+ All Categories
Home > Documents > Phase I and Pharmacological Study of the Pulmonary ... · rationale for the clinical evaluation of...

Phase I and Pharmacological Study of the Pulmonary ... · rationale for the clinical evaluation of...

Date post: 25-Aug-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
9
[CANCER RESEARCH 5.1. 1794-1801. April 15. 1993) Phase I and Pharmacological Study of the Pulmonary Cytotoxin 4-Ipomeanol on a Single Dose Schedule in Lung Cancer Patients: Hepatotoxicity Is Dose Limiting in Humans1 Eric K. Rowinsky,2 Dennis A. Noe, David S. Ettinger, Michaele C. Christian, Barbara G. Lubejko, Elliot K. Fishman, Susan E. Sartorius, Michael R. Boyd, and Ross C. Donehower Division of Pharmacology and Experimental Therapeutics, The Johns Hopkins Oncolog\ Center /£. K. K.. D. A. M, B. G. L., R. C. D.J and the Department of Radiologi' ¡E.K. F.I. The Johns Hopkins Medical Institutions. Baltimore. Maryland 21287; and the Cancer Therapy and Evaluation Program [M. C. C.] and The Laboratory of Drug Discovery Research and Development. Developmental Therapeutics Program ¡M.R. B. /. Division of Cancer Treatment. National Cancer Institute. Bethesda, Maryland 20892 ABSTRACT 4-Ipomeanol (IPO), a naturally occurring pulmonary toxin, is the first cytotoxic agent to undergo clinical development based on a biochemical- biological rationale as an antineoplastic agent targeted specifically against lung cancer. This rationale is based on preclinical observations that met abolic activation and intracellular binding of IPO, as well as cytotoxicity, occurred selectively in tissues and cancers derived from tissues that are rich in specific P450 mixed function oxidase enzymes. Although tissues capable of activating IPO to cytotoxic intermediates in vitro include liver, lung, and kidney, IPO has been demonstrated in rodents and dogs to undergo in situ activation, bind covalently, and induce cytotoxicity pref erentially in lung tissue at doses not similarly affecting liver or kidneys. Although the drug was devoid of antitumor activity in the conventional murine preclinical screening models, cytotoxic activity was observed in human lung cancers in vitro and in human lung cancer xenografts in vivo, adding to the rationale for clinical development. Somewhat unexpectantly, hepatocellular toxicity was the dose-limiting principal toxicity of IPO administered as a 30-min infusion every 3 weeks to patients with lung cancer. In this study, 55 patients received 254 courses at doses almost spanning 3 orders of magnitude, 6.5 to 1612 mg/m2. Transient and isolated elevations in hepatocellular enzymes, predomi nately alanine aminotransferase, occurred in the majority of courses of IPO at 1032 mg/m2, which is the recommended IPO dose for subsequent phase II trials. At higher doses, hepatocellular toxicity was more severe and was often associated with right upper quadrant pain and severe malaise. Toxic effects were also noted in other tissues capable of activating IPO, including possible nephrotoxicity in a patient treated with one course of IPO at 154 mg/m2 and severe, reversible pulmonary toxicity in another patient who received nine courses of IPO at doses ranging from 202 to 826 mg/m2. Although individual plasma drug disposition curves were well described by a two-compartment first order elimination model, the rela tionship between IPO dose and area under the disposition curve was curvilinear, suggesting saturable elimination kinetics. At the maximum tolerated dose, the mean half-lives (Al and A2) were 6.7 and 114.5 min, respectively. Renal excretion of parent compound accounted for less than 2% of the administered dose of IPO. An unidentified metabolite was detected in the plasma of patients treated at higher doses. No objective antitumor responses were observed; however, stable disease persisted for at least eight courses in 27% of patients. The preponderance of clinical toxicity observed in liver rather than lung suggests that IPO may be preferentially activated and bound in liver rather than lung or other tissues in humans or that human lung tissue is more effective at detoxifying and/or is more tolerant to activated IPO than other species. In any event, these observations suggest further that the rationale for the clinical evaluation of IPO should be extended to include liver cancers and possibly renal cancers, as well as lung cancers. Received 11/4/92: accepted 3/8/93. The costs of publication of this article were defrayed in part by the payment of page charges. This article must therefore be hereby marked advertisement in accordance with 18 U.S.C. Section 1734 solely to indicate this fact. . 1This study was supported by NIH Contract NO1-CM-57738. Presented in part at the annual meetings of the American Society of Clinical Oncology. San Francisco, CA, May 1989. and San Diego. CA. May 1992. 2 To whom requests for reprints should be addressed, at The Johns Hopkins Oncology Center. Division of Pharmacology and Experimental Therapeutics, 1-121, 600 North Wolfe Street. Baltimore. MD 21287-8934. INTRODUCTION IPO3 (Fig. 1) is a naturally occurring furan isolated from common sweet potatoes (Ipomoea batatas) infected with the fungus Fusarium sotanÃ-(1-3). It is the first agent to be developed under the auspices of the NCI based on a biochemical-biological rationale (4) as an anti neoplastic agent targeted specifically against lung cancer (reviewed in Ref. 5). IPO was isolated and identified by Boyd et al. (I, 2) as the agent responsible for outbreaks of a lethal pulmonary disease in cattle. IPO was demonstrated to be a tissue-specific cytotoxin which requires metabolic activation by cytochrome P450 mixed function oxidases to a highly reactive intermediate that covalently binds to macromole- cules (4-24). In lower mammals, such as rabbits, rats, guinea pigs, female mice, and dogs, IPO is preferentially activated in the lung, specifically in bronchiolar Clara cells, and to a lesser extent in type II pneumocytes, which are rich in the specific cytochrome P450 isoen- zymes required for IPO activation (4). The predominant toxicity in these species is pulmonary (4-27). IPO binds immediately to nucleo- philic macromolecules at the site of activation (4), leading to bron chiolar epithelial necrosis preferentially involving the Clara cells (4, 25, 26). Maximal expression of lung toxicity occurs between days 1 and 5 after treatment (12, 25, 26). In dogs treated with lethal doses, radiographie findings are consistent with a interstitial pneumonitis including severe acute inflammation with neutrophil and monocytic infiltration of alveoli and bronchioles, a proteinaceous alveolar exú date, capillary congestion, and alveolar cell necrosis (5, 25-27). In lower mammals the slope of the dose-toxicity curves are steep, with small differences between doses associated with reversible pulmonary toxicity and death ( 12, 25, 26). The relative tissue content and substrate specificities of various isoforms of cytochrome P450 may vary considerably among different species, different organs within a given species, and among different types of cells within a given organ. Such differences may explain, at least in part, certain species differences in the relative tissue specific ities of IPO. For example, histopathological evidence of toxicity has been documented not only in the lungs but also in other tissues, such as the liver (hamster) and the kidneys (male mice), which also possess cytochrome P450 isoenzymes capable of activating IPO to cytotoxic intermediates (14, 22). However, the only species in which pulmonary toxicity is entirely absent and in which hepatotoxicity predominates is in birds. Interestingly, avian lungs are devoid of P450 enzymes for metabolic activation of IPO and they lack terminal airways and both the Clara and type II alveolar cells that typically line these structures in other species (23, 24). ' The abbreviations used are: IPO. 4-ipomeanol; NCI, National Cancer Institute: IC50, 50% inhibitory concentration; FEV,, forced expiratory volume over I s; POi, arterial partial pressure of oxygen; DLCO, pulmonary diffusion capacity for carbon monoxide; DLCO/ALV. DLCO adjusted for alveolar volume; MTD, maximum tolerated dose; FVC. forced vital capacity; LCT. limited computerized tomography; AUC, area(s) under the curve; CT, computerized tomography; ALT. alanine aminotransferase; AST, aspanate aminotransferase. 1794 Research. on December 14, 2020. © 1993 American Association for Cancer cancerres.aacrjournals.org Downloaded from
Transcript
Page 1: Phase I and Pharmacological Study of the Pulmonary ... · rationale for the clinical evaluation of IPO should be extended to include liver cancers and possibly renal cancers, as well

[CANCER RESEARCH 5.1. 1794-1801. April 15. 1993)

Phase I and Pharmacological Study of the Pulmonary Cytotoxin 4-Ipomeanol on a

Single Dose Schedule in Lung Cancer Patients: Hepatotoxicity Is DoseLimiting in Humans1

Eric K. Rowinsky,2 Dennis A. Noe, David S. Ettinger, Michaele C. Christian, Barbara G. Lubejko, Elliot K. Fishman,

Susan E. Sartorius, Michael R. Boyd, and Ross C. DonehowerDivision of Pharmacology and Experimental Therapeutics, The Johns Hopkins Oncolog\ Center /£. K. K.. D. A. M, B. G. L., R. C. D.J and the Department of Radiologi' ¡E.K.F.I. The Johns Hopkins Medical Institutions. Baltimore. Maryland 21287; and the Cancer Therapy and Evaluation Program [M. C. C.] and The Laboratory of Drug DiscoveryResearch and Development. Developmental Therapeutics Program ¡M.R. B. /. Division of Cancer Treatment. National Cancer Institute. Bethesda, Maryland 20892

ABSTRACT

4-Ipomeanol (IPO), a naturally occurring pulmonary toxin, is the firstcytotoxic agent to undergo clinical development based on a biochemical-

biological rationale as an antineoplastic agent targeted specifically againstlung cancer. This rationale is based on preclinical observations that metabolic activation and intracellular binding of IPO, as well as cytotoxicity,occurred selectively in tissues and cancers derived from tissues that arerich in specific P450 mixed function oxidase enzymes. Although tissuescapable of activating IPO to cytotoxic intermediates in vitro include liver,lung, and kidney, IPO has been demonstrated in rodents and dogs toundergo in situ activation, bind covalently, and induce cytotoxicity preferentially in lung tissue at doses not similarly affecting liver or kidneys.Although the drug was devoid of antitumor activity in the conventionalmurine preclinical screening models, cytotoxic activity was observed inhuman lung cancers in vitro and in human lung cancer xenografts in vivo,adding to the rationale for clinical development.

Somewhat unexpectantly, hepatocellular toxicity was the dose-limitingprincipal toxicity of IPO administered as a 30-min infusion every 3 weeks

to patients with lung cancer. In this study, 55 patients received 254 coursesat doses almost spanning 3 orders of magnitude, 6.5 to 1612 mg/m2.

Transient and isolated elevations in hepatocellular enzymes, predominately alanine aminotransferase, occurred in the majority of courses ofIPO at 1032 mg/m2, which is the recommended IPO dose for subsequent

phase II trials. At higher doses, hepatocellular toxicity was more severeand was often associated with right upper quadrant pain and severemalaise. Toxic effects were also noted in other tissues capable of activatingIPO, including possible nephrotoxicity in a patient treated with one courseof IPO at 154 mg/m2 and severe, reversible pulmonary toxicity in another

patient who received nine courses of IPO at doses ranging from 202 to 826mg/m2. Although individual plasma drug disposition curves were well

described by a two-compartment first order elimination model, the rela

tionship between IPO dose and area under the disposition curve wascurvilinear, suggesting saturable elimination kinetics. At the maximumtolerated dose, the mean half-lives (Al and A2) were 6.7 and 114.5 min,

respectively. Renal excretion of parent compound accounted for less than2% of the administered dose of IPO. An unidentified metabolite wasdetected in the plasma of patients treated at higher doses. No objectiveantitumor responses were observed; however, stable disease persisted forat least eight courses in 27% of patients.

The preponderance of clinical toxicity observed in liver rather thanlung suggests that IPO may be preferentially activated and bound in liverrather than lung or other tissues in humans or that human lung tissue ismore effective at detoxifying and/or is more tolerant to activated IPO thanother species. In any event, these observations suggest further that therationale for the clinical evaluation of IPO should be extended to includeliver cancers and possibly renal cancers, as well as lung cancers.

Received 11/4/92: accepted 3/8/93.The costs of publication of this article were defrayed in part by the payment of page

charges. This article must therefore be hereby marked advertisement in accordance with18 U.S.C. Section 1734 solely to indicate this fact. .

1This study was supported by NIH Contract NO1-CM-57738. Presented in part at the

annual meetings of the American Society of Clinical Oncology. San Francisco, CA, May1989. and San Diego. CA. May 1992.

2 To whom requests for reprints should be addressed, at The Johns Hopkins OncologyCenter. Division of Pharmacology and Experimental Therapeutics, 1-121, 600 NorthWolfe Street. Baltimore. MD 21287-8934.

INTRODUCTION

IPO3 (Fig. 1) is a naturally occurring furan isolated from common

sweet potatoes (Ipomoea batatas) infected with the fungus Fusariumsotaní(1-3). It is the first agent to be developed under the auspices ofthe NCI based on a biochemical-biological rationale (4) as an anti

neoplastic agent targeted specifically against lung cancer (reviewed inRef. 5).

IPO was isolated and identified by Boyd et al. (I, 2) as the agentresponsible for outbreaks of a lethal pulmonary disease in cattle. IPOwas demonstrated to be a tissue-specific cytotoxin which requires

metabolic activation by cytochrome P450 mixed function oxidases toa highly reactive intermediate that covalently binds to macromole-cules (4-24). In lower mammals, such as rabbits, rats, guinea pigs,

female mice, and dogs, IPO is preferentially activated in the lung,specifically in bronchiolar Clara cells, and to a lesser extent in type IIpneumocytes, which are rich in the specific cytochrome P450 isoen-

zymes required for IPO activation (4). The predominant toxicity inthese species is pulmonary (4-27). IPO binds immediately to nucleo-

philic macromolecules at the site of activation (4), leading to bronchiolar epithelial necrosis preferentially involving the Clara cells (4,25, 26). Maximal expression of lung toxicity occurs between days 1and 5 after treatment (12, 25, 26). In dogs treated with lethal doses,radiographie findings are consistent with a interstitial pneumonitisincluding severe acute inflammation with neutrophil and monocyticinfiltration of alveoli and bronchioles, a proteinaceous alveolar exúdate, capillary congestion, and alveolar cell necrosis (5, 25-27). Inlower mammals the slope of the dose-toxicity curves are steep, with

small differences between doses associated with reversible pulmonarytoxicity and death ( 12, 25, 26).

The relative tissue content and substrate specificities of variousisoforms of cytochrome P450 may vary considerably among differentspecies, different organs within a given species, and among differenttypes of cells within a given organ. Such differences may explain, atleast in part, certain species differences in the relative tissue specificities of IPO. For example, histopathological evidence of toxicity hasbeen documented not only in the lungs but also in other tissues, suchas the liver (hamster) and the kidneys (male mice), which also possesscytochrome P450 isoenzymes capable of activating IPO to cytotoxicintermediates (14, 22). However, the only species in which pulmonarytoxicity is entirely absent and in which hepatotoxicity predominates isin birds. Interestingly, avian lungs are devoid of P450 enzymes formetabolic activation of IPO and they lack terminal airways and boththe Clara and type II alveolar cells that typically line these structuresin other species (23, 24).

' The abbreviations used are: IPO. 4-ipomeanol; NCI, National Cancer Institute: IC50,

50% inhibitory concentration; FEV,, forced expiratory volume over I s; POi, arterialpartial pressure of oxygen; DLCO, pulmonary diffusion capacity for carbon monoxide;DLCO/ALV. DLCO adjusted for alveolar volume; MTD, maximum tolerated dose; FVC.forced vital capacity; LCT. limited computerized tomography; AUC, area(s) under thecurve; CT, computerized tomography; ALT. alanine aminotransferase; AST, aspanateaminotransferase.

1794

Research. on December 14, 2020. © 1993 American Association for Cancercancerres.aacrjournals.org Downloaded from

Page 2: Phase I and Pharmacological Study of the Pulmonary ... · rationale for the clinical evaluation of IPO should be extended to include liver cancers and possibly renal cancers, as well

PHASE l AND PHARMACOLOGICAL STUDY OF 4-IPOMEANOL

OH

Fig. 1. Structure of 4-ipomeanol.

IPO was not active against the in vivo tumor murine models (e.g.,L12IO and P388 leukemias) nor in the human tumor xenografts usedin the NCI screen in the early 1980s (5, 27), and continuous treatmentwith high concentrations were required to inhibit colony formation ofmany types of human tumors in colony-forming assays (27). Severalhuman tumor types, including many non-small cell lung lines and theMCF-7 breast cancer line, as well as its doxorubicin-resistant variantMCF-7/ADR that expresses the multidrug-resistant phenotype. were

relatively sensitive to IPO in an automated cell culture growth inhibition assay (27); however. IC50 values were seldom less than 5mmol/liter, a concentration that is not likely to be achieved andsustained in vivo (5, 27). Since many of the tumors used in thetraditional screens were not likely to contain the metabolic apparatusnecessary to activate IPO, the rationale to develop IPO was furthersupported by some observations of activity in novel screening systems(27-30). For example. IPO was active against several other humannon-small cell lung cancers growing in vitro and in xenografts (27-

30). ¡nvitro, drug sensitivity correlated with the ability of these cellsto activate IPO, as well as their morphological resemblance to Claraand type II alveolar cells (28). In addition, IPO was demonstrated toinhibit the growth of non-small cell lung cancers implanted intrabron-

chially into athymic mice using a novel orthotopic xenograft modelpermitting cytotoxic testing in a milieu resembling the clinical setting(29, 30). Several lung cancer cell lines morphologically resemblingsmall cell cancer are also capable of activating and covalently bindingIPO (28). In addition, a diverse sampling of biopsied primary humanlung carcinomas, as well as adjacent lung tissue from fresh surgicalspecimens, have been demonstrated to be capable of activating andbinding IPO (31).

The decision to develop IPO as an anticancer drug was based on itsunique mechanism of action, as well as its potential tissue specificity.These characteristics, as well as the steep dose-toxicity relationship of

IPO, indicated that clinical trials should be performed cautiously,using such measures as stringent eligibility criteria, rigorous pulmonary monitoring, and a conservative dose escalation scheme. Thepurposes of this study were to: (a) determine the MTD of IPO givenby a single, brief i.v. infusion repeated every 3 weeks; (b) recommenda dose for phase II trials; (c) characterize the toxicities associated withthis schedule of administration; (d) seek preliminary evidence forantitumor activity: and (e) describe the pharmacology of IPO anddetermine whether it could be related to relevant clinical endpoints.

MATERIALS AND METHODS

Eligibility. Based on the lack of significant preclinical activity in extrapul-

monary neoplasms, only patients with histologically documented advancednon-small and small cell lung cancers were candidates for this study. Generaleligibility criteria included ages >18 years; no major surgery within 14 days or

wide field radiotherapy and/or chemotherapy within 28 days (or 6 weeks forthose treated with a nitrosourea or mitomycin); adequate hematological (WBCcount > 4000//J/ and platelet count > 100,000/nD, hepatic (total bilirubin <1.2mg/dl. and renal (creatinine <l.5 mg/dl) functions; and no other coexisting

medical problems of sufficient severity to prevent full compliance with the

study. Due to the potential pulmonary toxicity of IPO. eligibility was alsorestricted to patients with adequate pulmonary function as defined by: (a) aFEV, >1.5 liter; (b) a PO2 >70 mm Hg; and (c) a DLCO/ALV >75% of the

predicted value. All patients gave informed written consent.Dosage and Drug Administration. The starting dose of IPO. 6.5 mg/nr.

was equivalent to l()'/r of the dose that was lethal in 10% of female mice. It was

administered as a 30-min infusion every 3 weeks. Initially, doses were esca

lated using a modified Fibonacci search method to a dose of 104 mg/nr. Thiswas followed by 40% increments to 154, 216, 302, 422. 590. and 826 mg/m:.

Thereafter, a more conservative dosing scheme, in which doses were escalatedin increments of 25%. was used due to the occurrence of pulmonary toxicityin one patient treated at the 826 mg/nr dose level. At least three IPO-naive

patients were treated at each escalated dose level. As the MTD was approachedand potential dose-limiting toxicity was observed, at least six new patients

were treated. The MTD was defined as one dose level below the dose thatinduced greater than NCI grade 3 toxicity in more than one-third of IPO-naive

patients. Initially, dose escalation was not permitted in the same patient. However, the protocol was subsequently amended due to the lack of toxicity inpatients treated up to the 46 mg/m2 dose level. Intrapatient dose escalation was

then permitted if the patient received at least two courses at the lower doselevel without toxicity and if two new patients had already been treated at thenext highest dose. Dose escalation in the same patient proceeded until pulmonary toxicity was observed at 826 mg/m2. Thereafter, intrapatient dose esca

lation was not permitted. The protocol permitted dose reductions by one to twodose levels for patients experiencing the following toxicities during a previouscourse: (<;) NCI grade 2 pulmonary toxicity defined as moderate pulmonarysymptoms; or (b) a decrease in FEV,. FVC, PO,, or DLCO/ALV of 25 to 50%.Dose modifications were not performed for other toxic effects.

IPO was supplied by the Division of Cancer Treatment. NCI (Bethesda.MD), in 2-ml vials containing a mixture of 10 mg IPO/ml in 0.9% sodium

chloride injection, with sodium hydroxide added to adjust the pH to 6. IPO wasinitially reconstituted with 100 ml of 0.9% sodium chloride solution and thenadministered over 30 min. but IPO was administered undiluted at higher doses.The first dose was administered on the inpatient units of The Johns HopkinsOncology Center, and all subsequent treatment was given in the clinic. Patientswere treated at approximately 10 a.m. to avoid potential variability due toOrcadian fluctuations in microsomal P450 enzymes and tissue glutathionewhich are involved in IPO metabolic activation and detoxification, respectively. Medications, including steroids and H2 antagonists, barbituates, andphenytoin that are known to modulate the activity of microsomal enzymes,were avoided.

Pretreatment and Follow-up Studies. Histories, physical examinations,

and routine laboratory studies were performed prior to treatment and weeklyduring therapy. Routine laboratory studies included complete blood cell anddifferential WBC counts, electrolytes, urea, creatinine. glucose, total protein,albumin, calcium, phosphate, uric acid, alkaline phosphatase, total and directbilirubin, AST (serum glutamic oxaloacetic transaminase), ALT (serumglutamic pyruvic transaminase). prothrombin time, and urinalysis. Liver function tests were also performed on days 2 and 4 after acute enzyme elevationswere initially observed. In addition, pulmonary testing was performed at regular intervals including: («)pulmonary function testing (spirometry. FVC,FEV,. lung volumes, vital capacity, alveolar volume, and single breath DLCO)

prior to treatment and on days 2, 4. 8, 15. and 22 during course 1 and prior totreatment and weekly during subsequent courses; (hi arterial blood gases (pH,pO2. and pCOi) prior to treatment and on days I, 2. 3. 4, 8, 15. and 22 duringcourse 1 and weekly during subsequent courses; (<•)chest radiographs prior to

treatment and on days 1. 2. 3. 4, 8, 15. and 22 during course I and weeklyduring subsequent courses: and (d) LCT of the lungs prior to treatment and onday 2. 4. 8, 15. and 22 during course 1 and weekly during subsequent courses.LCT was obtained from a level just below the lung apex through the diaphramat 15-mm intervals on either a Siemens Somatom Plus or DRH scanners

(Siemens Medical Systems, Iselin, NJ); scan parameters were 1 s, 125 kVp.250 mA. and 4-mm collimation or 4 s, 125 kVp, 250 mA. and 4-mm colli-

mation. respectively. All scans were reconstructed with a high spatial frequency reconstruction algorithm to optimize definition of the lung parenchyma. The lung parenchymal density in Hounsfield units was recorded at eachlevel. Formal tumor measurements were performed after every two courses,and patients were able to continue treatment if they did not develop progressivedisease. A complete response was defined as complete disappearance of all

1795

Research. on December 14, 2020. © 1993 American Association for Cancercancerres.aacrjournals.org Downloaded from

Page 3: Phase I and Pharmacological Study of the Pulmonary ... · rationale for the clinical evaluation of IPO should be extended to include liver cancers and possibly renal cancers, as well

PHASE [ AND PHARMACOLOGICAL STUDY OF 4-IPOMEANOL

active disease and a partial response required a 50% or greater reduction in thesum of the product of the bidimensional measurements of all measurablelesions.

Pharmacological Analyses. Pharmacological studies were performed during the initial course of therapy in all patients. Blood samples in heparanizedtubes were collected before the infusion, at 10 and 20 min during the infusion,

and at the end of infusion. Samples were also collected at 1,2, 5, 10, 20, and30 min and 1. 2, 4, 6, 10, 24, and 48 h after the end of infusion. Urine wascollected continuously for 48 h following drug administration. Plasma wasseparated by centrifugation and was stored at -20°C until analysis was per

formed. IPO is stable for at least 6 months under these conditions (data notshown).

IPO concentrations in plasma and urine were measured by gas chromatog-raphy using 1-undecanol as an internal standard. To each 1 ml plasma sample,3 ml of benzene were added, along with 40 ul 1-undecanol 10 mmol/liter in

methanol. After being vortexed for 30 s and centrifuged for 10 min, 2 ml of thebenzene supernatant were removed and transferred to clean tubes. Next, 25 (jlof heptfluorobutylimidazole were added to the supernatant, and the mixturewas heated in a heating block at 50°Cfor 15 min. After 0.5 ml of water was

added to neutralize the derivatizing agent, the mixture was mechanicallyshaken for 5 min, and 4 ml of 5% ammonium hydroxide were added. Forty ulof the sample were then diluted with 1 ml of benzene in an autosampler vial.A 1-ul sample was then injected into a Varian Model 3400 gas chromotograph(Varian, Palo Alto. CA) fitted with a Restek (Bellefonte, PA) RX-1 megabore

60 m (length) x 0.53 mm (inside diameter) x 1 urn (outside diameter) column.The injector was maintained at 100°C,while the column was maintained at135°Cfor 7 min and increased at a rate of 25°C/minto 285°Cwhich was held

for 2 min. Nitrogen was the carrier gas at a flow rate of 30 ml/min anddetection was accomplished by electron capture. The retention times of thefluorinated derivatives of IPO and the internal standard were 4.8 and 5.9 min,respectively. Peak areas were quantitated using a Nelson 3000 integrationsystem (Perkin Elmer Nelson System, Cupertino, CA), and drug concentrationswere determined from linear regression equations derived from calibrationcurves prepared with samples between 0.2 and 16 umol/liter.

Individual plasma drug disposition curves were fit using a pharmacokineticmodel characterized by two-compartment distribution of drug and first order

elimination of drug from the central compartment. The values of the followingkinetic parameters were estimated for each disposition curve: the dispositionrate constants and associated half-lives, central volume of distribution, steady-

state volume of distribution. AUC, and plasma clearance rate.Pharmacokinetic modeling was performed by nonlinear regression analysis

using PCNONLIN (Statistical Consultants, Lexington, KY). Correlationsamong kinetic parameter values and categorical toxicity data were calculatedusing Spearman's rank-order correlation statistics.

RESULTS

Fifty-five patients received 254 total courses of IPO through 17

dose levels (Table 1). All courses were évaluablefor toxicity. Themedian number of courses administered/patient was 3 and rangedfrom 1 to 18. Twenty-six patients received only 1 or 2 courses but 15

patients received more than 5 courses and 8 patients received morethan 10 courses. The median cumulative dose of IPO was 826 mg/m2and ranged from 13 to 11,092 mgs/m2. Thirteen patients had IPO

escalated due to minimal or no toxicity, including seven, two, one,two, and one patients who received IPO at two-, three-, four-, five-,and seven-dose levels, respectively. One patient was taken off study

for severe, but reversible, pulmonary toxicity. Patient characteristicsare listed in Table 2. Six patients had advanced small cell cancer thathad progressed on or within 1 to 6 months of treatment with platinum-based therapy, and 49 had non-small cell cancer. No objective re

sponses were observed; however, stable disease persisted for at leasteight courses in 15 of the 55 patients (27%).

Hepatotoxicity. Hepatotoxicity was the dose-limiting toxicity of

IPO on this schedule of administration. Hepatotoxicity was characterized by isolated elevations in the hepatocellular enzymes ALT andAST. Bilirubin and alkaline phosphatase levels remained normal even

Table 1 Dose escalations

ModifiedFibonacci40%

Increments25%

IncrementsTotalsDose(mg/m2)6.51322334ft5978104154216302422590826103212901612New

patients3334333333333374155Patientsescalated111233756Totalpatients333434445661089741Totalcourses2310111213910101013142316452672254

Table 2 Patient characteristics

Characteristic No.

No. of patients (évaluable)No. of courses (évaluable)Sex: male:femaleMedian age (range)Eastern Oncology Cooperative Group

performance status

Previous therapyRadiotherapyChemotherapy + radiotherapyNoneChemotherapy

HistologyAdenocarcinomaLarge cell carcinomaSquamous cell carcinomaSmall cell carcinomaNon-small carcinoma

Poorly differentiated

55 (55)254 (254)

43:1257 (31-75)

0-241-292-2

24149

2587663

in courses associated with NCI grade 3 (5.1-20-fold above normallimits) and grade 4 (>20-fold above normal limits) hepatocellular

enzyme elevations. Elevations in ALT were more pronounced thanelevations in AST, with peak ALT levels generally greater than peakAST levels by 1.5-2-fold. The onset of hepatocellular toxicity was

noted as early as day 2, and elevations in both AST and ALT weremaximal on day 4. The enzyme abnormalities typically resolved byday 15, and values were rarely abnormal after day 22. Hepatotoxicitywas not cumulative in that the magnitude of ALT and AST elevationswere not generally more pronounced with successive courses of therapy.

Table 3 depicts the grade of hepatocellular toxicity as a function ofIPO dose. Liver enzyme elevations were noted in only one patientduring one course administered at IPO doses below 826 mg/m2.

However, reversible and asymptomatic elevations in hepatocellularenzymes were common at IPO doses >826 mg/m2, with both the

frequency and severity of IPO-induced hepatocellular enzyme eleva

tions progressively increasing with higher drug doses. Grade 3 and 4toxicities occurred in 3 of 7 courses involving 3 of 4 patients receiving1230 mg/m2 and in 2 courses involving a single patient treated at the1612 mg/m2 dose level. At IPO doses of 1230 and 1612 mg/m2, right

upper quadrant pain and/or moderate to severe malaise were associated with grades 3 and 4 hepatotoxicity during 4 of 5 courses involv-

1796

Research. on December 14, 2020. © 1993 American Association for Cancercancerres.aacrjournals.org Downloaded from

Page 4: Phase I and Pharmacological Study of the Pulmonary ... · rationale for the clinical evaluation of IPO should be extended to include liver cancers and possibly renal cancers, as well

PHASE I AND PHARMACOLOGICAL STUDY OF 4-IPOMEANOL

Table 3 Hepatotoxidty

Dose(mg/m")104154216302422590826103212901612Totalcourses1010131423164526720910131423163781HepatotoxicityGrade1

2315

48451

2 3"1°41"

" Four courses in which hepatotoxicity was associated with reversible RUQ pain and

severe malaise.

¡ng3 of 4 patients. At 1032 mg/m2, grade 3 hepatotoxicity was

experienced by 3 of 7 patients during 5 of 26 courses. Additionally, alltoxic episodes were asymptomatic and of brief duration at this dose.Therefore, based on the lack of other hepatic and constitutional manifestations associated with hepatocellular toxicity at the 1032 mg/m2

dose level, as well as the brief and noncumulative nature of IPO-induced hepatic dysfunction, 1032 mg/m2 was determined to be the

MTD for IPO on this schedule of administration.Pulmonary Toxicity. Rigorous pulmonary monitoring by pulmo

nary function tests, arterial blood gases, and LCT revealed that IPOinduced neither dose-dependent subclinical nor clinical pulmonary

toxicity on this schedule of administration. The percentage of changein several spirometrie (FVC and FEV, ) and gas exchange (DLCO andDLCO/ALV) parameters as a function of total cumulative IPO doseare depicted in Fig. 2, A-D. Although many patients developed cu

mulative decrements in these functions of at least 20% compared withbaseline values, most decrements occurred in the setting of lung tumorprogression and were attributed to progressive disease. In addition,decrements in these spirometrie and gas exchange parameters werenot related to cumulative dose. In fact, the magnitude of decrementsin these parameters were similar among patients receiving cumulativedoses which spanned four orders of magnitude (13 to 11,092 mg/m2).

Pulmonary toxicity was clearly documented in only one of 55patients, a previously untreated 42-year-old male with adenocarci-

noma of the lung. Neither clinical nor subclinical evidence of pulmonary toxicity were noted during his first three courses of IPO administered at doses of 202 and 422 mg/m2. However, he developed

moderate pleuritic chest pain and 15% decrements in FVC, FEV,, andDLCO on day 22 of his fourth course at 422 mg/m2. Additionally,

despite a normal chest radiograph, CT scanning revealed a new pleu-ral-based pulmonary paranchymal infiltrate which was contralateral tohis primary neoplasm (Fig. 3, A-B). All symptoms, as well as pulmo

nary function and radiographie abnormalities, resolved by day 28, andthe patient was subsequently treated with four additional courses ofIPO uneventfully, including two courses each at 422 and 590 mg/m2.

However, 2 to 4 days after receiving a ninth course of IPO at 826mg/m2, he developed transient myalgias and progressive dyspnea. On

day 8, hypoxia was noted (PO2 = 50 mm Hg), and pulmonary func

tion testing revealed 50% decrements in FEV,, FVC, and DLCO.Diffuse pulmonary infiltrates were also evident on plain chest radiographs and on CT scan. His symptoms progressively worsened untilday 13, at which time his PO2 was 31 mm Hg, and panlobar infiltratesand bilateral effussions were noted on CT scanning (Fig. 3, C-D).Following an unremarkable bronchoscopic examination, transbron-

chial biopsy, and bronchoalveolar lavage, methylprednisolone, 125mg i.v. every 6 h, was begun. Thereafter, his symptoms, as well asradiographie and spirometrie abnormalities progressively improved,with complete resolution noted by day 35. IPO was not readminis-

tered.Renal Toxicity. Possible renal toxicity occurred in a 48-year-old

male with an adenocarcinoma of the lung and extensive pleural involvement. The patient developed gross, painless hematuria on days16 and 22 following treatment with his first course of IPO at 154mg/m2. The hematuria lasted 24 to 48 h on both occasions. Urinalyses

and microscopic examinations revealed proteinuria, glycosuria, andlarge numbers of RBC, RBC casts. WBC, and clumps of WBC. Aurine protein electrophoresis revealed a glomerular pattern of proteinuria. Further studies included an abdominal CT scan which revealed normal kidneys and no evidence of intraabdominal disease;unremarkable bacterial, fungal and viral urinary cultures; and stableserum levels of creatinine and urea. All gross and microscopic abnor-

Fig. 2. Changes in various pulmonary functionsincluding FVC, FEV,, DLCO. and DLCO/ALV asa function of the cumulative dose of IPO in allpatients. There was no relationship between worsening of these parameters and cumulative IPOdose. Disease progression (P) accounted for themajority of episodes in which pulmonary funcitonparameters decreased by at least 20% comparedwith baseline values. 7", parameters of the single

patient who developed severe pulmonary toxicity(see text).

100Tso

•¿�so.«••|Vo

0C8-20S°--ta•K

•¿�-90

-FVC1

I80604020TT¡¡i•¿�'••¿�npinrmiFIBS?DÃœ"FEV,nnS

nn•- Di r -°°m •¿�IT Pirnuu!iiuuyouü""IMII " 1 1 H " 1 20 ' tiu 1 "Me" H u H HHTÕp"

1 E t 1 IPTP M ÃœQp S p ?D?PPP B p -«o•¿�'•,--60-113

58 M ¡OB B» 1290 2064 «29« 11092 13 W 88 208 8» 12» 2064 4298 110K8060«

*°S

20PercentCh

S8o-00-80DLCO

Mjso40DLCO/ALVfj

, jo 1 RO•

1 1 II il il0p

pIH

.20-i. . 1 . .« . L ulk -CD. Ila•MI)¡iio-inB•¿�g -o yüyypuy|•¿�i *i-40

- pP•60-13

5fl 88 208 826 1290 2064 4298 11092 13 50 88 208 828 1290 2064 4298 ilOU?CumulativeDose (mg/m'2) Cumulative Dose <mg/m*>

1797

Research. on December 14, 2020. © 1993 American Association for Cancercancerres.aacrjournals.org Downloaded from

Page 5: Phase I and Pharmacological Study of the Pulmonary ... · rationale for the clinical evaluation of IPO should be extended to include liver cancers and possibly renal cancers, as well

PHASE l AND PHARMACOLOGICAL STUDY OF 4-1POMEANOL

Fig. 3. CT lung scans of a 42-year-old male with metastatic adenocarcinoma of the lung. A. CT scan before course 4 of IPO administered at 422 mg/irr. B. CT scan on day 22 ofcourse 4 revealing a pleural-based infiltrate (arrows) associated with pleuritic chest pain and modest decrements in pulmonary function tests. C. CT before course 8 of IPO. D. CTon day 8 of course 8 revealing diffuse pulmonary infiltrate involving the left lung and a large pleural effusion. This was associated with substantial decrements in pulmonar)' functionlests and severe hypoxia.

malities resolved by day 35. The patient was not retreated with IPOdue to tumor progression.

Miscellaneous Toxicities. Other toxicities, including nausea, vomiting, myalgias, malaise, and myelosuppression. were infrequent, clinically insignificant, and occurred at all dose levels. Modest leukopeniaand thrombocytopenia were observed in two patients, including oneminimally pretreated patient with nadir WBC counts of 3900/ul and2900/ul, respectively, during her first and second courses of IPO at422 mg/m2 and a nadir platelet count of 81.000/ul during course 2. A

second heavily pretreated patient developed mild leukopenia with anadir WBC count of 3500/ul during his only course at 522 mg/m2.

Thirteen patients experienced mild to moderate nausea and/or vomiting. Nausea without vomiting occurred during 12 courses, while vomiting occurred during 10 courses. Nausea and vomiting were not doserelated and generally occurred for 1 to 4 h immediately after drugadministration. Symptoms were never protracted and antiemetics wererarely required. In addition, one patient complained of mild myalgiaslasting 2 to 4 days after each of nine courses of IPO administered atdoses of 202, 422, 590, and 826 mg/m2. Two patients also complainedof moderate malaise during two courses at 1290 mg/m2 and severe

malaise was experienced by one patient during two courses at 1612mg/m2.

Pharmacokinetics. Complete IPO plasma disposition curves were

obtained on 44 subjects. Seven of the curves (doses 13 to 33 mg/m2)

were generated using an early version of the IPO assay which did notyield values consistent with the final version of the assay. Those datawere not included in the analysis. Individual plasma dispositioncurves are well described by a two-compartment, first order elimina

tion model (Fig. 4). The mean kinetic parameter values obtained usingthat model are listed in Table 4. At the MTD. 1032 mg/nr/day, themean half-lives, \¡and A2,are 6.7 and 114.5 min, respectively, and theplasma clearance rate is 1.16 liter/min/m2.

While individual plasma disposition curves do not suggest thepresence of nonlinear kinetics, inspection of the scatterplot of IPOdose versus AUC reveals a curvilinear relationship between the two(Fig. 5), suggesting the presence of saturable elimination kinetics. Toexplore that possibility, the average plasma concentration data of thesubjects at three different dose levels (33, 826, and 1290 mg/m2) were

simultaneously fit using a two-compartment pharmacokinetic modelwith elimination of drug treated as a Michaelis-Menten process. This

model fit the disposition data well. In addition, the AUC generated bysimulations using the model describe the observed AUC versus doserelationship very well (Fig. 4). The parameter values for the modelare: central volume of distribution. 27.5 liter/m2: steady state volumeof distribution, 210.7 liter/m2 (intercompartmental clearance rate, 1.95liter/min/m2); Vmax, 22.7 umol/liter/min-m2; and A"m,5.86 umol/liter.

1798

Research. on December 14, 2020. © 1993 American Association for Cancercancerres.aacrjournals.org Downloaded from

Page 6: Phase I and Pharmacological Study of the Pulmonary ... · rationale for the clinical evaluation of IPO should be extended to include liver cancers and possibly renal cancers, as well

PHASE l AND PHARMACOLOGICAL STUDY OF 4-lPOMEANOL

300

100-

30-

C.2 10-

gOU

3-I

1-

0.3100 200 300 400 500

time (min)600 700

Fig. 4. IPO plasma disposition curves for two representative subjects. The fits of theindividual data by a two-compartment model of IPO pharmacokinetics with first orderelimination are shown as continuous lines.

Urine IPO determinations in five subjects (doses 216 to 422 mg/m2)

revealed minimal parent drug excretion equivalent to less than 2% ofthe administered dose. Subjects treated at higher doses (>590 mg/m2)

exhibited an unidentified Chromatographie peak with kinetic characteristics consistent with a metabolite (i.e., the peak was not present inpredose plasma and increased, then decreased in magnitude roughly inparallel with the parent drug concentrations). Using IPO dispositioncurves as the input function and assuming one-compartment distribu

tion and first order formation and elimination of the species, itspharmacokinetics were modeled. The mean elimination half-life for

the species was determined to be 33 min (12 cases analyzed; range,7-88 min). Structural identification of the metabolite is pending.

Pharmacodynamics. While liver toxicity was seen only at higherdoses and larger AUC, hepatotoxicity grade did not correlate witheither IPO dose or AUC in patients who manifested hepatotoxicity,

DISCUSSION

The original rationale (4) and subsequent impetus to develop IPO asan anticancer agent specific for lung cancer was based on the novelmechanism of action of the agent, as well as the potential of IPO asa unique tissue-specific cytotoxin (5-31). Additionally, the demon

stration that both normal human lung tissue and lung cancers arecapable of metabolically activating IPO in vitro and the documentation of pulmonary toxicity as the predominant toxicity of IPO in dogsand rodents in preclinical studies suggested that pulmonary toxicitywould likely be the principal toxicity of IPO in humans. Similar topreclinical studies, toxicological effects in this study were observed inhuman lung, liver, and kidney, which presumably all contain thecytochrome P450 isoforms capable of metabolically activating IPO.However, hepatocellular toxicity, rather than pulmonary toxicity, wasthe principal toxicity in humans when the IPO was administered as a30-min infusion every 3 weeks. At 1032 mg/m2, the MTD and rec

ommended dose for subsequent phase studies of IPO on this administration schedule, hepatic toxicity was characterized by isolated elevations in hepatocellular enzymes, particularly ALT. Although NCIgrade 3 toxicity was noted in 5 of 26 courses and 3 of 7 patients at1032 mg/m2, an incidence which exceeded the criteria initially estab

lished for the MTD, liver enzyme abnormalities at this dose resolved

completely within 1 to 2 weeks, and were neither cumulative norassociated with other toxic sequalae, such as severe malaise and rightupper quadrant pain. These other toxic manifestations were commonlyassociated with hepatocellular enzyme elevations in patients receivinghigher doses of IPO.

The explanation for the difference in target organ-specific toxicity

in humans versus other species is not known. While the lack ofterminal airways containing Clara and type II alveolar cells or anyother cells capable of metabolically activating IPO in situ has adequately explained the exclusive occurrence of hepatotoxicity, ratherthan pulmonary toxicity, in birds (23, 24), all mammalian lungs studied to date, including human lungs (31), possess these specializedcells. In addition to pulmonary toxicity, hepatic binding and occasional necrosis has been noted in hamsters (14) but liver toxicity haslargely been subclinical in this species, and birds are still the onlynonhuman species known in which hepatotoxicty predominates (23,24). Although there may be substantial interspecies differences in thepharmacological behavior of IPO that may partially explain the widelydisparate MTDs between species, significantly greater and not lesspulmonary toxicity would be predicted in humans if the relative susceptibility of various species to lung toxicity were solely due toquantitative differences in pharmacological exposure to IPO. Actually,the mean AUC achieved in mice at the murine dose that was lethal in10% of mice and in dogs at the highest safe dose during preclinicalstudies were 144 ug/ml/min (860 umol/liter/min) and 55.4 ug/ml/min(330 umol/liter/min), respectively, which are equivalent to only 13.4and 5.1% of the mean AUC (6531 umol/liter/min) achieved at theMTD (1032 mg/m2) in humans (32). It is possible that inherent in

terspecies differences in the IPO activating potential and/or protectivemechanisms of lung tissues account for the different susceptibilities ofhuman, dogs, and rodents to pulmonary toxicity. This may not be thecase for hepatotoxicity since mild and subclinical elevations of hepatocellular enzymes were observed in dogs and rodents treated at IPOdoses associated with severe pulmonary toxicity, and therefore, acomparable degree of drug-induced hepatic dysfunction might be

anticipated to occur in humans, dogs, and rodents at similar IPO AUCif dosing in the lower mammals were not limited by severe pulmonarytoxicity.

This study demonstrated that IPO administered as a 30-min infusiondid not induce as a prominent feature either dose-related acute or

cumulative pulmonary toxicity, as assessed by sequential monitoringof radiographie, functional, and clinical indices. However, severepulmonary toxicity clearly occurred in one of 55 patients after thepatient received several courses of IPO at doses ranging from 202 to826 mg/m2. Except for the development of transient pleural-based

infiltrates on CT during a previous course, which were also associatedwith mild pleuritic chest pain and modest decrements in pulmonaryfunction tests, no other historical or clinical characteristics were identified as possibly predisposing the patient to the subsequent episode ofsevere pulmonary toxicity. In addition, the pharmacokinetic profile ofthe patient did not significantly differ from that of the other patients.Even if the true incidence of IPO-induced pulmonary toxicity in

humans is very low, as suggested by this study, the sporadic nature andseverity of the toxicity may have several pertinent implications withrespect to the design of subsequent clinical trials. First, the potentialseverity of the toxicity mandates that the eligibility criteria usedduring subsequent studies be somewhat restrictive to patients withsignificant pulmonary dysfunction who may not be able to toleratefurther respiratory compromise. Second, it suggests that a combination of clinical, radiographie, and functional pulmonary monitoringshould be incorporated into these investigations. The assessment ofmultiple pulmonary parameters (e.g., LCT, pulmonary function tests,physical examinations) should be continued during the next phase of

1799

Research. on December 14, 2020. © 1993 American Association for Cancercancerres.aacrjournals.org Downloaded from

Page 7: Phase I and Pharmacological Study of the Pulmonary ... · rationale for the clinical evaluation of IPO should be extended to include liver cancers and possibly renal cancers, as well

PHASE l AND PHARMACOLOGICAL STUDY OF 4-IPOMEANOL

Table 4 Mean phurmacokinetic parameter values

ModelOne-compartmentTwo-compartmentDose(mg/m~f3346597810415421630242259082610321290No.of

patients3122211323374y.«(liter/m2)25.920.527.031.465.640.016.421.427.014.125.350.739.8vss(liter/m2)53.3138.3129.349.574.5151.9125.3IV,\¡(min)2.81.83.11.72.63.03.84.24.53.85.76.77.8r/!À2(min)41.591.2106.645.264.2114.5132.3Clearance(liter/min/m2)2.81.83.11.72.63.01.51.82.01.41.31.20.9

' V0 central volume of distribution; Vss. steady state volume of distribution: i":. half-lives.

12000

10000

=£. 8000

I"O 6000

3.

^ 4000

2000

200 400 600 800 1000

Dose (mg/m2)1200 1400

Fig. 5. Relationship between IPO AUC and dose. •¿�individual subject data. Continuous line, relationship predicted by a two-compartment model of IPO pharmacokineticswith Michaelis-Menten type elimination. The Michaelis-Menten parameter values yielding the fit shown are: Vma,. 22.7 umol/min-m2; and K,n. 5.86 umol/liter.

developmental clinical trials until adequate data concerning pulmonary toxicity are collected and a scheme for optimal clinical monitoring can be developed. Finally, the occurrence of a severe episodeduring a latter course also suggests that pulmonary monitoring andclose patient follow-up continue beyond the initial one to two courses

of treatment.Nephrotoxicity also occurred in only one patient who was treated at

a relatively low IPO dose. As with the patient who developed clinically significant pulmonary toxicity, this patient did not have anyidentifiable risk factors for nephrotoxicity, nor was his pharmacoki-

netic profile sufficiently different from that of other subjects treated atidentical doses. Glomerular proteinuria, RBC, WBC, and WBCclumps in the urine suggested a primary renal process, possibly affecting the glomeruli. tubules, and/or interstitium. In preclinical studies, pathological evidence of nephrotoxicity. such as striking anddiscrete necrosis in the proximal cortical tubules with sparing of theadjacent tubules, glomeruli. and blood vessels, were limited to adultmale mice and could not be induced in either adult female or immaturemice (14, 25). Although nephrotoxicity was self limited and largelyinconsequential in this patient, tumor progression precluded retreatment, and therefore, information concerning the effects of subsequenttherapy in patients who develop IPO-induced nephrotoxicity is not

available.Although stable disease persisted for at least eight treatment courses

in 27% of the 55 total lung cancer patients studied, no objectiveantitumor responses were observed. Additionally, the eligibility requirements for the study were uniquely restrictive, thereby limiting

accrual to patients expected to possess biologically favorable characteristics, and hence, a longer survival. All patients had a good toexcellent initial performance status, and the group included 33 and 17individuals with no prior chemotherapy and radiotherapy, respectively. In essence, the large number of patients with non-small cell

lung cancer and an ideal performance status may inadvertently permita rough assessment of drug activity in non-small cell lung cancer. This

trial, albeit phase I in design, revealed little evidence of activity of IPOon this particular schedule in non-small cell lung cancer. No objectiveresponses occurred in 12 patients with non-small cell lung cancertreated with IPO doses >1032 mg/irr, the recommended phase II dose,nor in 15 patients with non-small cell lung cancer treated with IPOdoses ranging from 826 mg/m2 ( 1 dose level below the MTD) to 1612mg/m2 (2 dose levels above the MTD).

The preponderance of clinical toxicity observed in the present studyin liver rather than lung suggests that, in vivo. IPO may be preferentially activated and bound in liver rather than in lung or other tissuesin humans. However, as with the lungs of all other species exceptavians, human lung clearly has very substantial enzymatic IPO activating potential (31). Therefore, one alternative explanation for thelack of lung toxicity in this study could be that human lung tissue isrelatively more effective at detoxifying IPO than is the liver in humansor the lungs in other species. For example, it may be relevant thatpulmonary susceptibility to IPO toxicity is exquisitely sensitive totissue gluthathione status (33, 34). In turn, the pulmonary glutathionestatus may be modulated not only by enviornmental and other exogenous influences such as prior drug treatment or irradiation, but alsoby diverse endogenous or constitutive factors. Conjugation of activated IPO with gluthatione (35) prevents tissue alkylation and toxicity(33, 34). Any compromise or enhancement of this pathway may sensitize or densensitize, respectively, the lung tissue susceptibility toIPO. Also, a more specific histológica! and/or biochemical subtypingor selection of patients as to lung cancers that might be most likely torespond to IPO was not attempted in this phase I study. However, suchsubtyping is feasible using surgical specimens from individual patients (31) and may be a consideration for the design of future efficacystudies.

Irrespective of the above considerations concerning possible factorsregulating the specificity of toxicity of IPO to normal tisuses ofhumans compared with other species, the relevance to any potential HIvivo cytotoxic activity or selectivity (or modulation thereof) of IPOagainst tumors originating from lungs or other tissues is unknown.However, it remains that many human lung cancers apparently dopossess at least one fundamental attribute necessary for susceptibility;a diverse group of fresh human lung tumor biopsies were found tohave very substantial capacity to metabolically activate IPO, andindeed some tumors were even more metabolically competent thanadjacent lung tissue from the same patients (31 ). Interestingly, a more

1800

Research. on December 14, 2020. © 1993 American Association for Cancercancerres.aacrjournals.org Downloaded from

Page 8: Phase I and Pharmacological Study of the Pulmonary ... · rationale for the clinical evaluation of IPO should be extended to include liver cancers and possibly renal cancers, as well

PHASE [ AND PHARMACOLOGICAL STUDY OF 4-IPOMEANOL

recent study of fresh surgical specimens from diverse human lungcancers also revealed that such tumors contained gluthathione levels atleast equal to or exceeding adjacent normal pulmonary tissue (36).

Finally, the observations reported herein support the view that therationale for clinical evaluation of IPO in humans should be extendedto include liver cancers, and possibly also renal cancers, in addition tolung cancers. Accordingly, we have initiated an efficacy trial of IPOin hepatocellular cancer using the presently described schedule andrecommended phase II dose.

ACKNOWLEDGMENTS

The authors wish to thank Mary Duerr for assistance with data collection;Lisa Hurowit?., Chris Baba, and Tian-Ling Chen for technical assistance with

pharmacological assays; and the referring medical staff and medical and nursing staffs of the Johns Hopkins Oncology Center for the care of the patients in

this study.

REFERENCES1. Boyd. M. R., Burka. L. T.. Harris. T. M.. and Wilson. B. J. Lung-toxic furanolerpe-

noids produced by sweet potatoes (Ipomea baiatasi following microhioal infection.Biochim. Biophys. Acta. Ì37:184-195. 1974.

2. Boyd. M. R., and Wilson. B. J. Isolation and characterization of 4-ipomeanol. alung-toxic furanoterpenoid produced by sweet potatoes (Ipmeoea bátalas). J. Agrie.Food Chem.. 20: 428-430. 1972.

3. Boyd, M. R.. Wilson. B. J.. and Harris. T. M. Confirmation by chemical synthesis ofthe structure of 4-ipomeanol. a lung-toxic metabolite of the sweet potato. Ipomoeabatatas. Nature N. Biol.. 236: 158-159. 1972.

4. Boyd. M. R. Evidence for the clara cell as a site of cytochrome P450-dependent mixedfunction oxidase activity in lung. Nature (Lond.). 269; 713-715, 1977.

5. Christian. M. C. Wittes. R. E.. Leyland-Jones. B.. McLemore. T. L.. Smith. A. C.Grieshaber, C. K.. Chabner. B. A., and Boyd. M. R. 4-ipomeanol: a novel investiga-lional new drug for lung cancer. J. Nati. Cancer Inst.. 81: 1133-1143. 1989.

6. Boyd, M. R. Biochemical mechanisms of chemical-induced lung injury: roles ofmetabolic activation. CRC Crii. Rev. Toxicol.. 7: 103-176, 1980.

7. Boyd, M. R., and Reznik-Schuller. H. M. Metabolic basis for the pulmonary Clara cellas a target for pulmonary carcinogenesis. Toxicol. Palhol., 12: 56-61, 1984.

8. Boyd, M. R. Role of metabolic activation in the pathogenesis of chemically inducedpulmonary disease: mechanism of action of the lung toxic furan, 4-ipomeanol. Environ. Health Perspect., 16: 127-138, 1976.

9. Boyd. M. R. Metabolic activation and lung toxicity: a basis for cell selective pulmonary damage by foreign chemicals. Environ. Health Perspect.. 55: 47-51. 1984.

10. Boyd. M. R.. Grygiel. J. G.. and Minchin. R. F. Metabolic activation as a basis fororgan-selective toxicity. Clin. Exp. Pharmacol. Physiol., 10: 87-107. 1983.

11. Boyd. M. R., Burka. L. T. and Wilson. B. J. Distribution, excretion, and binding ofradioactivity in the rat after intraperiloneal administration of the lung-toxic furan|MC]4-ipomeanol. Toxicol. Appi. Pharmacol., 32: 147-157. 1975.

12. Boyd, M. R.. and Burka. L. T In vivo studies on the relationship between target organalkylation and the pulmonary toxicity of a chemically reactive metabolite of 4-ipomeanol. J. Pharmacol. Exp. Ther.. 207.- 687-697, 1978.

13. Boyd, M. R., Burka. L. T. Wilson. B. J.. and Sasame. H. A. In vitro studies on themetabolic activation of the pulmonary toxin 4-ipomeanol by rat lung and livermicrosomes. J. Pharmacol. Exp. Then. 207: 677-686, 1978.

14. Dutcher. J. S.. and Boyd. M. R. Species and strain differences in target organatkylation and toxicities by 4-ipomeanol. Predictive value of covalent binding instudies of target organ toxicities by reactive metabolites. Biochem. Pharmacol., 28:3367-3372. 1979.

15. Boyd. M. R.. Sasame. H. A., and Franklin. R. B. Comparison of ratios of covalentbinding to total metabolism of the pulmonary toxin. 4-ipomeanol, in vitro in pulmonary and hepatic microsomes and the effects of pretreatment with phenobarbitol or3-methylchlorantrene. Biochem. Biophys. Res. Commun., 193: 1167-1172. 1980.

16. Devereux. T. R.. Jones. K. G.. Bend, J. R.. Fonts. J. R.. Statham. C. N.. and Boyd. M.

R., In vitro metabolic activation of the pulmonary toxin, 4-ipomeanol. in noncilialed

hronchiolar epithelial (Clara) and alveolar type II cells isolated from the rabbit lung.J. Pharmacol. Exp. Ther. 220: 223-227. 1982.

17. Parandoosh. Z.. Fujita. V. S.. Coon. M. J.. and Philpot. R. M. Cytochrome P-450isozymes 2 and 5 in rabbit lung and liver. Drug Metab. Disposition. /5: 59-67. 1987.

18. Slaughter. S. R.. Statham. C. N.. Philpot, R. M.. and Boyd, M. R. Covalent bindingof metabolites of 4-ipomeanol to rabbit pulmonary and hepatic microsomal proteinsand to the enzymes of the pulmonary cytochrome P450-dependent monooxygenasesystem. J. Pharmacol. Exp. Ther, 224: 252-257. 1983.

19. Boyd. M. R. Effects of inducers and inhibitors on drug metabolizing enzymes anddrug toxicity in extrahepatic tissues. Ciba Found. Symp.. 76: 43-66, 1980.

20. Wolf. C. R., Slatham, C. N.. McMenamin, M. G.. Bend. J. R.. Boyd, M. R.. andPhilpot. R. M. The relationship between the catalytic activities of the lung-specifictoxicity of the furan derivative. 4-ipomeanol. Mol. Pharacol.. 22: 738-744. 19X2.

21. Sasame H. A.. Gillette. J. R.. and Boyd. M. R. Effects of anti-NADPH-cytochrome creducíaseand anti-cylochrome b5 antibodies on the hepatic and pulmonary microsomal metabolism and covalent binding of the pulmonary toxin 4-ipomeanol. Biochem.Biophys. Res. Commun.. K4: 389-395. 1978.

22. Boyd. M. R.. and Dutcher, J. S. Renal toxicity due to reactive metabolites formed insitu in the kidney: investigations with 4-ipomeanol in the mouse. J. Pharmacol. Exp.Ther.. 216: 640-647. 1981.

23. Buckpitt. A. R.. Statham. C. N.. and Boyd. M. R. In vivo studies on the target tissuemetabolism, covalent binding, glutathione depletion, and toxicily of 4-ipomeanol inbirds, species deficient in pulmonary enzymes for metabolic activation. Toxicol. Appi.Pharmacol.. 65: 38-52. 1982.

24. Buckpitt. A. R.. and Boyd, M. R. Metabolic activation of 4-ipomeanol by avian tissuemicrosomes. Toxicol. App. Pharmacol.. 65: 53-62, 1982.

25. Durham. S. K., Boyd. M. R.. and Castleman, W. L. Pulmonary endoihelial andbronchiolar epithelial lesions induced by 4-ipomeanol in mice. Am. J. Pathol., Ufi:66-75. 1985.

26. Smiih. A. C.. Barrel. D.. Sledham. M. A., et al. Preclinical toxicology studies of4-ipomeanol: a novel candidate for clinical evaluation in lung cancer. Cancer Treat.Rep.. 71: 1157-1164, 1987.

27. Clinical Brochure. 4-lpomeanol NSC 349438. Belhesda. MD: Division of CancerTrealmenl. National Cancer Instiiute. February. 1988.

28. Falzon. M. McMahon. J. B.. Schuller. H. M.. and Boyd. M. R. Metabolic activationand cytotoxicity of 4-ipomeanol in human non-small cell lung cancer lines. CancerRes.. 46: 3483-3489. 1986.

29. McLemore T. L.. Liu. M. C.. Blacker, P. C., Gregg. M.. Alley. M. C.. Abbotl. B. J..Shoemaker. R. H.. Bohlman. M. E.. Liltersl. C. C.. Hubbard. W. C.. Brennan. R. H..McMahon. J. B.. Fine. D. L.. Eggleston. J. C.. Mayo. J. G.. and Boyd. M. R. Novelintrapulmonary model for orthotopic propagation of human lung cancers in athymicnude mice. Cancer Res.. 47: 5132-5140. 1987.

30. McLemore, T, Coudert, B.. Adelberg, S., Liu. M. C.. Hubbard. W. C.. Litters!, C. C..Eggleston, J. C.. and Boyd. M. R. Metabolic activation of 4-ipomeanol by humanpulmonary carcinoma cells propagated in vitro and intrabronchially in nude mice.Clin. Res., 36: 498A, 1988.

31. McLeMore, T. L, Litterst. C. C., Coudert. B. P., Liu. M. C.. Hubbard. W. C.. Adelberg,S., Czerwinki. M., McMahon. N. A., Eggleston. J. C.. and Boyd, M. R. Metabolicactivation of 4-ipomeanol in human lung, primary pulmonary carcinomas, and established human pulmonary carcinoma cell lines. J. Nail. Cancer Inst.. 82: 1420-1426.1990.

32. Preclinical toxicology evalualion of 4-ipomeanol (NSC-349438) in CD2F| mice,Fischer 344 rals, and beagle dogs. Summary, pharmacokinelics. and mice lethalitystudies. Final Report. Vol. I. Kansas City. MO: Midwest Research Institute. December 4. 1986.

33. Boyd. M. R., Sliko. A. S.. Statham. C. N.. and Jones. R. B. Protective role ofendogenous pulmonary glutathione and other sulfhydryl compounds against lungdamage by alkylating agents. Investigations with 4-ipomeanol in the rat. Biochem.Pharmacol.. 31: 1579-1583. 1982.

34. Slatham. C. N.. and Boyd. M. R. Distribution and metabolism of the pulmonaryalkylating agent and cytotoxin. 4-ipomeanol. in control control and diethylmaleate-treated rats. Biochem. Pharmacol.. 31: 1585-1589. 1982.

35. Buckpitt. A. R.. and Boyd. M. R. The in vitro formation of gluthathione conjugateswith the microsomally activated pulmonary bronchiolar alkylating agent and cyto-loxin. 4-ipomeanol. J. Pharmacol. Exp. Then. 2/5: 97-103. 1980.

36. Cook. J. A.. Pass, H. I., lype. S. N., Friedman. N.. DeGraff. W.. Russo. A., andMitchell. J. B. Cellular gluthathione and thiol measurements from surgically resectedhuman lung tumor and normal lung tissue. Cancer Res., 51: 4287^4294. 1991.

1801

Research. on December 14, 2020. © 1993 American Association for Cancercancerres.aacrjournals.org Downloaded from

Page 9: Phase I and Pharmacological Study of the Pulmonary ... · rationale for the clinical evaluation of IPO should be extended to include liver cancers and possibly renal cancers, as well

1993;53:1794-1801. Cancer Res   Eric K. Rowinsky, Dennis A. Noe, David S. Ettinger, et al.   Patients: Hepatotoxicity Is Dose Limiting in Humans4-Ipomeanol on a Single Dose Schedule in Lung Cancer Phase I and Pharmacological Study of the Pulmonary Cytotoxin

  Updated version

  http://cancerres.aacrjournals.org/content/53/8/1794

Access the most recent version of this article at:

   

   

   

  E-mail alerts related to this article or journal.Sign up to receive free email-alerts

  Subscriptions

Reprints and

  [email protected] at

To order reprints of this article or to subscribe to the journal, contact the AACR Publications

  Permissions

  Rightslink site. Click on "Request Permissions" which will take you to the Copyright Clearance Center's (CCC)

.http://cancerres.aacrjournals.org/content/53/8/1794To request permission to re-use all or part of this article, use this link

Research. on December 14, 2020. © 1993 American Association for Cancercancerres.aacrjournals.org Downloaded from


Recommended