You are on page 1of 11

Annals of Oncology 13: 1841–1851, 2002

Review DOI: 10.1093/annonc/mdf337

Irinotecan: mechanisms of tumor resistance and novel strategies


for modulating its activity
Y. Xu & M. A. Villalona-Calero*
Department of Medicine and the Experimental Therapeutics Program, Comprehensive Cancer Center, Arthur G. James Cancer Hospital and
Richard J. Solove Research Institute, Ohio State University, Columbus, OH, USA

Received 27 February 2002; revised 1 July 2002; accepted 17 July 2002

Camptothecins are broad-spectrum anticancer drugs that specifically target DNA topoisomerase I
(Topo I). The formation of a cleavable drug–Topo I–DNA complex results in lethal double-strand DNA
breakage and cell death. However, de novo or acquired clinical resistance to camptothecins is common.
Studies of the camptothecin analog irinotecan suggest the following general mechanisms of resistance:
(i) variable levels of the enzymes involved in the conversion of irinotecan; (ii) reduced cellular accu-
mulation from active drug efflux; (iii) reduced levels of Topo I expression; (iv) alterations in the
structure of Topo I from different mutations; (v) alterations in the cellular response to camptothecin–
Topo I–DNA complex formation, which involves proteasome degradation of Topo I and/or enhanced
DNA repair; and (vi) activation of the transcription factor nuclear factor kappa B by DNA damage and
subsequent suppression of apoptosis. Multiple approaches using pharmacological and biological
modulation to circumvent the above mechanisms of resistance have been incorporated into ongoing
clinical trials and are expected to enhance the antitumor activity of irinotecan and reduce its systemic
toxicity.
Key words: apoptosis, irinotecan, modulation, nuclear factor kappa B, proteasome, resistance

Introduction failure of 5-fluorouracil (5-FU) and leucovorin [2], and more


recently it has been approved for use in combination with
Irinotecan (CPT-11) is a semisynthetic analog of campto-
5-FU/leucovorin as a first-line treatment for this disease [3, 4].
thecin, originally isolated from the Chinese/Tibetan ornamen-
Studies evaluating irinotecan as adjuvant chemotherapy in
tal tree Camptotheca acuminata. It is a chemotherapy agent
node-positive colorectal cancer after resection are underway.
that causes S-phase-specific cell killing by poisoning topo-
Likewise, phase II studies on advanced esophageal and gastric
isomerase I (Topo I) in the cell. It was first discovered and
cancer also showed encouragingly high response rates [5, 6].
synthesized in Japan in 1983, and it has now demonstrated
In addition, the combination of irinotecan and cisplatin is cur-
potent antitumor activity against a wide range of tumors
[1]. Numerous studies have been done to uncover possible rently being compared with cisplatin and etoposide in patients
mechanisms for the cellular resistance to this agent, and mul- with extensive stage small-cell lung cancer in randomized
tiple experimental approaches are being tested in clinical trials trials in the United States, based on exciting results from a
for their potential ability to circumvent individual mechan- Japanese randomized trial. This combination showed marked
isms for this resistance. This paper summarizes the recent superior response rate (84.4% versus 67.5%, P = 0.02) and
advances in this field. prolonged survival (median survival 12.8 versus 9.4 months,
P = 0.002) [7]. In non-small-cell lung cancer, the combination
Clinical significance and toxicity of cisplatin and irinotecan is at least comparable with other
active combination regimens containing cisplatin [8].
Irinotecan has shown activity against colorectal, esophageal,
The major toxicities of irinotecan in clinical use are
gastric, non-small-cell and small-cell lung cancers, leukemia
myelosuppression and diarrhea. It can cause either acute
and lymphomas, as well as central nervous system malignant
diarrhea related to a cholinergic surge from inhibition of
gliomas [1]. Initial approval in the United States was as a
second-line treatment for metastatic colorectal cancer, after acetylcholinesterase, or a delayed diarrhea syndrome, which
is possibly related to the accumulation of the active metabolite
of irinotecan in the bowel [9]. On occasion, these side-effects
*Correspondence to: Dr M. A. Villalona-Calero, B406 Starling-Loving can be life-threatening, calling for pharmacological studies to
Hall, 320 West 10th Avenue, Columbus, OH 43210, USA.
Tel: +1-614-293-7511; Fax: +1-614-293-7529;
optimize dosing and schedules or for biological modulation of
E-mail: Villalona-1@medctr.osu.edu undesirable effects.

© 2002 European Society for Medical Oncology


1842

Pharmacology by uridine diphosphate glucoronosyltransferase isoform 1A1


(UGT 1A1) in the liver to an inactive form, SN-38 glucuron-
Understanding the pharmacology of irinotecan provides in-
ide (SN-38G). This isoenzyme, which is also responsible for
sights into the basis of irinotecan toxicity and tumor resistance
the glucuronidation of bilirubin, is mutated in Gilbert’s syn-
[10]. Irinotecan is a unique semisynthetic camptothecin,
drome [17] and deficient in Crigler–Najjar syndrome type I
characterized by the presence of a bulky piperidino side-chain
at the C-10 position [11]. This side-chain can be cleaved [18]. As a result, patients with these disorders are at increased
enzymatically by carboxylesterase to 7-ethyl-10-hydroxy- risk for severe irinotecan-induced toxicity. The deficiency in
camptothecin (SN-38), which is 1000-fold more potent than Gilbert’s syndrome is due to a homozygous TA insertion in
irinotecan [12]. Both irinotecan and SN-38 are in equilibrium the TATAA promoter of UGT 1A1, leading to a mutated allele
with their active lactone and inactive carboxylate forms, an [19]. Pretreatment of rats with valproic acid caused 99%
equilibrium that is pH- and protein-dependent [13, 14]. The inhibition of the formation of SN-38G by inhibiting UGT 1A1
metabolism and mechanism of action of irinotecan are sum- enzyme activity, leading to a 270% increase in the area under
marized in Figure 1. the plasma concentration–time curve (AUC) [20]. In contrast,
Carboxylesterase activity is found in serum, liver, intestine pretreatment with phenobarbital, an inducer of UGT 1A1,
and other sites [15]. Genetic variability of carboxylesterase caused a 1.7-fold increase in the AUC of SN-38G, and 31 and
expression and/or activity is suspected, as variations of SN-38 59% reduction in the AUC of SN-38 and irinotecan, respect-
levels are observed from individual to individual for a given ively [20].
dose of irinotecan [16]. After conversion from irinotecan by The elimination of irinotecan and SN-38 is through biliary
carboxylesterase, SN-38 is deactivated through conjugation excretion, which is dependent on the canalicular multispecific

Figure 1. Depicted are the steps involved in the activation and metabolism of irinotecan, as well as its mechanism of action. Numerals reflect areas in
which enhancement (+) or interference (–) with irinotecan action and toxicity have been reported: 1, phenytoin and carbamazepine (–); 2,
carboxylesterase (CE) gene transfer (+); 3, genetic profiling (+), phenobarbital and dexamethasone (–), valproic acid (+); 4, antisense canalicular
multispecific organic anion transporter (cMOAT) (+), cyclosporin (+); 5, antibiotics (–); 6, intestinal alkalinization (–); 7, multidrug resistance-
associated protein, P-glycoprotein and breast cancer resistance protein inhibitors (+); 8, topoisomerase I (Topo I) increase (+), Topo I mutation (–);
9, proteasome inhibition (+); 10, gene transfer (IκBα) (+), proteasome inhibition, thalidomide and cyclooxygenase inhibition (+). APC,
aminopentanecarboxylic acid; B Gluc, β-glucuronidase; CYP 3A4, cytochrome P-450; ds, double strand; NF-κB, nuclear factor kappa B;
NPC, 7-ethyl-10[4-(1-piperidino)-1-amino]-carbonyloxycamptothecin; SN-38, 7-ethyl-10-hydroxycamptothecin; SN-38G, SN-38 glucuronide;
ss, single strand; SUMO, small ubiquitin-like modifier; UGT, uridine diphosphate glucoronosyltransferase.
1843

organic anion transporter (cMOAT), a member of the trans- is inactivated to SN-38G by glucuronidation by UGT 1A1,
porters with an ATP cassette [21, 22]. Treatment with cyclo- overexpression of UGT at the protein and mRNA levels was
sporin A, which decreases biliary flow and inhibits cMOAT, found to account for SN-38 resistance in a human lung cancer
increases the AUC of irinotecan and SN-38 [23]. Subse- cell line [31].
quently, SN-38G is deconjugated to SN-38 by β-glucuro- Irinotecan’s complex metabolism makes this agent a target
nidase produced by the intestinal bacterial flora, which may of potential interactions with other medications. The most com-
account for the late SN-38 double peaks in the plasma, and pelling example of interference with irinotecan pharmaco-
may be responsible for the delayed intestinal toxicity of irino- kinetics comes from studies in patients with central nervous
tecan [24]. Indeed, β-glucuronidase activity is correlated with system tumors [32]. Irinotecan’s clearance in these patients
irinotecan-induced cecal damage in the rat [25]. The toxicity was approximately two-fold higher than reported in previous
in this model was attenuated by antibiotic administration, trials. The observation that the overwhelming majority of
which presumably inhibits the intestinal microflora prolifera- these patients were receiving enzyme-inducing antiepileptic
tion [25]. drugs (phenytoin, carbamazepine and phenobarbital) led to
A second major metabolite of irinotecan is aminopentane- the suspicion that induction of hepatic cytochrome P-450
carboxylic acid (APC), produced by oxidation of the terminal enzymes, including CYP3A4, resulted in increased conver-
piperidine ring by the cytochrome P-450 CYP3A4 enzyme sion of irinotecan to its APC inactive metabolite. In addition,
[26]. APC does not hydrolyze to SN-38 and is a poor inhibitor phenobarbital and dexamethasone are known to induce
of the Topo I–DNA cleavable complex. glucoronyl transferase, enhancing the conversion of SN-38 to
SN-38G [20]. Conversely, another anticonvulsant, valproic
Mechanism of action acid, inhibits SN-38 conjugation, decreasing therefore irinote-
can clearance [20]. A phase I study in patients with malignant
Irinotecan interacts with cellular Topo I–DNA complexes and glioma receiving irinotecan and concurrent enzyme-inducing
has S-phase-specific cytotoxicity [27]. Topoisomerases reduce antiepileptic drugs demonstrated that doses of irinotecan that
DNA twisting and supercoiling that occur in selected regions are two-fold higher than the usually recommended doses can
of DNA as a result of essential cellular processes such as trans- be administered safely to these patients [33].
cription, replication and repair recombination. They cleave
and reseal the phosphodiester backbone of DNA, and form a
covalent enzyme–DNA linkage, which allows the passage of Decreasing the intracellular level of irinotecan by
another single- or double-stranded DNA through the nicked active efflux
DNA. Topo I binds to single-strand DNA breaks, and the Multidrug resistance, a major obstacle encountered in cancer
reversible Topo I–irinotecan–DNA cleavable complex is not chemotherapy, is characterized by overexpression of ATP-
lethal to the cells by itself. However, upon their collisions with binding cassette (ABC) transmembrane transporters such as
the advancing replication forks, the formation of a double- P-glycoprotein (P-gp) and multidrug resistance-associated
strand DNA break occurs, leading to irreversible arrest of the protein (MRP), which actively transport chemotherapy agents
replication fork and cell death [27]. The collision of the irino- out of the cell [34]. P-gp accepts amphipathic cationic or
tecan–Topo I complex with the replication fork also results in neutral compounds as substrate, whereas MRP acts as a
G2 arrest/delay by signaling the presence of DNA damage to glutathione S-conjugate (GS-X) export pump. Several ABC
an S-phase checkpoint mechanism [28]. At higher concentra- proteins have been shown to be able to efflux camptothecins.
tions of irinotecan, non-S-phase cells can also be killed. The Using membrane vesicles from human cell lines overexpress-
mechanism of non-S-phase cell killing appears to be related to ing P-gp and MRP, it was found that both P-gp and MRP are
transcriptionally mediated DNA damage, and through the involved in the active efflux of SN-38 and irinotecan [35]. The
mechanism of apoptosis [29]. involvement of MRP was further shown in MRP-overexpress-
ing transfectants, where the ATP-dependent efflux of irino-
tecan and SN-38 was inhibited by its specific inhibitor
Mechanisms of irinotecan resistance PAK-104P [36]. Another ATP-dependent transporter, cMOAT
(also known as MRP2), which shares 49% amino acid identity
Pharmacokinetics and metabolism of irinotecan
with MRP1, also exhibits GS-X pump activity [37]. cMOAT
Irinotecan is subject to extensive metabolic conversion by is predominantly located in hepatic cells, where it is involved
various enzymatic systems in the body. The conversion of in the excretion of irinotecan to the bile. Transfection of
irinotecan to the more active form SN-38 requires carboxy- cMOAT antisense cDNA into a human hepatic cancer cell line
lesterase. A clear relationship between carboxylesterase level results in a reduction in cMOAT protein and increased sensit-
and the chemosensitivity of human small-cell and non-small- ivity to camptothecin derivatives [38].
cell lung cancer cell lines has been demonstrated in vitro, In addition, the breast cancer resistance protein (BCRP),
where irinotecan resistance was encountered in cell lines with another member of the ABC family of drug transporters, also
low carboxylesterase expression [30]. Furthermore, as SN-38 mediates resistance to camptothecins [39]. The affinities to
1844

BCRP are as follows in descending order: SN-38, topotecan, found in the area that may alter DNA cleavage or Topo I–
9-aminocamptothecin (9-AC), irinotecan. Administration of a DNA–irinotecan interactions, so that camptothecin can no
BCRP inhibitor, GF120918, reverses the resistance almost longer enter the complex [50]. Mutations immediately flank-
completely at 100 nM, further confirming the involvement of ing the catalytic tyrosine may affect the catalytic activity of
BCRP in irinotecan efflux [40]. Overexpression of BRCP has the enzyme, altering its ability to relax supercoiled DNA [51].
been described in ovarian, breast, colon and gastric cancer cell Recently, the crystal structure of Topo I–DNA covalent com-
lines [40], implying their chemoresistance. In summary, the plexes became available, which will allow structural mapping
above results indicate that the P-gp and MRP family of of these mutations and may shed light on the functional roles
transporters play important roles in the efflux and active of the mutations [52, 53].
excretion of irinotecan and can be targets for pharmacological
modulation. DNA repair: ubiqitin/26S proteasome-mediated
degradation of Topo I
Quantitative Topo I level
Another cellular mechanism of resistance to irinotecan is
As Topo I is the cellular target of irinotecan, it is conceivable repair of irinotecan-induced DNA damage; a mechanism which
that the cellular level of Topo I would be proportional to is coupled with RNA transcription [27]. According to this
irinotecan cytotoxic effects. This notion is supported by model, the collision between the elongation RNA polymerase
experimental evidence from yeast systems and mammalian complex and the Topo-I cleavable complex (on the template
cell lines [41, 42]. In irinotecan-resistant cell lines rendered strand) results in transcription arrest and the formation of a
resistant by stepwise, continuous treatment with irinotecan, Topo I linked single-strand break. This collision triggers
the total activity of Topo I was shown to be reduced compared degradation of Topo I through an ubiquitin/26S proteasome-
with the irinotecan-sensitive parental cell line [43]. Theoret- dependent system [54–56]. Subsequent to Topo I destruction,
ically, the expression of Topo I in tumor specimens may serve repair of the single-strand break can presumably occur. The
as a predictor for sensitivity to irinotecan chemotherapy. model is illustrated by studies in breast cell lines, demonstrat-
Increased levels of Topo I are demonstrated in colon cancer ing that the most sensitive cell line was completely defective
and prostate cancer, but not in renal cancer [43, 44]. Topo I in irinotecan-induced Topo I reduction, whereas the least-
levels are higher in colorectal tumors compared with normal sensitive line exhibited effective Topo I reduction. Tumor
colon mucosa, suggesting a favorable therapeutic index [43]. sensitivity to proteasome degradation can therefore serve as
Pathological specimens from breast cancer and seminoma an important parameter for determining irinotecan sensitivity/
examined for Topo I expression using immunohistochemical resistance [55]. However, studies on a panel of breast, colon
staining showed great variability, with increased expression and leukemia cell lines have demonstrated that irinotecan-
observed in only 41% of the breast and 30% of the seminoma induced Topo I reduction is defective in most tumor cell lines
specimens tested, respectively [45, 46]. Although no consist- [27], whereas Topo I reduction in peripheral blood mono-
ent association has been described between pretreatment nuclear cells has been observed in patients treated with
tumoral Topo I expression and antitumor response to irino- irinotecan [57]. It remains to be established whether irino-
tecan, this question has not been properly addressed in pro- tecan-induced Topo I reduction occurs preferentially in normal
spectively designed clinical trials. In addition, tumor cells cells to protect them from irinotecan toxicity, while tumor
may escape irinotecan cytotoxic effects by reducing the level cells remain sensitive.
of Topo I expression. Therefore, dynamic studies, evaluating In addition to the ubiquitin degradation pathway, a small
the behavior of Topo I in cancer patients during a treatment ubiquitin-like modifier (SUMO-1) can also conjugate Topo I
period are warranted. [58]. The role of SUMO–protein conjugation is still unclear,
Multiple mechanisms contribute to the regulation of Topo I and diverse functions have been proposed. One speculation is
level in the cell. In tumors demonstrating increased levels of that SUMO may target Topo I for relocation to a different
enzyme expression, increased levels of mRNA were also cellular compartment, so it cannot participate in the formation
observed [44]. This would indicate either increased trans- of Topo I cleavable complexes [58]. Likewise, a gene that
cription or increased mRNA stability. On the other hand, encodes an enzyme that can hydrolyze the bond between Topo
specimens with reduced expression were found to have a non- I–DNA has been isolated in yeast systems [59]. Enzyme-
productive rearrangement of the Topo I genome, leading to defective mutants of this gene are hypersensitive to treatments
decreased transcription and thus reduced enzymatic produc- that increase the amount of covalent complexes. Identification
tion of Topo I [47]. The rearranged allele in these specimens is of this gene in humans may have implications for the effect-
hypermethylated, resulting in transcriptional silencing. iveness of Topo I interactive agents in the clinic.

Topo I mutations Nuclear factor kappa B activation


Certain cell lines resistant to camptothecin have been found to As a result of irinotecan–Topo I–DNA complex formation and
harbor mutations in Topo I [48, 49]. Mutations have been its collision with the replication fork, a double-stranded DNA
1845

break ensues. The double-stranded DNA damage is lethal and identify tumors deficient in carboxylesterase and stimulate its
causes cell-cycle arrest and apoptosis. Recently, multiple expression. A gene-therapy-based approach of transferring the
experiments also showed that treatment of irinotecan activates carboxylesterase gene directly to the tumor followed by sys-
nuclear factor kappa B (NF-κB), an ubiquitous transcription temically administered irinotecan has been attempted in cell
factor which controls the transcription of a wild variety of lines and a nude mouse tumor model with success [71, 72].
genes involved in inflammation and immunity [60–64]. The This strategy may have potential as a local therapy for solid
activation was dependent on initial nuclear DNA damage, tumors, such as the treatment of intrabronchial tumors, lung
followed by cytoplasmic signaling events [62]. This very tumors invading into the chest wall and inoperable intra-
important discovery has raised awareness of another mechan- pulmonary tumors. Another very interesting application of
ism of chemoresistance, as many experiments on the diverse this gene therapy-based approach is to purge contaminating
roles of activated NF-κB indicated that NF-κB acts as an tumor cells from their mixture with CD34+ cells while prepar-
antiapoptotic factor, especially in early transforming tumor ing for autologous stem cell transfer. Adenovirus-mediated
cells. Activated NF-κB can suppress the apoptotic cascade transfer of carboxylesterase followed by exposure to irino-
induced by tumor necrosis factor-alpha (TNF-α), oncogenic tecan successfully killed neuroblastoma cells mixed with
Ras, and chemotherapy agents, particularly irinotecan [61–65]. CD34+ cells in vitro, while the cytotoxic effect did not affect
Inhibition of NF-κB activation augments irinotecan-induced the colony formation by CD34 cells [73]. This approach
apoptosis [63, 64]. In addition, NF-κB knockout mice die shows potential clinical utility to reduce the tumor burden of
embryonically from extensive apoptosis within the liver [66], peripheral stem cells in autologous transplantation.
indicating again that NF-κB may normally be antiapoptotic.
How does NF-κB protect cells from apoptosis? It has been Increased cellular level of Topo I
suspected that the antiapoptotic effect of NF-κB involves the
regulation and activation of pro-survival genes; however, only In preclinical systems, coadministration of irinotecan and mito-
a few of them have been identified. Major recent discoveries mycin C (MMC) had a synergistic effect [74]. It was further
have revealed that the NF-κB cell-survival pathway cross- demonstrated that MMC increases Topo I catalytic activity as
talks with the c-Jun N-terminal kinase (JNK) pathway, sub- measured by relaxation of supercoiled DNA, although Topo I
stantially blunting this pathway [67, 68]. JNKs are part of the protein level was not increased [75]. Based on these pre-
evolutionarily conserved mitogen-activated protein kinase clinical data, our group has evaluated the combination of
family, and are implicated in cell-death pathways stimulated MMC and irinotecan (given 24 h after MMC) in patients with
by environmental stresses and TNF-α. Two JNK inhibitory refractory solid malignancies [76]. Topo I gene expression
proteins have been discovered. In one model, activation of was measured in pre-MMC and post-MMC peripheral mono-
NF-κB mediates transcriptional increase of a protein called nuclear specimens by RT-PCR studies and quantified with a
gadd45b/myd118, which then lowers JNK signaling induced competitive, reaction-specific internal standard and analysis
by the TNF-α receptor [67]. The other protein (NF-κB- by capillary electrophoresis with laser-induced fluorescence
induced X-chromosome-linked inhibitor of apoptosis) is [77]. This combination chemotherapy showed promising
switched on by TNF-α in an NF-κB-dependent manner and activity in heavily pretreated gastric, esophageal, breast and
also blunts JNK activation [68]. These discoveries undoubtedly non-small-cell lung cancers. Interestingly, patients experi-
shed light on elucidating the downstream events of NF-κB encing a major response to treatment (complete and partial
activation, and its role of antiapoptosis and cell survival responses, n = 5) had a 46-fold induction of Topo I at 3 h after
following TNF-α stimulation and chemotherapy. MMC compared with baseline, and a 312-fold induction at the
end of the irinotecan infusion (P = 0.00021, ANOVA) [77]. In
contrast, non-responding patients (stable disease or tumor
Strategies to overcome resistance to progression, n = 29) did not experience significant Topo I
irinotecan induction (P = 0.64). Although these results are of interest, it
is not certain if the dynamics of Topo I following MMC in
In vivo transfer of carboxylesterase cDNA to solid peripheral mononuclear cells adequately reflect intratumoral
tumors events. Ongoing phase II tumor-specific studies with the
Irinotecan is activated by carboxylesterase to SN-38 in vivo. combination, in which tumor tissues are sampled pre- and
Although plasma and liver carboxylesterase presumably pro- post-treatment for Topo I and carboxylesterase gene expres-
duce sufficient SN-38 for cytotoxicity, a lack of such conver- sion, will help clarify this issue.
sion at the tumor site may result in decreased sensitivity to Studies on another Topo I-interactive agent, topotecan,
irinotecan. Conversely, selective increase of carboxylesterase noted the occurrence of depleted cellular levels of Topo I in
levels in tumors may produce tumor-specific activity. Carb- relation to prolonged infusion of topotecan (for 21 days). Topo
oxylesterase expression has been detected in some, but not all, I levels decreased progressively in a weekly fashion and
squamous cell lung carcinomas and adenocarcinomas at vari- returned to baseline levels 1 week after stopping the infusion
ous levels [69, 70]. Thus, conceivably it would be desirable to [78, 79]. This association indicates that prolonged infusions
1846

may have advantages because of their ability to target NF-κB from IκB results in the passage of NF-κB into the
maximum amounts of free Topo I to form DNA–drug–Topo I nucleus, where it binds to specific sequences in the promoter
complexes. This regimen was used in a phase II trial as regions of target genes [64]. As has recently been shown, the
second-line therapy in patients with previously treated ovarian NF-κB complexes negatively cross-talk to the JNK cascade,
cancers [79]. The regimen was well tolerated, and produced a reducing the TNF-α-mediated JNK activation [67, 68].
response rate of 35%, which is at the upper level for topotecan A number of strategies have been employed to inhibit the
therapy in this group of patients. Laboratory data suggested activation of NF-κB. The first approach is the molecular gene-
that the depletion of free Topo I was partly due to the forma- therapy approach of introducing a dominant negative variant
tion of cleavable complexes [78]. However, this experiment of IκBα via adenovirus vector to tumor cells [63]. This form
cannot distinguish from two other reported possibilities of free of IκBα is a ‘super-repressor’, which is resistant to phos-
Topo I depletion: (i) translocation of Topo I from the nucleus phorylation by IκB kinase, and presumably remains bound to
to cytoplasm in the presence of low doses of topotecan [80], NF-κB, thus sequestering NF-κB in the cytoplasm. This
and (ii) the transcription-triggered proteasome degradation of approach has been tested in multiple tumor cell lines and
Topo I from the Topo I–DNA complex as described above mouse tumor models (fibrosarcoma, colorectal cancer) by
[54–56]. Randomized studies are required to elucidate the different groups. It consistently demonstrated increased cell
therapeutic advantage and true mechanism in this administra- death or tumor regression in the presence of irinotecan and
tion schedule. increased induction of apoptosis [63, 87]. Inhibition of NF-κB
using the ‘super-repressor IκBα’ was found to sensitize non-
Sequential treatment with Topo I and Topo II small-cell lung cancer cells to gemcitabine-induced apoptosis
inhibitors [88]. However, using the same approach, enhanced chemo-
sensitivity was not confirmed in similar studies done by two
Cell lines selected for resistance to camptothecin often demon-
other groups [89, 90]. This discrepancy may represent the
strate collateral sensitivity to Topo II poisons and increases in
intricate cellular protein interaction and control of apoptosis.
Topo II enzyme activity or content [48]. Reciprocal changes in
NF-κB has been connected with multiple aspects of oncogen-
Topo I activity have been observed in cell lines resistant to
esis, including promotion of cell proliferation, cell migration,
Topo II poisons [81]. This phenomenon was also observed in
angiogenesis, metastasis and blockage of cell differentiation
cancer patients. Several patients demonstrated decreased
and apoptosis [64]. In many late-stage tumor cell lines and
levels of Topo I accompanied by increased levels of Topo II
other experimental tumor systems, NF-κB appears to be
when treated with the Topo I inhibitor 9-AC [82]. In colon
proapoptotic [64].
cancer cell lines, sequential exposure to camptothecin and
As the use of gene therapy to deliver NF-κB inhibitors is
etoposide leads to additive cytotoxicity [83]. Sequential
relevant to certain cancers, but is limited when considering
treatment of athymic mouse bearing human colon tumor with
widely disseminated metastases, an alternative approach is to
topotecan followed by etoposide also demonstrated increased
use pharmacological inhibitors of proteasome function. The
tumor sensitivity to etoposide [84]. However, results from a
inhibition of proteolytic function effectively blocks degrada-
recent phase I translational study of sequential administration
tion of cellular proteins that have undergone ubiquitination,
of the Topo I and II inhibitors, topotecan and etoposide, did
such as IκB [91]. MG-132, an inhibitor of proteasomes, has
not fully support the above proposed mechanism. By assaying
been shown to significantly inhibit NF-κB activation induced
Topo I and II levels in the tumor specimens after treatment, the
by irinotecan [90]. However, in that experimental system,
increase in Topo II was only clearly demonstrated in one out
apoptosis was decreased, again raising the question of the role
of six patients tested [85]. Thus, the rationale of sequential
of NF-κB to be pro- or antiapoptotic in that tumor line.
administration of Topo II and Topo I requires further invest-
PS-341, a potent boronic acid dipeptide that is highly selective
igation.
for proteasome inhibition, was also tried, in order to inhibit
proteasome function in human colon cancer cell lines [92].
Inhibition of the activation of NF-κB to enhance
Pretreatment of cancer cells with PS-341 can effectively block
irinotecan-induced apoptosis
the activation of NF-κB that is induced by exposure to SN-38/
Irinotecan exposure frequently results in increased production irinotecan. NF-κB inhibition using PS-341 also augmented
of TNF-α, interleukin-1 (IL-1), phorbol esters and lipopoly- the sensitivity of tumor cells to chemotherapy in tumor xeno-
saccharides, as well as in activation of the transcription factor graft models. The increased antitumor effect was attributable
NF-κB [86]. Activation of NF-κB leads to inhibition of apop- in part to regulation of apoptosis, with markedly increased
tosis. NF-κB is a heterodimer consisting of two proteins, p65 levels of apoptosis compared with single-agent treatment
and p50. In unstimulated cells, NF-κB is located in the alone [92]. In pancreatic xenografts, the combination of
cytoplasm and is bound to IκBα and IκBβ, which prevents it PS-341 and irinotecan results in 89% inhibition of tumor
from entering the nuclei. The external stimuli modulate signal growth and increased cellular apoptosis [93]. While using
transduction pathways leading to IκB phosphorylation, PS-341 appears to be very promising, it is important to keep in
causing its rapid degradation by proteasomes. The release of mind that the increased antitumor effect may in part be attrib-
1847

uted to the prolonged binding of irinotecan to Topo I, rendered with solid malignancies in which the combination of irino-
by the inhibition of degradation of bound Topo I–irinotecan tecan with thalidomide and subsequently the triple drug com-
complex, in the presence of proteasome inhibition [54, 55]. In bination irinotecan/thalidomide/celecoxib are evaluated. The
addition, proteasome inhibition may also stabilize a wide principal end point is demonstration of inhibition of NF-κB
range of key cell-cycle regulatory proteins, although it has nuclear translocation in peripheral mononuclear cells. Pre-
been shown to be at least independent of p53 [93]. liminary results in 10 patients receiving irinotecan 125 mg/m2
on days 1 and 8 every 3 weeks in combination with thalido-
Thalidomide and cyclooxygenase-2 inhibitors mide 400 mg daily confirms the attenuated toxicity profile
previously reported with the irinotecan/thalidomide combina-
The teratogenic sedative and antinausea drug thalidomide has
tion and no effect of thalidomide on the pharmacokinetics of
come back to the clinic based on its anti-inflammatory and
irinotecan.
anti-oncogenic properties. It is an inhibitor of TNF-α by
enhancing the degradation of TNF-α mRNA [94]. Thalido-
mide has been recently demonstrated to block NF-κB activa- Pharmacological modulations to reduce
tion through a mechanism that involves the inhibition of irinotecan toxicity
activity of the IκB kinase [95]. Thalidomide could inhibit the
activation of NF-κB induced by stimulation with TNF-α and Pharmacological modulation to reduce irinotecan toxicity has
IL-1β, and could inhibit the transcription of many endogenous attracted many investigations. Myelosuppression and diarrhea
downstream genes controlled by NF-κB [95]. This study are the two major toxicities associated with irinotecan treat-
suggests the inhibition of NF-κB as the mediator for the anti- ment. It is believed that toxicity correlates with levels of
inflammatory and anti-oncogenic properties of thalidomide; SN-38 in the plasma and bowel. Systemic SN-38 levels
in addition, it also provides a rationale for the use of thalido- depend on the expression and activity of carboxylesterases in
mide in combination with irinotecan to modulate irinotecan- serum and liver, which convert irinotecan into SN-38 [15],
induced chemoresistance. A clinical trial with the combination and on hepatic glucoronosyltransferase (UGT 1A1) activity,
of irinotecan and thalidomide in 5-FU-refractory metastatic which converts SN-38 into its glucoronidated and inactive
colorectal cancer is ongoing at the University of Arkansas. form SN-38G [18]. Although genetic variability in carboxyl-
Patients receive 350 mg/m2 irinotecan every 3 weeks and esterase expression is suspected, and wide variability on the
400 mg thalidomide daily. Preliminary results are encouraging, ability of carboxylesterases to metabolize irinotecan has been
including one complete response and three partial responses documented among its different isoenzymes [100, 101], no
among 14 evaluable patients. Interestingly, grade 3–4 diarrhea, genetic disorders of deficiency of this enzyme have been
a toxicity frequently associated with single-agent irinotecan, characterized. UGT 1A1, however, is mutated in Gilbert’s
has rarely been observed [96]. syndrome and deficient in Crigler–Najjar type 1 syndrome,
Aspirin and non-steroidal anti-inflammatory drugs resulting in increased risk for irinotecan-induced severe toxic-
(NSAIDs) inhibit cyclooxygenases (COXs) to prevent prosta- ity [17, 18]. A clinical trial has been planned at the University
glandin synthesis. They also play important roles in the of Chicago to test the predictive value of hepatic UGT 1A1
chemoprevention of colon cancer. COX-1 is constitutively genotyping for irinotecan pharmacokinetics and toxicity [19].
expressed in many types of tissue, while COX-2 expression is In addition, pharmacological modulation with cyclosporin A,
induced by cytokines and growth factors, and is increased in an inhibitor of the active irinotecan/SN-38 transporter
90% of sporadic colon cancers and 40% of colonic adenomas, cMOAT on hepatic cells [23], and phenobarbital, an inducer
but is not elevated in the normal colonic epithelium [95]. of UGT 1A1 [20] has been tested in cancer patients in a phase
Both the non-specific COX inhibitor sulindac and the COX-2 I clinical trial [102]. Cyclosporin A significantly inhibited
inhibitor celecoxib have been shown to cause regression of the clearance of irinotecan while phenobarbital induced
polyps in patients with familial adenomatous polyposis [97, glucuronidation of SN-38, both consistent with preclinical
98]. The mechanism of chemoprevention by COX-2 inhibitors findings.
may involve an increase in apoptosis [97]. In addition, SN-38G is largely eliminated by biliary excretion [103]. In
NSAIDs also act through COX-independent mechanisms by the bowel, fecal bacteria-derived β-glucuronidase converts
inhibiting the NF-κB pathway [99]. The NSAIDs specifically the non-active SN-38G to active SN-38, which can produce
bind to IκB kinase b, inhibiting its kinase activity. This inhibi- direct injury to the mucosa. Since the conversion to SN-38 is
tion results in prevention of IκB degradation and nuclear believed to be a major factor in the diarrhea produced by
translocation of NF-κB. The above studies provide evidence irinotecan, inhibition of bowel β-glucoronidase is a potential
that NSAIDS, or more specifically COX-2 inhibitors, are strategy to reduce irinotecan toxicity. In a small controlled
potential agents that may inhibit the activation of NF-κB. trial, the broad-spectrum antibiotic neomycin was given to
In order to evaluate if the target of thalidomide and COX-2 patients receiving irinotecan in order to inhibit microflora in
inhibitors mediating favorable interactions with irinotecan is the intestines. Amelioration of diarrhea was found in six out of
NF-κB, we are conducting a phase I clinical trial in patients seven patients treated [104]. However, a recent study has
1848

detected up to 30% of irinotecan excreted unchanged in the In addition, based on this knowledge, pharmacologically
bile [105]. Thus, the potential exists for direct bowel con- based or biologically based novel therapeutic approaches to
version of irinotecan into SN-38 by the recently identified modulate irinotecan activity and toxicity have been incorpor-
intestinal carboxylesterases [101]. In the absence of a direct ated into clinical trials. Although none of the novel strategies
inhibitor of intestinal carboxylesterases, strategies aimed at discussed in this manuscript has yet been validated, modula-
ameliorating irinotecan-induced diarrhea must consider the tion of NF-κB activity and Topo I expression and activity
contribution of intestinal irinotecan conversion. appears to hold the most promise for improvement in clinical
The pH dependency of the equilibrium between the active activity in early clinical trials. Randomized phase II trials
lactone forms and less active carboxylate forms of both SN-38 evaluating several novel strategies may help to identify the
and irinotecan [13] provides an additional opportunity to use most promising regimens for further study. It is the bias of the
the understanding of irinotecan pharmacology to ameliorate authors that these approaches should first be tested in irino-
toxicity. Under acidic conditions, the lactone form is favored, tecan-naive patients and later in irinotecan-refractory patients,
but at physiological or higher pH, the lactone form is unstable, in the event that encouraging activity is found on the former.
and hydrolysis to open the lactone ring into its carboxylate Given the broad spectrum of activity of irinotecan, validation
form is favored. In addition, the lactone forms are passively of these approaches in any of these settings might translate
transported in intestinal cells, resulting in an intestinal uptake into significant advances in clinical practice.
10 times greater than the carboxylate forms [106]. Therefore,
intestinal alkalinization has been proposed as a potential strat-
egy to decrease irinotecan gastrointestinal toxicity [107, 108]. Acknowledgements
In a phase II trial, oral alkalinization utilizing sodium bicarbon-
We would like to thank J. Kuhn, the University of Texas
ate, magnesium oxide, base water and ursodeoxycolic acid
Health Science Center at San Antonio, for critically reviewing
combined with ‘controlled’ defecation resulted not only in
this manuscript.
a decreased incidence of diarrhea compared with a non-
randomized control group, but also in decreased myelo-
suppression, while maintaining anticancer activity [108].
Finally, oral administration of immunomodulators, which References
induce endogenous IL-15 for protection of intestinal epithel- 1. Rothenberg ML. Irinotecan (CPT-11): recent developments and
ium, and Chinese/Japanese herbal concoctions with an unclear future directions—colorectal cancer and beyond. Oncologist 2001; 6:
mechanism of action have been used [109, 110]. A need for 66–80.
carefully designed randomized clinical trials will be needed to 2. Cunningham D, Pyrhonen S, James RD et al. Randomised trial of
irinotecan plus supportive care versus supportive care alone after
assess the efficacy (decreased toxicity without affecting anti-
fluorouracil failure for patients with metastatic colorectal cancer.
cancer activity) of all the strategies discussed above. In addi-
Lancet 1998; 352: 1413–1418.
tion, feasibility aspects need to be considered, since efficacy 3. Saltz LB, Cox JV, Blanke C et al. Irinotecan plus fluorouracil and
and simplicity will both be needed to guarantee the success of leucovorin for metastatic colorectal cancer. Irinotecan Study Group.
these approaches in clinical practice. N Engl J Med 2000; 343: 905–914.
4. Douillard JY, Cunningham D, Roth AD et al. Irinotecan combined
with fluorouracil compared with fluorouracil alone as first-line treat-
Future directions ment for metastatic colorectal cancer: a multicentre randomised trial.
Lancet 2000; 355: 1041–1047.
The intricate pathways of metabolic activation/deactivation of
5. Ilson DH, Saltz L, Enzinger P et al. Phase II trial of weekly irinotecan
irinotecan make this agent especially suitable for interactions
plus cisplatin in advanced esophageal cancer. J Clin Oncol 1999; 17:
with commonly used medications. Therefore, it is extremely 3270–3275.
important to be mindful of the potential for increased clear- 6. Ilson D, Enzinger P, Saltz L et al. Phase II trial of weekly irinotecan +
ance or increased toxicity that these medications may have cisplatin in advanced gastric cancer. Proc Am Soc Clin Oncol 1999;
when used in combination with irinotecan. However, know- 18: 259a (Abstr).
ledge of these metabolic pathways can be used to identify 7. Noda K, Nishiwaki Y, Kawahara M et al. Irinotecan plus cisplatin
modulation strategies aimed at increasing the antitumor activ- compared with etoposide plus cisplatin for extensive small-cell lung
ity and reducing the toxicity of this unique agent. Similarly, cancer. N Engl J Med 2002; 346: 85–91.
with the increasing understanding of irinotecan intracellular 8. Jagasia MH, Langer CJ, Johnson DH et al. Weekly irinotecan and
cisplatin in advanced non-small cell lung cancer: a multicenter
interactions and the intracellular mechanisms mediating
phase II study. Clin Cancer Res 2001; 7: 68–73.
resistance to this agent, the potential to prospectively identify
9. Hecht JR. Gastrointestinal toxicity of irinotecan. Oncology 1998; 12:
patients with tumors especially susceptible to irinotecan 72–78.
therapy warrants careful attention. Clinical trials of irinotecan 10. Mathijssen RH, van Alphen RJ, Verweij J et al. Clinical pharmaco-
should incorporate, whenever feasible, analyses of potential kinetics and metabolism of irinotecan (CPT-11). Clin Cancer Res
markers of clinical activity and toxicity. 2001; 7: 2182–2194.
1849

11. Kunimoto T, Nitta K, Tanaka T et al. Antitumor activity of 7-ethyl- 29. Morris EJ, Geller HM. Induction of neuronal apoptosis by campto-
10-[4-(1-piperidino)-1-piperidino]carbonyloxy-camptothecin, a novel thecin, an inhibitor of DNA topoisomerase-I: evidence for cell cycle-
water-soluble derivative of camptothecin, against murine tumors. independent toxicity. J Cell Biol 1996; 134: 757–770.
Cancer Res 1987; 47: 5944–5947. 30. van Ark-Otte J, Kedde MA, van der Vijgh WJ et al. Determinants of
12. Kawato Y, Aonuma M, Hirota Y et al. Intracellular roles of SN-38, a CPT-11 and SN-38 activities in human lung cancer cells. Br J Cancer
metabolite of the camptothecin derivative CPT-11, in the antitumor 1998; 77: 2171–2176.
effect of CPT-11. Cancer Res 1991; 51: 4187–4191. 31. Takahashi T, Fujiwara Y, Yamakido M et al. The role of glucuronida-
13. Akimota K, Kawai A, Ohya K. Kinetic studies of the hydrolysis and tion in 7-ethyl-10-hydroxycamptothecin resistance in vitro. Jpn
lactonization of camptothecin and its derivatives, CPT-11 and SN-38, J Cancer Res 1997; 88: 1211–1217.
in aqueous solution. Chem Pharm Bull 1994; 42: 2135–2138. 32. Friedman H, Petros W, Friedman et al. Irinotecan therapy in adults
14. Burke T, Munshi C, Mi Z. The important role of albumin in determin- with recurrent or progressive malignant glioma. J Clin Oncol 1999;
ing the relative human blood stabilities of camptothecin anticancer 17: 1516–1525.
drugs. J Pharm Sci 1995; 84: 518–519. 33. Prados M, Kuhn J, Yung W et al. A phase-I study of CPT-11 given
15. Kaneda N, Nagata H, Furuta T, Yokokura T. Metabolism and every 3 weeks to patients with recurrent malignant glioma. A North
pharmacokinetics of the camptothecin analogue CPT-11 in the American Brain Tumor Consortium (NABTC) study. Proc Am Soc
mouse. Cancer Res 1990; 50: 1715–1720. Clin Oncol 2001; 19: 162 (Abstr 627).
16. Ohe Y, Sasaki Y, Shinkai T et al. Phase I study and pharmacokinetics 34. Loe DW, Deeley RG, Cole SP. Biology of the multidrug resistance-
of CPT-11 with 5-day continuous infusion. J Natl Cancer Inst 1992; associated protein, MRP. Eur J Cancer 1996; 32A: 945–957.
84: 972–974. 35. Chu XY, Suzuki H, Ueda K et al. Active efflux of CPT-11 and its
17. Wasserman E, Myara A, Lokiec F et al. Severe CPT-11 toxicity in metabolites in human KB-derived cell lines. J Pharmacol Exp Ther
patients with Gilbert’s syndrome: two case reports. Ann Oncol 1997; 1999; 288: 735–741.
8: 1049–1051. 36. Chen ZS, Furukawa T, Sumizawa T et al. ATP-dependent efflux of
18. Iyer L, King CD, Whitington PF et al. Genetic predisposition to the CPT-11 and SN-38 by the multidrug resistance protein (MRP) and its
metabolism of irinotecan (CPT-11). Role of uridine diphosphate inhibition by PAK-104P. Mol Pharmacol 1999; 55: 921–928.
glucuronosyltransferase isoform 1A1 in the glucuronidation of its 37. Kool M, de Haas M, Scheffer GL et al. Analysis of expression of
active metabolite (SN-38) in human liver microsomes. J Clin Invest cMOAT (MRP2), MRP3, and MRP4, and MRP5, homologues of the
1998; 101: 847–854. multidrug resistance-associated protein gene (MRP1) in human
19. Innocenti F, Iyer L, Ratain MJ. Pharmacogenetics of anticancer cancer cell lines. Cancer Res 1997; 57: 3537–3547.
agents: lessons from amonafide and irinotecan. Drug Metab Dispos 38. Koike K, Kawabe T, Tanaka T et al. A canalicular multispecific
2001; 29: 596–600. organic anion transporter (cMOAT) antisense cDNA enhances drug
20. Gupta E, Wang X, Ramirez J, Ratain MJ. Modulation of glucuronida- sensitivity in human hepatic cancer cells. Cancer Res 1997; 57:
tion of SN-38, the active metabolite of irinotecan, by valproic acid 5475–5479.
and phenobarbital. Cancer Chemother Pharmacol 1997; 39: 440–444. 39. Schellens JH, Maliepaard M, Scheper RJ et al. Transport of topoiso-
21. Chu XY, Kato Y, Ueda K et al. Biliary excretion mechanism of merase I inhibitors by the breast cancer resistance protein. Potential
CPT-11 and its metabolites in humans: involvement of primary active clinical implications. Ann N Y Acad Sci 2000; 922: 188–194.
transporters. Cancer Res 1998; 58: 5137–5143. 40. Maliepaard M, van Gastelen MA, Tohgo A et al. Circumvention
22. Chu XY, Kato Y, Sugiyama Y. Multiplicity of biliary excretion of breast cancer resistance protein (BCRP)-mediated resistance to
mechanisms for irinotecan, CPT-11, and its metabolites in rats. camptothecins in vitro using non-substrate drugs or the BCRP
Cancer Res 1997; 57: 1934–1938. inhibitor GF120918. Clin Cancer Res 2001; 7: 935–941.
23. Gupta E, Safa AR, Wang X, Ratain MJ. Pharmacokinetic modulation 41. Reid RJ, Benedetti P, Bjornsti MA. Yeast as a model organism for
of irinotecan and metabolites by cyclosporin A. Cancer Res 1996; 56: studying the actions of DNA topoisomerase-targeted drugs. Biochim
1309–1314. Biophys Acta 1998; 1400: 289–300.
24. Gupta E, Lestingi TM, Mick R et al. Metabolic fate of irinotecan in 42. Kanzawa F, Sugimoto Y, Minato K et al. Establishment of a
humans: correlation of glucuronidation with diarrhea. Cancer Res camptothecin analogue (CPT-11)-resistant cell line of human non-
1994; 54: 3723–3725. small cell lung cancer: characterization and mechanism of resistance.
25. Takasuna K, Hagiwara T, Hirohashi M et al. Involvement of Cancer Res 1990; 50: 5919–5924.
beta-glucuronidase in intestinal microflora in the intestinal toxicity of 43. Giovanella BC, Stehlin JS, Wall ME et al. DNA topoisomerase
the antitumor camptothecin derivative irinotecan hydrochloride I-targeted chemotherapy of human colon cancer in xenografts.
(CPT-11) in rats. Cancer Res 1996; 56: 3752–3757. Science 1989; 246: 1046–1048.
26. Santos A, Zannetta S, Cresteil T et al. Metabolism of irinotecan 44. Husain I, Mohler JL, Seigler HF, Besterman JM. Elevation of topo-
(CPT-11) by CYP3A4 and CYP3A5 in humans. Clin Cancer Res isomerase I messenger RNA, protein, and catalytic activity in human
2000; 6: 2012–2020. tumors: demonstration of tumor-type specificity and implications for
27. Liu LF, Desai SD, Li TK et al. Mechanism of action of camptothecin. cancer chemotherapy. Cancer Res 1994; 54: 539–546.
Ann N Y Acad Sci 2000; 922: 1–10. 45. Lynch BJ, Bronstein IB, Holden JA. Elevations of DNA topoisomer-
28. Shao RG, Cao CX, Zhang H et al. Replication-mediated DNA ase I in invasive carcinoma of the breast. Breast J 2001; 7: 176–180.
damage by camptothecin induces phosphorylation of RPA by DNA- 46. Coleman LW, Perkins SL, Bronstein IB, Holden JA. Expression of
dependent protein kinase and dissociates RPA: DNA-PK complexes. DNA topoisomerase I and DNA topoisomerase II-alpha in testicular
EMBO J 1999; 18: 1397–1406. seminomas. Hum Pathol 2000; 31: 728–733.
1850

47. Tan KB, Mattern MR, Eng WK et al. Nonproductive rearrangement 67. De Smaele E, Zazzeroni F, Papa S et al. Induction of gadd45beta by
of DNA topoisomerase I and II genes: correlation with resistance to NF-kappaB downregulates pro-apoptotic JNK signalling. Nature
topoisomerase inhibitors. J Natl Cancer Inst 1989; 81: 1732–1735. 2001; 414: 308–313.
48. Saleem A, Edwards TK, Rasheed Z, Rubin EH. Mechanisms of 68. Tang G, Minemoto Y, Dibling B et al. Inhibition of JNK activation
resistance to camptothecins. Ann N Y Acad Sci 2000; 922: 46–55. through NF-kappaB target genes. Nature 2001; 414: 313–317.
49. Rubin EH, Li TK, Duann P, Liu LF. Cellular resistance to topoiso- 69. Takaoka K, Ohtsuka K, Jin M et al. Conversion of CPT-11 to its
merase poisons. Cancer Treat Res 1996; 87: 243–260. active form, SN-38, by carboxylesterase of non-small cell lung
50. Li XG, Haluska P Jr, Hsiang YH et al. Identification of topoisomerase cancer. Proc Am Soc Clin Oncol 1997; 16: 252a (Abstr).
I mutations affecting both DNA cleavage and interaction with 70. Kameyama M, Shoichi I, Ohtsuka K et al. Activation and detoxifica-
camptothecin. Ann N Y Acad Sci 1996; 803: 111–127. tion of CPT-11 by human lung cancer cells. Proc Am Soc Clin Oncol
51. Fujimori A, Harker WG, Kohlhagen G et al. Mutation at the catalytic 2000; 19: 505a (Abstr 1975).
site of topoisomerase I in CEM/C2, a human leukemia cell line resist- 71. Kojima A, Hackett NR, Ohwada A, Crystal RG. In vivo human
ant to camptothecin. Cancer Res 1995; 55: 1339–1346. carboxylesterase cDNA gene transfer to activate the prodrug CPT-11
52. Redinbo MR, Stewart L, Kuhn P et al. Crystal structures of human for local treatment of solid tumors. J Clin Invest 1998; 101: 1789–
topoisomerase I in covalent and noncovalent complexes with DNA. 1796.
Science 1998; 279: 1504–1513. 72. Kojima A, Hackett NR, Crystal RG. Reversal of CPT-11 resistance of
53. Stewart L, Redinbo MR, Qiu X et al. A model for the mechanism of lung cancer cells by adenovirus-mediated gene transfer of the human
human topoisomerase I. Science 1998; 279: 1534–1541. carboxylesterase cDNA. Cancer Res 1998; 58: 4368–4374.
54. Desai SD, Liu LF, Vazquez-Abad D, D’Arpa P. Ubiquitin-dependent 73. Meck MM, Wierdl M, Wagner LM et al. A virus-directed enzyme
destruction of topoisomerase I is stimulated by the antitumor drug prodrug therapy approach to purging neuroblastoma cells from
camptothecin. J Biol Chem 1997; 272: 24 159–24 164. hematopoietic cells using adenovirus encoding rabbit carboxyleste-
55. Desai SD, Li TK, Rodriguez-Bauman A et al. Ubiquitin/26S rase and CPT-11. Cancer Res 2001; 61: 5083–5089.
proteasome-mediated degradation of topoisomerase I as a resistance
74. Kano Y, Suzuki K, Akutsu M et al. Effects of CPT-11 in combination
mechanism to camptothecin in tumor cells. Cancer Res 2001; 61:
with other anti-cancer agents in culture. Int J Cancer 1992; 50:
5926–5932.
604–610.
56. Beidler DR, Chang JY, Zhou BS, Cheng YC. Camptothecin resist-
75. Gobert C, Bracco L, Rossi F et al. Modulation of DNA topoisomerase
ance involving steps subsequent to the formation of protein-linked
I activity by p53. Biochemistry 1996; 35: 5778–5786.
DNA breaks in human camptothecin-resistant KB cell lines. Cancer
76. Villalona-Calero MA, Kuhn JG, Drengler R et al. Pharmaco-
Res 1996; 56: 345–353.
logically-based phase I study of mitomycin C (MMC) as a modulator
57. Murren JR, Beidler DR, Cheng YC. Camptothecin resistance related
of irinotecan (CPT-11) antitumor activity. Proc Am Soc Clin Oncol
to drug-induced down-regulation of topoisomerase I and to steps
2001; 20: 101a (Abstr 400).
occurring after the formation of protein-linked DNA breaks. Ann N Y
77. Kolesar J, Miller J, Duan W, Villalona-Calero M. TI induction in
Acad Sci 1996; 803: 74–92.
PBMC by mitomycin-C predicts irinotecan response. Proc Am Soc
58. Mo Y, Yu Y, Shen Z, Beck W. Nuclear delocalization of human topo-
Clin Oncol 2002; 21: 115 (Abstr 456).
isomerase I in response to topotecan correlates with sumoylation of
the protein. J Biol Chem 2002; 277: 2958–2964. 78. Hochster H, Liebes L, Speyer J et al. Effect of prolonged topotecan
infusion on topoisomerase 1 levels: a phase I and pharmacodynamic
59. Pouliot JJ, Yao KC, Robertson CA, Nash HA. Yeast gene for a
Tyr-DNA phosphodiesterase that repairs topoisomerase I complexes. study. Clin Cancer Res 1997; 3: 1245–1252.
Science 1999; 286: 552–555. 79. Hochster H, Wadler S, Runowicz C et al. Activity and pharmaco-
60. Nitiss J, Wang JC. DNA topoisomerase-targeting antitumor drugs dynamics of 21-day topotecan infusion in patients with ovarian
can be studied in yeast. Proc Natl Acad Sci USA 1988; 85: 7501– cancer previously treated with platinum-based chemotherapy.
7505. New York Gynecologic Oncology Group. J Clin Oncol 1999; 17:
61. Piret B, Piette J. Topoisomerase poisons activate the transcription 2553–2561.
factor NF-kappaB in ACH-2 and CEM cells. Nucleic Acids Res 80. Danks MK, Garrett KE, Marion RC, Whipple DO. Subcellular
1996; 24: 4242–4248. redistribution of DNA topoisomerase I in anaplastic astrocytoma
62. Huang TT, Wuerzberger-Davis SM, Seufzer BJ et al. NF-kappaB cells treated with topotecan. Cancer Res 1996; 56: 1664–1673.
activation by camptothecin. A linkage between nuclear DNA damage 81. Riou JF, Grondard L, Petitgenet O et al. Altered topoisomerase I
and cytoplasmic signaling events. J Biol Chem 2000; 275: 9501– activity and recombination activating gene expression in a human
9509. leukemia cell line resistant to doxorubicin. Biochem Pharmacol
63. Wang CY, Cusack JC Jr, Liu R, Baldwin AS Jr. Control of inducible 1993; 46: 851–861.
chemoresistance: enhanced anti-tumor therapy through increased 82. Rubin E, Wood V, Bharti A et al. A phase I and pharmacokinetic
apoptosis by inhibition of NF-kappaB. Nat Med 1999; 5: 412–417. study of a new camptothecin derivative, 9-aminocamptothecin. Clin
64. Baldwin AS. Control of oncogenesis and cancer therapy resistance Cancer Res 1995; 1: 269–276.
by the transcription factor NF-kappaB. J Clin Invest 2001; 107: 83. Bertrand R, O’Connor PM, Kerrigan D, Pommier Y. Sequential
241–246. administration of camptothecin and etoposide circumvents the
65. Mayo MW, Wang CY, Cogswell PC et al. Requirement of antagonistic cytotoxicity of simultaneous drug administration in
NF-kappaB activation to suppress p53-independent apoptosis slowly growing human colon carcinoma HT-29 cells. Eur J Cancer
induced by oncogenic Ras. Science 1997; 278: 1812–1815. 1992; 28A: 743–748.
66. Beg AA, Baltimore D. An essential role for NF-kappaB in preventing 84. Whitacre CM, Zborowska E, Gordon NH et al. Topotecan increases
TNF-alpha-induced cell death. Science 1996; 274: 782–784. topoisomerase IIalpha levels and sensitivity to treatment with
1851

etoposide in schedule-dependent process. Cancer Res 1997; 57: 98. Steinbach G, Lynch PM, Phillips RK et al. The effect of celecoxib,
1425–1428. a cyclooxygenase-2 inhibitor, in familial adenomatous polyposis.
85. Hammond LA, Eckardt JR, Ganapathi R et al. A phase I and trans- N Engl J Med 2000; 342: 1946–1952.
lational study of sequential administration of the topoisomerase I and 99. Yamamoto Y, Yin MJ, Lin KM, Gaynor RB. Sulindac inhibits
II inhibitors topotecan and etoposide. Clin Cancer Res 1998; 4: 1459– activation of the NF-kappaB pathway. J Biol Chem 1999; 274: 27
1467. 307–27 314.
86. Baldwin AS Jr. The NF-kappa B and I kappa B proteins: new dis- 100. Humerickhouse R, Lorhrbach K, Li L et al. Characterization of
coveries and insights. Annu Rev Immunol 1996; 14: 649–683. CPT-11 hydrolysis by human liver carboxylesterase isoforms hCE-1
87. Cusack JC Jr, Liu R, Baldwin AS Jr. Inducible chemoresistance to and hCE-2. Cancer Res 2000; 60: 1189–1192.
7-ethyl-10-[4-(1-piperidino)-1-piperidino]-carbonyloxycamptothecin 101. Kanna R, Morton C, Danks M, Potter P. Proficient metabolism of
(CPT-11) in colorectal cancer cells and a xenograft model is over- irinotecan by a human intestinal carboxylesterase. Cancer Res 2000;
come by inhibition of nuclear factor-kappaB activation. Cancer Res 60: 4725–4728.
2000; 60: 2323–2330. 102.Ratain MJ, Goh BC, Lalitha I et al. A phase I study of irinotecan
88. Jones DR, Broad RM, Madrid LV et al. Inhibition of NF-kappaB (CPT-11) with pharmacokinetic modulation by cyclosporine A
sensitizes non-small cell lung cancer cells to chemotherapy-induced (CSA) and phenobarbital (PB). Proc Am Soc Clin Oncol 1999; 18:
apoptosis. Ann Thorac Surg 2000; 70: 930–936. 202a (Abstr 777).
89. Bentires-Alj M, Hellin AC, Ameyar M et al. Stable inhibition of 103. Atsumi R, Suzuki W, Hakusui H. Identification of the metabolites of
nuclear factor kappaB in cancer cells does not increase sensitivity to irinotecan, a new derivative of campothecin, in rat bile and its biliary
cytotoxic drugs. Cancer Res 1999; 59: 811–815. excretion. Xenobiotica 1991; 21: 1159–1169.
90. Tabata M, Tabata R, Grabowski DR et al. Roles of NF-kappaB and 26 104. Kehrer DF, Sparreboom A, Verweij J et al. Modulation of irino-
S proteasome in apoptotic cell death induced by topoisomerase I and tecan-induced diarrhea by cotreatment with neomycin in cancer
II poisons in human nonsmall cell lung carcinoma. J Biol Chem 2001; patients. Clin Cancer Res 2001; 7: 1136–1141.
276: 8029–8036. 105. Slatter J, Schaaf L, Sams J et al. Pharmacokinetics, metabolism,
91. Palombella VJ, Rando OJ, Goldberg AL, Maniatis T. The ubiquitin- and excretion of irinotecan (CPT-11) following i.v. infusion of
proteasome pathway is required for processing the NF-kappa B1 [14C]CPT-11 in cancer patients. Drug Metab Dispos 2000; 28:
precursor protein and the activation of NF-kappa B. Cell 1994; 78: 423–433.
773–785. 106. Kobayashi K, Bouscarel B, Matsuzaki Y et al. pH-dependent uptake
92. Cusack JC Jr, Liu R, Houston M et al. Enhanced chemosensitivity to of irinotecan and its active metabolite, SN-38, by intestinal cells. Int
CPT-11 with proteasome inhibitor PS-341: implications for systemic J Cancer 1999; 83: 491–496.
nuclear factor-kappaB inhibition. Cancer Res 2001; 61: 3535–3540. 107. Ikegami T, Ha L, Arimori K et al. Intestinal alkalization as a possible
93. Shah SA, Potter MW, McDade TP et al. 26S proteasome inhibition preventive mechanism in irinotecan (CPT-11) diarrhea. Cancer Res
induces apoptosis and limits growth of human pancreatic cancer. 2002; 62: 179–187.
J Cell Biochem 2001; 82: 110–122. 108. Takeda Y, Kobayashi K, Akiyama Y et al. Prevention of irinotecan
94. Moreira AL, Sampaio EP, Zmuidzinas A et al. Thalidomide exerts its (CPT-11)-induced diarrhea by oral alkalization combined with con-
inhibitory action on tumor necrosis factor alpha by enhancing mRNA trol of defecation in cancer patients. Int J Cancer 2001; 92: 269–275.
degradation. J Exp Med 1993; 177: 1675–1680. 109. Shinohara H, Killion J, Bucana C et al. Oral administration of the
95. Keifer JA, Guttridge DC, Ashburner BP, Baldwin AS Jr. Inhibition of immunomodulator JBT-3002 induces endogenous interleukin 15 in
NF-kappa B activity by thalidomide through suppression of IkappaB intestinal macrophages for protection against irinotecan-mediated
kinase activity. J Biol Chem 2001; 276: 22 382–22 387. destruction of intestinal epithelium. Clin Cancer Res 1999; 5: 2148–
96. Govindarajan R. Irinotecan/thalidomide in metastatic colorectal 2156.
cancer. Oncology 2002; 16: 23–26. 110. Mori K, Hirose T, Machida S, Tominaga K. Kampo medicines for
97. Janne PA, Mayer RJ. Chemoprevention of colorectal cancer. N Engl the prevention of irinotecan-induced diarrhea in advanced non-small
J Med 2000; 342: 1960–1968. cell lung cancer. Gan To Kagaku Ryoho 1998; 25: 1159–1163.

You might also like