; Casazza, A. M. Cancer Chemother. Pharmacol., Buchanan, G. R. J. Clin. Oncol., 1993, 11, 209-217.
1994, 34, S26-S31. [139] Sandoval, C.; Pui, C. H.; Bowman, L. C.; Heaton, D.; Hurwitz, C.
[103] Lee, K. H. J. Biomed. Sci., 1999, 6, 236-250. A.; Raimondi, S. C.; Behm, F. G.; Head, D. R. J. Clin. Oncol.,
1993, 11, 1039-1045.
372 Curr. Med. Chem. – Anti-Cancer Agents, 2005, Vol. 5, No. 4 Baldwin and Osheroff

[140] Smith, M. A.; Rubinstein, L.; Anderson, J. R.; Arthur, D.; [156] Strissel, P. L.; Strick, R.; Rowley, J. D.; Zeleznik-Le, N. J. Blood,
Catalano, P. J.; Freidlin, B.; Heyn, R.; Khayat, A.; Krailo, M.; 1998, 92, 3793-3803.
Land, V. J.; Miser, J.; Shuster, J.; Vena, D. J. Clin. Oncol., 1999, [157] Lovett, B. D.; Lo Nigro, L.; Rappaport, E. F.; Blair, I. A.; Osheroff,
17, 569-577. N.; Zheng, N.; Megonigal, M. D.; Williams, W. R.; Nowell, P. C.;
[141] Andersen, M. K.; Christiansen, D. H.; Jensen, B. A.; Ernst, P.; Felix, C. A. Proc. Natl. Acad. Sci. USA, 2001, 98, 9802-9807.
Hauge, G.; Pedersen-Bjergaard, J. Br. J. Haematol., 2001, 114, [158] Whitmarsh, R. J.; Saginario, C.; Zhuo, Y.; Hilgenfeld, E.;
539-543. Rappaport, E. F.; Megonigal, M. D.; Carroll, M.; Liu, M.;
[142] Pedersen-Bjergaard, J.; Philip, P. Blood, 1991, 78, 1147-1148. Osheroff, N.; Cheung, N. K.; Slater, D. J.; Ried, T.; Knutsen, T.;
[143] Pedersen-Bjergaard, J. Leuk. Res., 1992, 16, 733-735. Blair, I. A.; Felix, C. A. Oncogene, 2003, 22, 8448-8459.
[144] Felix, C. A.; Winick, N. J.; Negrini, M.; Bowman, W. P.; Croce, C. [159] Felix, C. A.; Walker, A. H.; Lange, B. J.; Williams, T. M.; Winick,
M.; Lange, B. J. Cancer Res., 1993, 53, 2954-2956. N. J.; Cheung, N. K.; Lovett, B. D.; Nowell, P. C.; Blair, I. A.;
[145] Felix, C. A.; Hosler, M. R.; Winick, N. J.; Masterson, M.; Wilson, Rebbeck, T. R. Proc. Natl. Acad. Sci. USA, 1998, 95, 13176-13181.
A. E.; Lange, B. J. Blood, 1995, 85, 3250-3256. [160] Stanulla, M.; Wang, J.; Chervinsky, D. S.; Thandla, S.; Aplan, P.
[146] Broeker, P. L.; Super, H. G.; Thirman, M. J.; Pomykala, H.; D. Mol. Cell. Biol., 1997, 17, 4070-4079.
Yonebayashi, Y.; Tanabe, S.; Zeleznik-Le, N.; Rowley, J. D. [161] Betti, C. J.; Villalobos, M. J.; Diaz, M. O.; Vaughan, A. T. Cancer
Blood, 1996, 87, 1912-1922. Res., 2001, 61, 4550-4555.
[147] Rowley, J. D. Annu. Rev. Gen., 1998, 32, 495-519. [162] Betti, C. J.; Villalobos, M. J.; Diaz, M. O.; Vaughan, A. T. Cancer
[148] Libura, J.; Slater, D. J.; Felix, C. A.; Richardson, C. Blood, 2004. Res., 2003, 63, 1377-1381.
[149] Super, H. J.; McCabe, N. R.; Thirman, M. J.; Larson, R. A.; Le [163] Li, T. K.; Chen, A. Y.; Yu, C.; Mao, Y.; Wang, H.; Liu, L. F.
Beau, M. M.; Pedersen-Bjergaard, J.; Philip, P.; Diaz, M. O.; Genes Dev., 1999, 13, 1553-1560.
Rowley, J. D. Blood, 1993, 82, 3705-3711. [164] Solovyan, V. T.; Bezvenyuk, Z. A.; Salminen, A.; Austin, C. A.;
[150] Domer, P. H.; Head, D. R.; Renganathan, N.; Raimondi, S. C.; Courtney, M. J. J. Biol. Chem., 2002, 277, 21458-21467.
Yang, E.; Atlas, M. Leukemia, 1995, 9, 1305-1312. [165] Felix, C. A.; Hosler, M. R.; Slater, D. J.; Parker, R. I.; Masterson,
[151] Schichman, S. A.; Canaani, E.; Croce, C. M. JAMA, 1995, 273, M.; Whitlock, J. A.; Rebbeck, T. R.; Nowell, P. C.; Lange, B. J. J.
571-576. Pediatr. Hematol. Oncol., 1998, 20, 299-308.
[152] Biondi, A.; Cimino, G.; Pieters, R.; Pui, C. H. Blood, 2000, 96, 24- [166] Felix, C. A.; Lange, B. J. Oncologist, 1999, 4, 225-240.
33. [167] Strick, R.; Strissel, P. L.; Borgers, S.; Smith, S. L.; Rowley, J. D.
[153] Negrini, M.; Felix, C. A.; Martin, C.; Lange, B. J.; Nakamura, T.; Proc. Natl. Acad. Sci. USA, 2000, 97, 4790-4795.
Canaani, E.; Croce, C. M. Cancer Res., 1993, 53, 4489-4492. [168] Ford, A. M.; Ridge, S. A.; Cabrera, M. E.; Mahmoud, H.; Steel, C.
[154] Felix, C. A.; Lange, B. J.; Hosler, M. R.; Fertala, J.; Bjornsti, M.-A. M.; Chan, L. C.; Greaves, M. Nature, 1993, 363, 358-360.
Cancer Res., 1995, 55, 4287-4292. [169] Ross, J. A. Proc. Natl. Acad. Sci. USA, 2000, 97, 4411-4413.
[155] Aplan, P. D.; Chervinsky, D. S.; Stanulla, M.; Burhans, W. C. [170] Ross, J. A.; Potter, J. D.; Reaman, G. H.; Pendergrass, T. W.;
Blood, 1996, 87, 2649-2658. Robison, L. L. Cancer Causes Control, 1996, 7, 581-590.
[171] Ross, J. A. Int. J. Cancer Suppl., 1998, 11, 26-28.

You might also like