Pharmac. Ther. Vol. 52, pp. 385-406, 1991 Printed in Great Britain. All rights reserved

0163-7258/91 $0.00+ 0.50 © 1992PergamonPress Ltd

Associate Editor: J. G. CORY

CISPLATIN RESISTANCE IN HUMAN CANCERS K. J. SCANLON,* M . KASHANI-SABET, T. TONE a n d T. FUNATO Biochemical Pharmacology, Department of Medical Oncology, Montana Building, City of Hope Medical Center, 1500 E. Duarte Road, Duarte, CA 91010, U.S.A.

Abstract---Cancer chemotherapeutic agents primarily act by damaging cellular DNA directly or indirectly. Tumor cells, in contrast to normal cells, respond to cisplatin with transient gene expression to protect and/or repair their chromosomes. Repeated cisplatin treatments results in a stable resistant cell line with enhanced gene expression but lacking gene amplification for the proteins that will limit cisplatin cytotoxicity. Recently, several new human cell lines have been characterized for cisplatin resistance. These cell lines have led to a better understanding of the molecular and biochemical basis of cisplatin resistance. The c-fos proto-oncogene, a master switch for turning on other genes in response to a wide range of stimuli, has been shown to play an important role in cisplatin resistance both in vitro and in patients. Based on these studies, new strategies have been developed to circumvent and/or exploit clinical cisplatin resistance.

CONTENTS I. Introduction 2. Mechanisms of Cisplatin Resistance 2.1. Cisplatin transport 2.2. Sulfhydryl peptides 2.3. Repair of cisplatin-DNA adducts 2.4. The signal transduction pathway 3. Clinical Resistance 3.1. Patient tumors 3.2. Early detection of drug resistance 4. Circumventing Cisplatin Resistance 4.1. Platin analogs 4.2. Modulators of cisplatin resistance 4.3. The fos oncogene in drug resistance 5. Exploiting the Mechanisms of Cisplatin Resistance 5.1. Thymidine salvage pathway 5.2. Nucleoside analogs 5.3. Pyrimidine pathway 6. Conclusions Acknowledgements References

1. I N T R O D U C T I O N There has been a rapid evolution in understanding the multistage process of cancer (Weinberg, 1989; Baker et al., 1990; Levine et al., 1991); unfortunately developing clinical strategies to exploit this finding have not followed as quickly (Antman, 1991; Perez et al., 1991a). In addition, both natural and acquired resistance to cancer chemotherapeutic agents have also limited the cure rates (Bertino, 1988). Patients treated with these agents have approximately a 50% response rate, but unfortunately, within 12 to 24 months the majority of these patients relapse

385 386 386 387 387 388 391 391 391 392 392 393 395 396 396 396 397 397 398 398

(DeVita, 1983; Frei, 1985; Louie et al., 1986). Cisdiamminedichloroplatin(II) (cisplatin) and its analogs are some of the most widely used cancer chemotherapeutic agents either alone or in combination with other agents in the treatment of cancer (Rosenberg, 1985). If one understands the molecular and biochemical basis for the development of cisplatin resistance then rational strategies to circumvent or exploit these properties could be developed for clinical protocols. During the past few years several reviews on cisplatin have been published (for an overview see Rosenberg, 1985), cisplatin structure (Sherman and 385

386

K.J. SCANLONet al.

Lippard, 1987); DNA repair in vitro (Fox and Roberts, 1987); DNA repair in patients (Scanlon et al., 1989a); clinical pharmacology (Loehrer and Einhorn, 1984; Ozois and Young, 1984; de Graeff et al., 1988; Muggia, 1991) cellular pharmacology (Reed and Kohn, 1990; Andrews and Howell, 1990); and for strategies to overcome cisplatin resistance (Scanlon et al., 1989b; Kelley and Rozencweig, 1989; Perez et al., 1990). The focus of this review will be to highlight new areas of cisplatin research that may provide novel treatment regimens to overcome cisplatin resistance in human tumors. This review will include only some references to cisplatin resistance in mouse model systems because of the poor correlation between cisplatin resistant properties in mouse cells (Sheibani and Eastman, 1990) when compared to human carcinomas, in vitro (Scanlon et al., 1989b) or patient tumors (KashaniSabet et al., 1990a). During the past few years a significant number of new human cell lines have been developed for studies on cisplatin resistance that reflect the type of patient tumors most widely treated with cisplatin. These new cisplatin-resistant human cell lines were made in vitro for the following types of tumors: ovarian carcinoma (Hamilton et al., 1985; Andrews et al., 1985; Kikuchi et al., 1986, 1990; Hill et al., 1987; Wolfet al., 1987; Lu et al., 1988; Kuppen et al., 1988; Hills et al., 1989; Terheggen et al., 1990; De Pooter et al., 1991; McLaughlin et al., 1991); head and neck carcinoma (Frei et al., 1985); prostate carcinoma (Metcalfe et al., 1986); lung carcinoma (Hospers et al., 1988, 1990b; Hong et al., 1988; Bungo et al., 1990; Kasahara et al., 1991); colon carcinoma (Scanlon et al., 1989c; Fram et al., 1990); breast carcinoma (Teicher et al., 1991b); bladder carcinoma (Bedford et al., 1987; Walker et al., 1990); fibrosarcoma (Chu and Chang, 1990); melanoma (Teicher et al., 1991b); testicular teratoma (Hill et al., 1990; Walker et al., 1990); and leukemia (Shionoya et al., 1986). These human cell lines resistant to cisplatin have been shown to have multifactorial mechanisms of resistance because these cells have had multiple types of cisplatin treatment. Administration of cisplatin has varied by either dose (high/low) and/or by exposure (continuous or pulse). Which mechanism is clinically relevant? Patients who are treated with cisplatin have been shown to attain serum levels of 10 to 50/~M with a clearance rate (T 1/2) of about 30 to 60 min (Rosenberg, 1985). Thus, human cell lines made resistant to cisplatin with one hour pulses of 50 # i cisplatin (Scanlon et aL, 1989b,c) may reflect the type of resistance found in patients treated with cisplatin (Kashani-Sabet et al., 1990a). Furthermore, unlike methotrexate resistance or the multidrug resistant phenotype where gene amplification and high levels of drug resistance (100-fold or greater) can be achieved in vitro, cisplatin resistance, in contrast, has relatively low levels of resistance (2- to 30-fold).

2. MECHANISMS OF CISPLATIN RESISTANCE 2.1. CISPLATINTRANSPORT Studies of cisplatin transport, efflux and accumulation in human cancer cells remain confusing because of terminology as well as the design of the transport experiments (Andrews et al., 1988; Andrews and Howell, 1990). Cisplatin, a highly reactive electrophile, interacts with the cellular molecules and this property of cisplatin will alter any interpretation of uptake and efflux measurements. Mammalian transport systems are usually characterized by non-metabolizable substrates, such as ct aminoisobutyric acid for the amino acid transport system (Shionoya et al., 1986). Unfortunately, an unreactive form of cisplatin is not available. Interpretation of cisplatin transport defects in resistant cells is further complicated by experiments using long incubation times (up to 24 hr) that do not discriminate between uptake from accumulation or the use of high concentrations of cisplatin (up to 1 mM) to measure uptake since patient tumors are rarely exposed to these concentrations of cisplatin. In spite of these limitations, several studies have suggested that a 25-75% decrease in cisplatin uptake contributes to cisplatin resistance in cell lines that have a 3- to 30-fold level of resistance. The clinical relevance of this small decrease in cisplatin accumulation as a mechanism of cisplatin resistance still remains unclear. Recently two laboratories have suggested that a membrane protein has been associated with cisplatin resistance. In one study with a specific type of mouse LI210 leukemia subline, a protein (200 kd) increased with a corresponding increase in cisplatin resistance (Kawai et al., 1990). The cell line from which this protein was isolated was 40-fold resistant to cisplatin but had only a 2-fold decrease in cisplatin accumulation. By contrast, in a second study, a protein was shown to decrease with increased cisplatin or methotrexate resistance (Bernal et al., 1990). The exact function of these proteins remains unknown. This is not the first time changes in membrane properties have been associated with cisplatin and methotrexate resistance (Scanlon et al., 1987; Kashani-Sabet and Scanlon, 1989). Several common pathways for methionine/folate metabolism and DNA synthesis/repair are activated with resistance to either cisplatin or methotrexate (see Sections 2.3 and 2.4). Thus, these proteins for cisplatin transport and/or efflux must be further characterized before any causal relationship can be established. Human ovarian carcinoma 2008 cells have also been made resistant to cisplatin (1.5- to 3.0-fold) in vivo by weekly courses of cisplatin (3 mg/kg) (Andrews et al., 1990). This resistant cell line was not stable and reverted to the sensitive phenotype within 20 to 40 days. A transport defect was suggested to explain that the resistance was associated with a 25% decrease in cisplatin accumulation. However, other

Cisplatin resistance in human cancers mechanisms of resistance were not investigated. Similar studies with the ovarian A2780 ceilswere also performed in athymic mice, but no mechanism of resistance was characterized (Rose and Bailer, 1990). In a third study, tumors were shown to develop mechanisms of cisplatin resistance which were found only in vivo (Teichcr et al., 1990), but not with in vitro mechanisms of cisplatin resistance (Teichcr et al., 1987). In contrast, tumors from ovary or colon patients who relapsed from cisplatinbased treatment have elevated levels of genes expressing enzymes associated with D N A repair (Kashani-Sabet et aI., 1990a) and these studies are consistent with data on ovarian and colon celllinesin culture (Scanlon et aI., 1989b). This topic will bc discussed in more detail in Section 3.1. 2.2. SULFHYDRYLPEPTIDES Glutathionine (GSH) is one of the most prevalent intracellular thiols and has been shown to be involved directly or indirectly with many cellular functions (Meister and Anderson, 1983). Human cells have been shown to have up to 3.3-fold increases in the level of GSH in cisplatin resistant cells; however these cells were also 30-fold resistant to cisplatin (Teicher, et al., 1987; Andrews and Howell, 1990; Perez et al., 1990). Only cisplatin resistant cells that have been exposed continuously to cisplatin (high dose or low dose) have shown these elevated levels of GSH. Correlating GSH levels with cisplatin resistance is problematic since there is such a relatively small concentration of cellular cisplatin in contrast to the relatively high levels of GSH in the cell. Clinically, the role of GSH in cisplatin resistance has also been inconclusive. The use of D,L-buthionine-(S,R)sulfoximine (BSO) has also shown to be effective in modulating melphalan-DNA damage but to a lesser extent cisplatin damage (Ozols et al., 1987). This will be discussed in more detail in Section 4.2. Metallothioneins are small proteins that are important in binding and detoxifying heavy metals such as zinc, cadmium and copper. Metallothionein gene expression can be modulated by the H-ras protooncogene (Hamer, 1986) and may be due to a responsive element in the promoter region of the metallothionein gene. Several reports (Andrews et al., 1987; Kelley et al., 1988; Schilder et al., 1990) have suggested that some but not all cell lines resistant to cisplatin may have increased expression for metallothioneins (for a review, Andrews and Howell, 1990). It has been difficult to correlate the level of metallothioneins and cisplatin resistance until recently when a human small cell lung carcinoma was shown to have elevated levels of metallothionein 1.6and 2.3-fold in 6.2- and 10.0-fold cisplatin resistant cells, respectively. However, no other mechanism of resistance was identified to account for the remaining resistance (Kasahara et ai., 1991). In contrast, 48 patients were recently analyzed for changes in metalJF'L" 52/3,--!

387

lothionein levels in ovarian tumors before and after chemotherapy. The conclusion from these studies was that metallothionein content was not a major determinant of tumor sensitivity to chemotherapy (Murphy et al., 1991). There may be several reasons for this observation including the fact that the metallothionein gene expression is transient and under the control of several types of promoters, in particular its AP-I binding site (see Section 2.4). Cisplatin can elicit a cellular response along the signal transduction pathway to express genes necessary to protect itself. Metallothionein is one of these genes and depending on the cell line and the level of transient expression, this could explain the variable expression of this gene in response to cisplatin (Lazo and Basu, 1991). The section on signal transduction pathways (2.4) will discuss this concept in more detail. 2.3. REPAIROF CISPLATIN-DNA ADDUCTS

It has been well established that cisplatin binds to DNA and these adducts contribute to cellular toxicity. In most studies it is believed that the cell's ability to repair either Pt--GG or Pt-GA adducts contributes to cisplatin resistance (Ducore et al., 1982; Fichtinger-Schepman et al., 1985; Eastman, 1986; Sherman and Lippard, 1987; Bedford et al., 1988; Eastman and Schulte, 1988; Schwartz et al., 1989; Hospers et al., 1990a; Lemaire et al., 1991; Shellard et al., 1991; Eastman, 1991). There have been several indirect correlations suggesting DNA repair plays an important role in cisplatin cytotoxicity. Repair-deficient human xeroderma pigmentosum fibroblasts are more sensitive to cisplatin than normal fibroblasts, possibly due to their limited excision repair capacity (Dijt et al., 1988; Hansson et al., 1991). There also appears to be a correlation between repair of cisplatin-DNA damage in vitro and cytotoxicity (Roberts and Friedlas, 1987; Scanlon et al., 1990b) including data from clinical samples (Reed et al., 1987, 1990; Kashani-Sabet et al., 1990a). In a methotrexate-resistant Chinese hamster ovary cell line with elevated dihydrofolate reductase, cisplatin-DNA adducts were removed faster in the transcribed region of dihydrofolate reductase gene and c-myc gene than in the non-coding region of the f o s gene (Jones et al., 1991). These data suggest that selective removal of these adducts may contribute to the resistant phenotype. It will be interesting to learn the mechanism of this preferential repair. Recently, more indirect evidence for repair of cisplatin-DNA adducts has been demonstrated using cisplatindamaged plasmids in cell lines sensitive and resistant to cisplatin. In vitro replicating systems such as the pRSVcat plasmid (Sheibani et al., 1989; Chao et al., 1990), or pSVO plasmid (Heiger-Bernays et al., 1990) or the pBR322 plasmid (Hansson and Wood, 1989) have been used in cells or cell lysates to measure repair of specific cisplatin-DNA damage. The

388

K.J. SCANLONet al.

plasmid DNA is an artificial system for measuring repair and the question still remains, do these cells repair their own DNA in a similar manner? In a recent series of papers, it appears that cells possess a protein that can recognize only cisplatinGG or cisplatin-GA adducts (Chu and Chang, 1988). These proteins have been termed damage recognition proteins (Toney et al., 1989; Donahue et al., 1990) or xeroderma pigmentation binding factors (Chu and Chang, 1990). Identification of these proteins was based on demonstrating changes in gel mobility of labeled cisplatin-DNA adducts with nuclear extracts (Andrews and Jones, 1991). The specificity of these proteins appears to be unique for cisplatin-DNA adducts and not for other types of platin adducts or UV adducts (Fujiwara et al., 1990). Although these studies are interesting, several questions still need to be addressed: (i) Have cells recently evolved a binding protein that only recognizes cisplatin adducts? (ii) why is there a lack of correlation between the amount of these proteins and cisplatin cytotoxicity in different cell types? Obviously, these binding proteins must recognize a change in the conformational structure of DNA, not a cisplatin-DNA adduct and unfortunately the level of these proteins does not correlate with the level of cisplatin resistance. The ability to repair cisplatin-DNA adducts presupposes the availability of all four nucleotide triphosphates. The patch size for repair of cisplatin-DNA adducts appears to be about 140-bases from plasmid studies (Hansson and Woods, 1989). In addition, when cells are exposed to cisplatin, there is enhanced degradation of the DNA perhaps to protect the chromosomes from cisplatin (Scanlon et al., 1990a). This degraded (i.e. decoy) DNA could be used by the cell to detoxify cisplatin in a similar manner to that by which GSH or metallothionein protects the cell. Why does cisplatin selectively perturb the thymidine pools? Previous studies have shown that cisplatin alters methionine uptake into the cells which stimulates the dTMP synthase cycle and produces more methionine (Kashani-Sabet and Scanlon, 1989). The catabolism of homocysteine and serine results in the accumulation of both methionine as well as methylenetetrahydrofolate, a key co-factor for dTMP synthase and DNA synthesis (see Section 5.1 also). This was the biochemical basis for cisplatin/ 5-fluorouracil synergy in vitro and clinically (see Section 5.3). The enzymes necessary to repair cisplatin-DNA damage may involve the DNA polymerases a and fl (Lai, et al., 1988, 1989; Scanlon et al., 1989a,b,c). However, 2008 ovarian carcinomas did not have elevated mRNA levels for DNA polymerase ~, and fl 18 hr after cisplatin treatment (Katz et al., 1990). Other studies have shown enhanced levels of DNA polymerase fl between 3 and 9 hr after cisplatin treatment (Scanlon et al., 1990b) suggesting transient expression in response to cisplatin. The role for a second type of repair enzyme, ECRR-I/2 has been

implicated in cisplatin resistance but the data is also not conclusive due in part to its limited expression (Hoeijmakers and Bootsma, 1990). Dihydrofolate reductase, dTMP synthase and topoisomerases have also been implicated in cisplatin resistance (Scanlon and Kashani-Sabet, 1988; Scanlon et al., 1990b, 1991). What do all these enzymes have in common? There are several steps by which cells repair D N A damage: These enzymes mentioned above and others could exist as a multienzyme complex that respond to nuclear oncogene expression and repair the damaged chromosomes (Reddy and Pardee, 1980; Pardee, 1987). In addition, some of the genes have AP-I binding domains which can respond to c-fos oncogene expression (Scanlon et al., 1991). If the cells respond to cisplatin through a cascade involving nuclear oncogenes then genes having an AP-1 binding site would be targets for initiating repair. This will be discussed in more detail in the next section. Finally, if these repair processes are important in cisplatin resistance then by modulating expression of these enzymes one should be able to reverse drug resistance (see Sections 4 and 5). 2.4. THE SIGNALTRANSDUCTIONPATHWAY DNA-damaging agents induce the expression of specific genes and such transcriptional modifications can influence cell cytotoxicity (Scanlon et al., 1991). Cell growth and differentiation can be initiated by signal cascades (Edelman et al., 1987; Hunter, 1991). Cancer chemotherapeutic agents have been shown to disrupt this signal transduction pathway which may contribute to the evolution of drug resistant clones (Tritton and Hickman, 1990). In turn, the nuclear oncogene c-los shows increased expression in cisplatin resistant cells in culture (Scanlon et al., 1988, 1989c; Hollander and Fornace, 1989) and in patients (Scanlon et al., 1988, 1989b; Kashani-Sabet et al., 1990a). If this oncogene plays a role in cisplatin resistance then modulating c-los RNA expression should change cisplatin's cytotoxicity. The c-los oncogene has been shown to modulate gene expression by stimulating other genes that possess the AP-I binding domain (i.e. dTMP synthase, topoisomerase I and metallothionein). In addition, a l o s ribozyme (catalytic RNA) has also been shown to completely reverse cisplatin resistance by modulating these repair genes (Scanlon et al., 1991; see Sections 4.2 and 4.3). Thus, what is the relationship between cisplatin resistance, nuclear oncogene expression and the signal transduction pathway? Regulation o f l o s gene expression is mediated by combinations of short DNA sequences that provide binding sites for sequence-specific transcription factors. The interaction between individual trans- and cis-acting elements which determine the rate of gene transcription are thought to represent the ultimate steps in various signal transduction pathways (Fig. 1). Two of the major classes of regulatory elements that

389

Cisplatin resistance in human cancers Genes with A P responsive elements •

Extracellular stimuli

(I) Growth factors (EGF)-~ TPA

(1mE)

l

Ca=+ionophoros D metal ions 3

Ca=÷

PKC ~

I TGACTCAI c-fos/c-lun ~- leuane Zlpl~r ~

(II) uv

DNA-damaging agents

cA.p

PKA

I ar carCAC I

• •

c-los TS DNA pol o • Topoi • hMTIIA o • GST • MDR o

~TIF/CRE)

(III)

Genes with ras responsive elements o FIG. 1. Signal transduction pathways involved in cisplatin resistance. (See Section 2.4 for abbreviations and details.) Cisplatin

~" methionine ~"

Hras

contribute to transcriptional regulation by extracellular signals are the AP-I/TRE and ATF/CRE sequence motifs. The AP-1/TRE element was originally defined as the activator protein 1 (AP-I) binding site or the phorbol 12-O-tetradecanoate 13-acetate (TPA) responsive element (TRE) (Lee et al., 1987b); the ATF/CRE element was defined as the activating transcription factor (ATF) binding site or the cAMP responsive element (CRE). The AP-I/TRE site is recognized by a group of proteins including those encoded by the c-fos and c-jun gone families (Abate et al., 1991). The ATF/CRE site is recognized by a family of proteins referred to as ATF or CRE binding proteins (CREB) (Hai and Curran, 1991). This family of proteins has been implicated in cAMP-, calciumand virus-induced alterations in transcription. AP-1 is the collective name for a class of transcription factors (i.e. serum responsive elements) that has been characterized by the ability to bind to the promoter/enhancer elements containing the TGACTCA sequence. The members of this family include c-Fos, FosB, Fra-1, Fra-2, c-Jun, JunB and JunD. All of these proteins share a conserved region, consisting of a leucine repeat dimerization domain (ieucine zipper). The functional form of AP-I is a dimer composed of a permutation of one or more of these proteins. Several such elements have been previously described: e.g. The hormone responsive element (HRE), the metal responsive element of the human metallothionein IIA gone (MRE); the TPA responsive element (TRE), the cAMP responsive element (CRE) and the estrogen responsive element (ERE). However, individual genes often respond to more than one extracellular agent each of which may use a different signal transduction pathway. The human metallothionein IIA gone (see Section 2.2) is induced by heavy metals (cisplatin), glucocorticoid hormones, phorbol esters, interleukin 1, interferon, various serum factors and u.v. light (Lee et aL, 1987a;

Buscher et al., 1988; Sassone-Corsi et al., 1988). This, in principle, could be accounted for by two basic mechanisms. First, either the gone possesses distinct cis-acting genetic elements, each one responding to a different stimulus and recognized by a unique transacting factor; or several signal pathways converge and operate through a single cis element recognized by one or several trans-acting factors whose activity is subject to complex control. Recently, an AP-I site has also been shown to be located on the MDR gone (Teeter et al., 1991) and this AP-I site may explain the cyclosporin A (CSA) effect on a wide range of cancer chemotherapeutic agents (see Section 4.2). An example of the second possibility is the transcription factor AP-2 which mediates gone induction by agents which elevate intracellular cAMP and by agents which activate protein kinase A (Buseher et al., 1988). Both types of mechanisms are involved in the activation of the protooncogene c-fos, which responds to a particularly large number of extracellular agents. These agents include growth factors, phorbol esters, Ca2+-ionophores, thyrotropic hormones, metal ions (such as K +, Cd 2+, Zn 2+ and cisplatin), cAMP, heat shock proteins, DNA damaging agents and u.v. irradiation (Buscher et al., 1988). Cisplatin has also been shown to stimulate heat shock protein (HSP35) gone expression (Oesterreich et aL, 1991). Induction of the cAMP signal-transduction pathway significantly increases c-fos transcription. Several studies have demonstrated that the transcriptional activator cAMP-responsive element (CRE)-binding protein (CREB) is responsible for c-fos stimulation upon cAMP induction (Foulkes et aL, 1991; Hartig et al., 1991). CREB protein binds as a dimer to CREs and exerts its transcriptional regulatory function upon several genes when in a phosphorylated form. At least 10 different genes encode CRE-binding proteins which include the genes for CREB, ATF-I, CRE-BPI, ATF-~, ATF-3 and ATF-4. A link

390

K.J. SCANLONet al.

between the action of the EIA proteins and the cAMP-dependent consensus cAMP response element (CRE) 5'-TGACGTCA-3' is located upstream of the E1A-inducible adenovirus early genes E2A, E2, E3 and E4. Cellular factors that recognize these sequences have been identified, including the CREB/ATF and AP-I families. Moreover, the combination of E IA protein and cAMP activates the E1A and E4 genes synergistically, indicating that the cAMP-mediated and E1A-mediated responses are somehow linked or interactive. EIA also acts together with cAMP to regulate the level of transcription factor AP-1 (Engel et al., 1991). AP-I activity has been implicated in both positive and negative transcriptional regulation as well as in chromosomal DNA replication (Murakami et al., 1991; Ginsberg et al., 1991). The role of DNA polymerase fl in the cell has been linked to DNA repair gap-filling synthesis through inhibitor studies and the enzyme also could have a role in gap-filling DNA synthesis (Wang, 1991; see Section 2.3). The core promoter for human DNA polymerase fl contains one 10-residue DNA palindromic sequence G T G A C G T C A C at position -49 to -40 that is similar to the binding site for ATF/CREB proteins (Widen and Wilson, 1991; Kedar et al., 1991; Englander et al., 1991). In experiments with a fl-pol promoter, the integrity of this palindromic element was essential for full promoter activity, EIA/E1B transactivation, stimulation by activation of H-ras protein and in mediating response to the DNA damaging agents. This site is known to mediate E1A inducibility in several viral promoters and to confer response to elevated levels of cAMP in many cellular genes. In these promoters, the sequence TGAGTCA is referred to as the cAMP responsive element (CRE) or activating transcription factor (ATF) binding site. DNA polymerase fl and metallothionein also have an H-ras responsive element that responds to changes in H-ras gene expression (Schmidt and Hamer, 1986; Kedar et al., 1990). Sklar (1988) and others (Niimi et al., 1991) have shown a correlation with cisplatin resistance and H-ras gene expression. These studies have been confirmed in cisplatin resistant human cells in vitro and from patients (Scanlon and KashaniSabet, 1989; Scanlon et al., 1989b,c; Kashani-Sabet et al., 1990a). The H-ras oncogene has also been shown to influence the methionine requirement in H-ras transformed cells (Vanhamme and Szpirer, 1987). These studies further establish the link between methionine/folate metabolisms, DNA repair systems and proto-oncogenes (see Section 2.3). Furthermore, H-ras may also enhance transcriptional activity of c-jun via specific changes in the phosphorylation state of the Jun protein (Bin6truy et al., 1991). Protein kinase C (PKC) has a crucial role in signal transduction for a variety of substrate proteins putatively important for cell differentiation and proliferation (Adunyah et al., 1991). Cells treated by various growth factors or tumor promoters exhibited the

resultant altered PKC activity, leading to the phosphorylation of a number of proteins and a dramatic shift of gene expression (McDonnell et al., 1990). The exact pathways of signal transduction from activation of PKC to genetic regulation in the nucleus are not well understood. However, some portion of the PKC activity is present in the nucleus as well as in the plasma membrane; hence, nuclear proteins involved in transcription, replication and repair may be subjected to modification. Of particular interest was the finding that PKC activated DNA topoisomerase I (Pommier et al., 1990), as well as DNA polymerase /~. Thus, signal transduction from PKC may genetically regulate the expression of DNA polymerase (Tokui et al., 1991). The expression of this gene is also regulated by promoter sequences for a transcription factor that may be a Fos-related protein and that the product of the c-fos gene may repress expression of the DNA polymerase /~ gene (Tokui et al., 1991). The expression of c-myc gene is also regulated by several mechanisms, including transcriptional initiation, transcription elongation, mRNA stability and translation. Transcriptional regulation of c-myc is responsive to many physiologic signals, such as growth factors and to biological agents. Transcription of c-myc starts from at least three initiation sites, with distinct promoters. PI and P2 are the major promoters located upstream of exon I and within exon I, respectively. The transcription of c-myc from P1 and P2 is regulated by a composite of positive and negative elements located both upstream and downstream of these promoters. A negative transcriptional element, which acts on both PI and I)2, has been mapped between -350 and -294 bp relative to the transcriptional starting site, P1. There are two biochemically distinct factors that bind specifically to the 26 bp DNase I footprinted region of the c-myc negative element. One of the factors was identified as the fos/jun (AP-I) complex (Takimoto et al., 1989) and the other as a ubiquitous octamer-binding protein. Interestingly, the two complexes bind highly overlapping sequences. The presence of interdigitating binding sites for distinct factors may confer multiple functions on a single cis-element or allow cooperative interactions between trans-acting factors. Fos/jun (AP-I) and the ubiquitous octamer-binding factor are known to participate in transcriptional activation and/or DNA replication; these studies suggest that repression of c-myc expression may employ components common to transcriptional activation and DNA replication/repair. This has been also noted in cisplatin resistant A2780DDP cells with c-fos/c-myc modulation by cyclosporin A (KashaniSabet et al., 1990b) or by a f o s ribozyme (Scanlon et al., 1991). H-ras has also been shown to enhance the transcriptional activity of c-jun via changes in the phosphorylation state of the c-jun protein (Bin6truy et al., 1991). Platelet-derived growth factor (PDGF) and mitogens stimulate expression of the proto-oncogenes

Cisplatin resistance in human cancers c-myc and c.fos in their respective target cells (Hall and Stiles, 1987). Although much data have accumulated concerning the inducibility of c-myc and c-fos, the precise role of these proto-oncogenes in the proliferative response to growth factors and mitogens is unknown. Agents which activate protein kinase C stimulate c-myc expression, but are only weak mitogens. Down regulation of protein kinase C (by prolonged exposure to a phorbol-based tumor promoter) diminishes the ability of PDGF to induce c-myc, but has no effect on the ability of P D G F to induce growth. Other data tend to dissociate c-los from any direct role in cell growth or differentiation. This second body of data suggests that induction of the c-myc and c-los gene products is neither necessary nor sufficient for mitogen to stimulate cell proliferation. Phorbol esters induce the growth and differentiation of hematopoietic cells. The mechanism by which phorbol esters stimulate gene transcription has recently been shown to involve the activation of specific enhancer sequences including AP-1, AP-2, AP-3 and NF-KB. To activate the AP-1 enhancer sequence, phorbol esters induce an increase in c-jun and c-los transcription. The protein products of these genes, Jun and Fos, contain a leucine repeat region, (leucine zipper), which allows them to dimerize. Control of transcription by the c-jun mRNA also involves an AP-I binding site and autoregulation by a posttranslationally modified Jun protein. In contrast, c-los m R N A synthesis is regulated by multiple enhancer regions. Regulation of AP-1 enhancer activity is extremely complex involving multiple binding proteins and both posttranslational and transcriptional control mechanisms (Adunyah et al., 1991). Epidermal growth factors (EGF) have also been shown to stimulate the expression of the protooncogenes c-myc and c-los in epidermal cells (Bravo et al., 1985). The binding of EGF to the membrane initiates stepwise biochemical changes in calcium fluxes, amino acid uptake and activation of the protein kinases which can make the cells sensitive to cisplatin (Amagase et al., 1989; Christen et al., 1990). In turn, cisplatin-resistant cells have lost this responsiveness to EGF and have altered protein kinase responses. Thus, agents which interact with the signal transduction pathway may have an inhibitory or stimulatory effect depending on the close and exposure time. These agents have been used to modulate cisplatin cytotoxicity with varying degrees of success (see Section 4.2). 3. CLINICAL RESISTANCE 3.1. PATIENTTUMORS Due to a lack of tissue samples and several technical difficulties, only a small number of patient tissues have been analyzed directly for cisplatin resistance. Presently, tumors from fifteen ovarian and colon patients have been analyzed for changes in gene

391

expression that are clinically resistant to cisplatin (Kashani-Sabet et al., 1988, 1990a; Scanlon et al., 1989a,b,c; Scanlon and Kashani-Sabet, 1989a; Kashani-Sabet et al., 1990a). These investigations have overcome the inherent difficulties in conducting in vivo drug resistance studies. These findings in patients corroborate the previous data on human ovarian and colon cell lines/n vitro (Scanlon et ai., 1989b). Human cells that become refractory to cisplatin based chemotherapy have enhanced expression of genes associated with DNA repair. In one study, multiple tumor samples from a patient treated with cisplatin were shown to have increased expression for oncogenes and enzymes associated with DNA synthesis and repair (Kashani-Sabet et al., 1990a). The polymerase chain reaction (PCR) assay has also overcome the problem of the limited number of patient cells that can be assayed to confirm these findings with conventional molecular biology assays. The PCR assay will be discussed in detail in the next section. Heterogeneity of the tumor sample still remains a problem especially if there are only quantitative changes rather than qualitative changes in gene expression. The use of normal tissue as a control is also problematic due to its non-proliferating properties. Finally, it is still difficult to uncouple the effects of tumor progression from drug resistance on gene expression in tumor cells in vivo. However, other mechanisms for cisplatin resistance have not yet been shown in fresh human tissue resistant to cisplatin. In a series of studies an ELISA assay was used to correlate increased cisplatin-DNA adducts in white blood cells of a patient treated with cisplatin or carboplatin with an increased patient response (Reed et al., 1986, 1987, 1988, 1990; Fichtinger-Sehepman et al., 1987; Parker et al., 1991). These studies had several limitations: (i) These were retrospective studies; (ii) the studies had a limited number of patient samples; (iii) adduct formation did not correlate with either the duration of the response or pharmacokinetits; and finally, (iv) specific adduct levels could not predict positive responses. In contrast, the PCR assay has become an indispensable tool in the laboratory, as well as for diagnostics. The use of the PCR assay for detection of drug resistance in patients will be discussed in the next section. 3.2. EARLYDETECTIONOF DRUG RESISTANCE Changes in mRNA for several genes are now known to predict drug resistance in patients. The appropriate screening for these tumor cell markers should be helpful clinically. However, multiple mechanisms of resistance can evolve in established malignancies; thus it would be important to test several genes for changes in expression in drug resistant cells. With the introduction of the Taq polymerase and an automated thermocycler, the PCR assay has become indispensable to most molecular biology laboratories (Guyer and Koshland, 1989; Erlich et al., 1991).

392

K.J. SCANLONet al.

Although a simple assay, PCR use as a diagnostic tool for human diseases prior to these advances was time consuming and tedious (Kashani-Sabet et al., 1988). For cancer diagnosis, the RNA PCR assay can identify which genes are expressed by the use of reverse transcriptase in a modified PCR assay. Studies in this laboratory with tumor samples from 18 patients (Kashani-Sabet et al., 1988; Scanlon and Kashani-Sabet, 1989; Scanlon et al., 1989a,b; Miyachi et al., 1991) have shown the PCR assay to be an effective device in detecting early failure of patients treated with cancer chemotherapeutic agents. This method has confirmed data obtained by conventional molecular biology techniques while circumventing the practical problems of time, sensitivity of detection and the small amount of patient tumor cells. In spite of these studies and others (Gilliland et al., 1990; Veelken et al., 1991), there has been scepticism about whether the PCR assay could quantitate mRNA analogous to Northern blots. A series of controls was introduced to quantitate RNA PCR: (i) Eliminating DNA contamination with DNase or primers that span an intron, (ii) use of an internal control gene that does not change its expression (i.e. phosphoglycerate kinase), (iii) standardizing the amount of RNA to demonstrate linearity for a specific number of PCR cycles, (iv) determining the authenticity of the DNA product by restriction enzyme digestion and (v) use of control cells whose gene expression does not change (Scanlon and KashaniSabet, 1989; Miyachi et al., 1991). Based on these criteria, we have quantitated gene expression in patient tumors for DNA synthesis genes (dihydrofolate reductase, dTMP synthase and DNA polymerase fl) and proto-oncogenes (c-fos, c-H-ras and c-myc). Changes in gene expression have also been shown by the PCR assay for the multidrug resistant phenotype. In a study with 202 patients, tumor samples from 26% of the patients were found to express this drug resistant gene (Noonan et al., 1990). In the future, more patient material will establish the PCR assay as a clinical diagnostic tool for patients with natural resistance or acquired resistance to these agents (Negrin and Blume, 1991). The advantage of measuring gene expression in contrast to gene amplification is that drug resistance in patients may not require gene amplification to achieve clinical resistance (Kashani-Sabet et al,, 1990a). 4. C I R C U M V E N T I N G CISPLATIN RESISTANCE 4.1. PLATINANALOGS Although cisplatin has proven effective against testicular, ovarian and bladder carcinoma, both host toxicity and natural/acquired resistance has limited its usefulness against some of the more common cancers, such as colon, lung and breast carcinomas. Thus, new cisplatin analogs have been actively syn-

thesized and tested alone or in combination with some success (Gill et ai., 1991; McKeage et al., 1991). Currently, the two most promising analogs from tissue culture studies are carboplatin and iproplatin but they may not be as effective as cisplatin against a wide spectrum of tumors (Harstrick et al., 1989). In the last few years, however, several new platin compounds have been synthesized and have shown promise in tissue culture against human carcinomas (Boven et al, 1985; de Graeff et al., 1988; Kuppen et al., 1988; Foster et al., 1990). Some of these agents may not trigger the cascade response induced by cisplatin (see Section 2.4). Tetraplatin (ormaplatin) has been shown to be more effective than cisplatin and less cross-resistant to cisplatin-resistant cells in several human cell lines, including ovarian carcinoma (Behrens et al., 1987; Scanlon et al., 1987, 1989b; Bhuyan et aL, 1991; Perez et al., 1991b); small cell lung carcinoma (Hospers et al., 1988); colon carcinoma and breast carcinoma (Scanlon et al., 1989b). This is due in part to a limited induction of the expression of genes involved in DNA synthesis and repair and a longer (2-6 hr) time required for this response (unpublished data). The d-transisomer of tetraplatin appears to be the more potent component from mouse studies (Bhuyan et al., 1991) and this agent is currently in phase I clinical trials. Oxalatoplatin, (1 R, 2R-1,2 cyclohexadiamine-N,N) oxalato(2-)-O,O')platinum, shows effectiveness that is similar to that of cisplatin against several transplanted murine tumors (Boughattas et al., 1989). Oxalatoplatin, at equally active doses of cisplatin, does not cause renal toxicity and has far less hematologic toxicity (Extra et al., 1990; Caussanei et al., 1990). Phase II clinical trials are underway using higher doses of oxalatoplatin resulting from optimization of circadian rhythm-based strategies and in combination with fluorouracil. Preliminary studies have shown good clinical responses to the 5-fluorouracil and oxalatoplatin combination (Levi et al., 1991). Bis(platinum) complexes [{cisPtCl2(NH3)}2 H2N(CH2)NNH 2 ] are a new series of anticancer agents in which two eis-diamine(platinum) groups are linked by an alkyidiammine of variable lengths. These complexes are potentially tetrafunctional and their interactions with DNA are significantly different from cisplatin (Roberts et al., 1989; Farrell et al., 1990). It appears that bis(platinum) adducts contain a high frequency of structurally novel interstrand cross links formed through binding of the two platinum centers to opposite DNA strands. In addition, bisplatinum complexes do not unwind supercoiled DNA, in a similar manner to cisplatin. Diastereoisomeric compounds [l,2-bis(2-hydroxyphenyl)ethylenediamine] dichloroplatinum(II) complexes, OL-3-PtC12 and meso-3-PtCl: compounds have been evaluated on the hormone-dependent, human MDA-MB231 breast cancer cell line, on the

Cisplatin resistance in human cancers cisplatin-sensitive and -resistant LI210 leukemia cell line, on the cisplatin-resistant human OVCAR3 ovarian cancer cell line and sensitive and resistant Ehrlich ascites tumors of the mouse. DL-3-PtCI2 produces a marked inhibitory effect on all tumor models. The diastereoisomer meso-3-PtC1 e is less active and more toxic than DL-3-PtClz. DL-3-PtC12led to a pronounced inhibition of all cisplatin-resistant tumors. At non-toxic concentrations DL-3-PtCI2 produces cytocidal effects on the NIH :OV-CAR 3 cell line. Due to the low toxicity and lack of cross resistance to cisplatin-resistant cells, DL-3-PtCi 2 is being further tested (Muller et al., 1990). A series of platinum(II) antitumor agents that are different from the classical structure-activity relationships established for platinum complexes have demonstrated activity against murine and human tumor systems cis[Pt(NH3)2(AM)CI] cations, in which AM is a derivative of pyridine, pyrimidine, purine, or aniline. They can inhibit simian virus 40 DNA replication in vitro and inhibit the action of DNA polymerases at individual guanine residues in replication mapping experiments. The platinumtriamine complexes do not produce the type of intrastrand cross-links on DNA that is characteristic of cisplatin and analogs with the general formula cis [Pt(amine)2X2]. These results indicate that cis[Pt(NH3)2(AMCi] + cations form monofunctional adducts on DNA. This is further supported by nuclear magnetic resonance spectroscopic and enzymatic digestion analyses of the products of the reactions of these triamine complexes with d(GpG) and dG, which also reveal monofunctional binding. The fact that trans[Pt(NH3)(4-Br-pyridine)Cl2] displays no anticancer activity, however, indicates that its formation from cis[Pt(NH3)2(4-Br-pyridine)Cl] + is not a significant component of the mechanisms of action of this platinum-triamine complex. Taken together, these findings again indicate that the cytotoxicity of cis[Pt(NH3)2(AM)C1] + complexes most likely arises from the formation of monofunctional adducts. The DNA binding properties associated with this new class of antitumor agents suggest that they may display an activity profile different from that of cisplatin and related analogs (Hollis et al., 1991). 4.2. MODULATORS OF CISPLATIN RESISTANCE As we elucidate the mechanisms of action for cisplatin, rational strategies will emerge to exploit or circumvent these phenotypic properties of cisplatinresistant cells. Several classes of biochemical modulators have been identified that modulate cisplatin acquired resistance by: (i) Interfering with the signal transduction pathway; (ii) inhibiting DNA repair; or (iii) acting through schedule-dependent cytotoxic combinations. Docosahexaenoic acid (DCHA) has been shown to change the lipid membrane composition of small cell lung carcinoma cells and enhance cisplatin activity in

393

the cisplatin-resistant cells (Timmer-Bosscha et al., 1989). This 3-fold reduction in cisplatin cytotoxicity in the resistant cells was not associated with drug transport or changes in the GSH levels, but was associated with a decreased ability to remove cisplatin DNA adducts. These results suggest that DCHA may modify the signal transduction pathway and thus limit the cells' ability to repair their DNA. Both the calcium channel blocker, nifedipine (Onoda et aL, 1986) and the calmodulin antagonist, naphthalenesulfonamide (Hidaka and Tanaka, 1983; Kikuchi et al, 1987) have been shown to have synergistic toxicity when either are combined with cisplatin. Another inhibitor of calmodulin, trifluoperazine, has also been shown to modulate cisplatin cytotoxicity about 2-fold in human ovarian carcinoma cells (Charp and Regan, 1985; Perez et al., 1990). In addition, calmodulin antagonists have potentiated DNA damage from irradiation and u.v. light (Rosenthal and Hait, 1988). Conversely, calcium-free media or EGTA caused calcium depletion in the cells and protects against the cytotoxic action of cancer chemotherapeutic agents, hyperthermia and gamma radiation (Bertrand et al., 1991). Protein kinases have been shown to play an important role in the signal transduction pathway, which involves membrane ion fluxes, modulation of AP responsive elements and direct effects on DNA synthesis enzymes (see Section 2.4). Thus, it is not surprising that heavy metals (Zn 2+, Cd 2+ and Hg 2÷) have been shown to inhibit TPA binding and alter protein kinase activity (Speizer et al., 1989). The protein kinase A in cisplatin-resistant cells is not responsive to forskolin in contrast to sensitive cells in which forskolin enhances cisplatin cytotoxicity (Christen et al., 1990; Mann et al., 1991). TPA, however, is a dynamic modulator depending on the dose and length of exposure. TPA can act directly on responsive elements of the DNA synthesis genes such as topoisomerase I (Pommier et al., 1990); the repair gene, gadd45 (Papathanasiou et al., 1991); the metallothionein gene (Hamer, 1986); and the c-fos oncogene and c-jun/AP1 activation (Huang et al., 1991; see Section 2.4). If cells have a short term exposure to TPA a transient level of increased gene expression is noted. In contrast, long term exposure to TPA decreased PKC activity and gene expression in the signal transduction pathway. TPA has been shown to transiently modulate cisplatin toxicity with human ovarian carcinoma cells (2.5-fold) (Isonishi et al., 1990) or cause a 9-fold increase in cisplatin cytotoxicity in HeLa cells with 24 hr exposure of TPA (Basu et al., 1990). Cyclohexamide did not influence TPA activity; similarly cyclohexamide does not effect the dimer formation of Fos/Jun. Inhibitors of protein kinase C; lyngbyatoxin A, quercetin and staurosporine (Hofmann et aL, 1988; Basu et al., 1991) also effect cisplatin cytotoxicity. The protein kinase inhibitor, H-7, has been shown to inhibit c-fos mRNA accumulation and the transcriptional

394

K.J. SCANLONet al.

activation of c-fos (Shibanuma et al., 1987). This would suggest that H-7 could potentiate cisplatin cytotoxicity. Glucocorticosteroids may also have activity for modulating cisplatin toxicity. CSA, used primarily as an immunosuppressant in transplantation, has been shown to reverse resistance to a variety of chemotherapeutic agents, including etoposide, adriamycin, vincristine, daunorubicin and cisplatin (Osieka et al., 1986; Twentyrnan et al., 1987; Sweet et al., 1989; Kashani-Sabet et aL, 1990b). It is intriguing that CSA can restore sensitivity to cisplatin-resistant cells because of the potential clinical implications. It has been suggested that CSA interacts either with plasma membrane potentials (Damjanovich et al., 1986), calcium--calmodulin pathways (Colombani et al., 1985), or acts by suppressing nuclear oncogene expression (Kashani-Sabet et al., 1990b; Scanlon et al., 1990b). In lymphocytes, CSA inhibits PHA-stimulated gene expression of growth factors (IL-2) and may interfere with Ca2+-depen dent pathways that regulate several transcription factors (Baldari et al., 1991). CSA forms an inhibitory complex with peptidyl-prolyl cis-trans isomerase which may block transcription factors (Gunter et al., 1989; Mattila et al., 1990; Flanagan et al., 1991). How this works in non-T cells has yet to be determined. These studies further support the role of the signal transduction pathway in cisplatin resistance (Section

2.4). The role of GSH in resistance to cisplatin has been studied in several human cell lines (see Section 2.2) and data remains inconclusive. However, pretreatment of cisplatin resistant cells with BSO has been shown to reduce the repair capacity of the cells and enhance cisplatin cytotoxicity 2- to 4-fold (Hamilton et al., 1985; Fojo et al., 1987; Lai et aL, 1989; Meijer et al., 1990). BSO treatment may play a role in eliminating GSH from the cytosol and enhancing the formation of DNA-platin adducts. This, however, has not been seen with all cisplatin-resistant cell lines and high levels of BSO are sometimes required for modulation of GSH levels to effect cisplatin cytotoxicity. In contrast to cisplatin, it appears that tetraplatin cytotoxicity is modulated more by BSO in some human ovarian tumors (Mistry et al., 1991). These strategies have only partially reversed cisplatin resistance in contrast to CSA or ribozymes. Aphidicolin (APC), an inhibitor of DNA polymerase ~ and fl, can dramatically reduce the level of DNA synthesis (Sheaff et aL, 1991) if the cells are continuously exposed to APC (Sorscher and Cordeiro-Stone, 1991). However, APC is a reversible inhibitor and once removed from the media DNA strand growth rapidly resumes to near normal levels. In spite of these limitations several studies have suggested that APC increases cisplatin cytotoxicity (Masuda et aL, 1988; 1990) while other studies have shown no activity with APC (Fram et al., 1987). The combination of BSO and APC has been shown to be additive with a 4-fold increase in cisplatin cytotox-

icity in human ovarian cells (Lai et al., 1989). These results could be reversed with the addition of GSH and suggest that GSH depletion could potentiate cisplatin cytotoxicity. BSO has also been shown to restore radiation sensitivity to melphalan-resistant A2780 cells. However, BSO treatment of cisplatinresistant A2780 cells was less effective and the authors suggested that elevated levels of GSH may not be the only mechanism of cisplatin resistance (Louie et aL, 1985). A number of studies has suggested that single dose irradiation can produce a clone of tumor cells that survive the treatment and are transiently resistant to cancer chemotherapeutic agents (Schwartz et aL, 1988; Sharma and Schimke, 1989; Hopwood and Moulder, 1989; Hopwood et al., 1990). In contrast, multiple exposures of tumor cells to X-irradiation treatment produce clones collaterally sensitive to chemotherapeutic agents (Bedford et al., 1987; Hill et al., 1988; Mivechi et al., 1990; Dempke et al., 1991) or collaterally resistant to other agents (Hill et al., 1988, 1990; De Pooter et al., 1991). These apparently contradictory findings may be explained by understanding the signal transduction pathway (Fig. 1). Ionizing radiation has been shown to transiently activate the fos/jun complex in several human cell lines in a dose-dependent fashion (Stein et al., 1989; Sherman et al., 1990). This would stimulate APresponsive elements in genes associated with DNA synthesis and repair and cause transient collateral resistance to cancer chemotherapeutic agents (Scanlon et al., 1991). Conversely, pretreatment of cells with cisplatin can sensitize the cells to Xirradiation (Begg et al., 1986). In contrast to these findings, repeated treatment of human cells with X-irradiation suppresses gene expression for DNA repair and the cells become collaterally sensitive to anti-metabolites or hyperthermia (Hill and Bellamy, 1984; Mivechi et al., 1990; Dempke et al., 1991). There could be several explanations for this decrease in DNA repair activity (Dewit, 1987), however, it would appear that the final effect of this decreased enzyme activity may be due to repression of the promoter elements of the DNA polymerase fl gene. It would be interesting to have these observations confirmed clinically (Ensley et al., 1984; Ervin et al., 1987). In order to improve treatment of tumors, a trimodality therapy has been tested experimentally with radiation, hyperthermia and cisplatin in a fibrosarcoma (Herman and Teicher, 1988). It appears that the sequencing of cisplatin first, hyperthermia second and then X-irradiation results in the optimal tumor cytotoxicity. Similar results have also been shown with hyperthermia and X-irradiation for leukemia cells (Mivechi et al., 1990), mouse tumors (Wallner et al., 1986; Mansouri et al., 1989) and fibrosarcomas (Herman et al., 1990). A platin analog, platinum rhodamine 123, has been shown to have cytotoxic activity and is an effective radiosensitizing agent both

Cisplatin resistance in human cancers in vivo and in vitro (Herman et al., 1988). In addition, this compound has minimal whole animal toxicity and may have a different mechanism of action at DNA than cisplatin.

4.3. TI-mfos ONCOGENEIN DRUG RESISTANCE Previous studies in a variety of cisplatin-resistant cell lines have suggested the importance of the los oncogene in drug resistance (reviewed by Scanlon et al., 1989b). Cisplatin resistant cell lines both in culture (Scanlon et al., 1989c; Kashani-Sabet, et al., 1990b) and from patient samples (Kashani-Sabet et al., 1990a) have been shown to exhibit elevated c-los gene expression, concomitant with over expression of several enzymes involved in DNA synthesis and repair processes (e.g. dTMP synthase and DNA polymerase fl). The hypothesis thus has been developed that the Fos protein functions in drugresistance by activating transcription of genes with a direct role in repair synthesis of DNA. This was supported by experiments demonstrating that administration of cisplatin leads to a cascade of gene expression, beginning with that of c-los (an earlyresponse gene) (Greenberg and Ziff, 1984) within 1 hr of drug treatment followed by that of dTMP synthase (at 2 hr) and DNA polymerase fl (at 3 hr), presumably effecting the cellular response to cisplatin damage. Studies by other laboratories have defined the two important functions of the Fos protein: Transcriptional activation and DNA synthesis (Cohen and Curran, 1989). For instance, Fos combines with the c-jun oncogene product to facilitate transcription of several genes (Rauscher et al., 1988; Kouzarides and Ziff, 1989; Chiu et al., 1988; Lee et al., 1987b). Moreover, studies using fos antisense RNA have implicatedfos in DNA synthesis and cellular proliferation (Holt et al., 1986; Riabowol et al., 1988; Ledwith et al., 1990). However, it was not clear how the gene-regulating function of Fos contributed to DNA synthesis. Our studies suggested that this was achieved by activation of enzymes with a direct role in DNA synthesis. In order to strengthen this hypothesis, we designed a ribozyme to cleave c-fos mRNA and investigate its effects on signal transduction and cisplatin resistance. The discovery of catalytic RNAs (or ribozymes) has suggested these agents as a new modulator of gene expression. Ribozymes (reviewed by Cech and Bass, 1986) act as site-specific RNAses whose specificity is determined by sequences flanking the catalytic moiety which are complementary to the target RNA. A particular sub-group, known as hammerhead ribozymes, consists of three stems and a catalytic center (Forster and Symons, 1987) and cleaves 3' to any G U N site (N being A, C, or U) (Haseloff and Gerlach, 1988). DNA encoding a hammerhead anti-fos ribozyme was cloned into the dexamethasone-inducible

395

pMAMneo plasmid and transfected into cisplatinresistant A2780DDP cells (Scanlon et al., 1991; Funato et al., 1992). The transformants demonstrated dramatic changes in morphology to resemble the parental drug-sensitive cell line. In addition, dexamethasone induction of the dbozyme completely reversed resistance to cisplatin in A2780DDP cells. Moreover, ribozyme induction and its effects on c-fos resulted in decreased expression of DNA synthesis enzymes, including dTMP synthase, DNA polymerase ~ and topoisomerase I, genes shown to be overexpressed in A2780DDP cells, hMTIIA, which has also been implicated in cisplatin resistance (Kelley et aL, 1988), is another known target of Fos/Jun action (Lee et aL, 1987a). Interestingly, hMTIIA expression is also lowered afterfos ribozyme activation in A2780DDP cells. These results, taken together, suggest that Fos is capable of controlling the cellular response to cisplatin-induced DNA damage by activation of several genes which have been separately implicated in cisplatin resistance. Importantly, dTMP synthase (Takeishi et aL, 1989) and topoisomerase I (Kunze et al., 1990) both contain AP-I binding sequences in their 5' regulatory sequences. DNA polymerase contains a ras-responsive element (Kedar et aL, 1990) andros has been shown to operate downstream of H-ras in signal transduction pathways (Ledwith et al., 1990). The fact that overexpression of several enzymes including dTMP synthase, DNA polymerase ~ and topoisomerase I is also implicated in resistance to 5-fluorouracil (Scanlon and Kasbani-Sabet, 1988), azidothymidine (Vazquez-Padua et aL, 1990) and camptothecin (Giovanella et al., 1989), respectively, broadens the spectrum of resistance of Fos action. The observation that both the GST gene (Friling et aL, 1992) and the MDR gene also contains an AP-I binding site (Teeter et aL, 1991) further reinforces the hypothesis that Fos may act as a master switch in the coordination of drug resistance (Scanlon and Kashani-Sabet, 1989), DNA repair and detoxification processes. It may require only a small modulation of Fos by the los ribozyme to have perterbations of the Fos/Jun complex which would either activate the wrong genes or inadequately turn on the correct genes. The aforementioned studies also introduce ribozymes as agents which can overcome drug resistance. That ribozymes can effectively curtail oncogene expression suggests that they may have roles either in reversing drug resistance or in acting as suppressors of tumor cell growth (Kasbani-Sabet et al., 1992). Ribozymes may have a useful advantage over antisense RNA for fos gene expression because of the short half-life of c-fos mRNA (Nishikura and Murray, 1987). These possibilities need to be further explored as a possibility for the development of tumor-selective chemotherapy. Recently a study (Isonishi et al., 1991) demonstrated that H-ras transfection into NIH3T3 cells resulted in an 8.1-fold resistance to cisplatin. The

396

K.J. SCANLONet al.

mechanisms of cisplatin resistance were only partially accounted for by a decrease in cisplatin transport (40%), a 17% decrease in cisplatin-DNA adduct formation, a 2-fold increase in metallothioneins and no change in glutathione content. There was no characterization of other H-ras responsive genes and/or c-fos oncogene expression. Surprisingly when the c-los oncogene was transfected into another mouse cell line, there was no change in cisplatin cytotoxicity. The 5-fold increase in the Fos protein had no effect on any f o s responsive genes. This was confusing since several genes with AP-I responsive elements that can modulate cisplatin cytotoxicity apparently did not respond to this increase in Fos (i.e. metallothionein, DNA polymerase fl, dTMP synthase), although they were not measured.

5. EXPLOITING THE MECHANISMS OF CISPLATIN RESISTANCE 5.1. THYMIDINESALVAGEPATHWAY Understanding the biochemical events that result in cisplatin resistance will likely provide targets for the development of schemes to circumvent resistance. Both the biochemical and tissue culture studies support the hypothesis that resistance to cisplatin and nucleotide analogs (fluorouracil, AZT) is a phenomenon related to the cells' capacity to repair DNA. Because DNA repair requires deoxyribonucleotides, the increased activity of deoxyribonucleotide metabolism was a likely component of enhanced DNA repair capacity (Scanlon et al., 1989b). In addition to expanded dNTP metabolism, it is likely that resistant cells have increased capacity for adduct excision, repair synthesis and ligation. In human colon carcinoma cells resistant to cisplatin (HCT8DDP), increased gene expression for DNA pol fl was observed (Scanlon et al., 1989c). These findings are similar to those obtained from several other cell lines resistant to cisplatin: mouse leukemia P388 cells, human leukemia K562 cells, human leukemia CEM cells and human ovarian carcinoma cells (Scanlon et al., 1989b). Based on these results, suicide substrates which are incorporated by DNA pol fl were incubated with cisplatin-resistant cells. Specifically, AZT, a derivative of thymidine metabolized via the thymidine metabolic pathway, as the triphosphate AZTTP acts as a suicide substrate causing chain termination (Scanlon et al., 1989c). DNA is degraded more rapidly in cisplatin-resistant cells than by the sensitive cells to remove the cisplatin-DNA adducts (Scanlon et al., 1990a). Repair gaps generated following cisplatin-DNA adduct removal would then provide sites for incorporation of suicide substrates, i.e. AZTTP. If cells were treated with cisplatin and then AZT, cisplatin-resistant cells are at a distinct disadvantage. HCT8DDP cells are prepared to repair cisplatin exposure effects due to the enhanced capacity to synthesize and repair cis-

platin-induced DNA damage, in contrast to the HCT8S cells. AZT may have additional advantages over other DNA synthesis inhibitors, such as aphidicolin. AZT is first phosphorylated by TK, which we have shown is overexpressed in cisplatin-resistant cells. Further, AZTTP can act as an inhibitor of DNA polymerase fl (similarly overexpressed in resistant cells), as well as DNA polymerase ct. AZT has advantages over dideoxythymidine (ddT), as ddT is poorly transported into human ovarian cells (Katz et al., 1990; Perez et al., 1990). Although A2780DDP cells have been shown to have enhanced machinery to repair cisplatin-induced DNA damage, it is still surprising that these cells can turn over 40% of the [t4C]-labeled DNA in the first 3 hr after cisplatin treatment (Scanlon et al., 1990a). The high turn over of labeled DNA in A2780DDP cells could have been due to this repair patch and/or the presence of thymidine rich patches. Consequently, the large quantity of degraded (i.e. decoy) DNA could have protected the A2780DDP cells from cisplatin. This concept has also been termed 'the cells tolerance to cisplatin'. Other studies suggest a similar role for metallothionein or glutathionine (Andrews and Howell, 1990; Scanlon et al., 1989b). An enhanced expression of genes necessary for DNA synthesis and repair has also been demonstrated in patients who failed cisplatin based chemotherapy. Thus, mechanisms for degrading DNA to protect drug resistant cells may have clinical relevance and could be easily exploited by this currently available nucleoside analog. 5.2. NUCLEOSIDEANALOGS Several DNA chain terminating nucleoside analogs have been developed in the past few years which have significant activity as antiviral agents. These analogs are activated by cellular enzymes to the triphosphate level. As the triphosphorylated derivatives, the modified nucleosides can serve as suicide substrates for DNA polymerases. Thus, increases in nucleoside metabolism in cisplatin resistant cells could potentially result in increased activation of these suicide substrates. Ganciclovir is an example used to inhibit DNA polymerase ct (Scanlon et al., 1989b). Increases in DNA polymerase levels in cisplatin-resistant cells could result in increased utilization of the drug and thus increased levels of truncated DNA molecules. Dideoxycytidine (ddC) is another nucleoside that has been studied in human ovarian carcinoma cells. A2780S cells have an ECs0 of 1.5 #M for continuous exposure to ddC. However, a suboptimal dose of cisplatin (1/~M) for l hr makes ddC eight times more potent in A2780S cells. Similar results have been obtained in human colon carcinoma HCT8 cells. In cisplatin resistant A2780DDP cells, cisplatin has only a small modulating effect on ddC, but in human colon carcinoma cisplatin-resistant cells, small doses of cisplatin are potent modulators of ddC (15-fold).

Cisplatin resistance in human cancers Additional studies to measure transport, metabolism and incorporation into DNA by sensitive and resistant cells are currently in progress. Studies with other nucleoside analogs which differ in their specificity as inhibitors of human polymerases are currently being investigated. These analogs offer another promising approach. By sequencing the order of administration of these analogs with cisplatin, synergistic combinations may lead to results similar to those described in the studies with fluoropyrimidines. The cytotoxic action of 1-fl-D-arabinofuranosylcytosine (araC, cytarabine) is complex and araCTP may exert its effect on multiple enzymes rather than just on DNA polymerase fl (Zittoun et al., 1989). Thus, demonstrating the interaction of araC with cisplatin has been confusing at times. Drug synergy has been reported for combinations of cytarabine and cisplatin in human carcinoma cells (Drewinko et al., 1976; Bergerat et al., 1981; Kingston et al., 1989) but this was not reproduced in human ovarian carcinoma 2008 cells (Howell and Gill, 1986) or in other laboratories (Fram et al., 1987). Kern and associates (1988) studied 55 human tumors for araC cytotoxicity and then optimized a 2 hr pretreatment of araC with cisplatin. Sequencing araC prior to cisplatin was most effective against melanoma and breast carcinomas. The reverse sequence decreased the synergism between araC and cisplatin. This observation was also confirmed by human colon cells in culture which showed a dependence on both the concentration and the exposure time for cisplatin to optimize the cytotoxicity of cisplatin with araC (Ozawa et al., 1989). Swinnen et al. (1989) expanded these studies of araC and cisplatin to include hydroxyurea (HU) in human colon carcinoma HT29 cells. This combination appeared to be effective due to the impaired repair capacity of HT29 cells. Carboplatin was substituted for cisplatin and these studies confirmed the clinical potential of this drug combination (Swinnen et al., 1991). This protocol was tested clinically and in the first of these studies 21 of 32 patients with previous treatment had a partial response. Nine of 32 patients and 3 of 8 patients with previous cisplatin treatment had significant responses (Albain et al., 1990). The combination of cisplatin and cyclophosphamide (CTX) has been well established in the treatment of ovarian carcinoma. In recent years, several clinical trials have been conducted comparing the efficacy of carboplatin with that of cisplatin given alone and in combination with CTX (Teeling and Carney, 1990) to patients with ovarian carcinoma. The 2-nitroimidazole radiosensitizer, etanidazole (ETA), is also a hypoxic-cell-selective cytotoxic agent (Teicher et al., 1991a) and a chemosensitizer or modulator of some antitumor drugs. Although the precise mechanism by which ETA acts as a chemosensitizer is not known, it appears to occur at the cellular level and may involve the interaction of a metabolite of the 2-nitroimidazole with the alkylating agent in the vicinity of DNA. Recent studies suggest

397

that the addition of ETA to clinical protocols may be beneficial for platin/CTX based chemotherapy. In a clinical setting, cisplatin and etoposide have been shown to act synergestically, against lymphomas, small cell lung carcinoma, non-small cell lung carcinoma and testicular cancer (Hainsworth et al., 1985; Evans et al., 1985) and in animal models (Schabel et aL, 1979), while with cell lines in culture, it has not been possible to reproduce these findings consistently (Tsai et al., 1989). 5.3. PYRIMIDINEPATHWAY Cisplatin and 5-fluorouracil are widely used antineoplastic agents and each has a broad spectrum of antitumor activity. Studies both in vitro and in tumor bearing animals have shown marked therapeutic synergy for these drugs (Schabel et al., 1979; Scanlon et al., 1986). In addition, the combination of these drugs has been very effective for the treatment of head and neck cancer, cervix and upper digestive tract cancers (Etienne et al., 1991; Decker et al., 1983; Khansur and Kennedy, 1991). Because there have been conflicting studies, new types of protocols and schedule combinations with optimal dose sequences and exposure times for this drug regimen are necessary in order to accentuate the biochemical events that support this potentiation (Rowland et al., 1986; Galligioni et al., 1987; Paredes et al., 1988; Recondo et al., 1989; Schilsky et al., 1990; Kemeny et al., 1990). In vitro tissue culture studies (Rosowsky et al., 1987; Newman et al., 1988) have shown that carcinomas which are resistant to cisplatin are collaterally resistant to anti-metabolites because of enhanced dTMP synthase activity. The clinical protocols with cisplatin and 5-fluorouracil may have limited effectiveness in patients because of these changes in the dTMP synthase cycle.

6. CONCLUSIONS In the treatment of cancer there has been no clinical breakthrough to rival combination chemotherapy in the early 1960s. There have also been no major drug advances since the anthracyclines, platin analogs and the podophyllotoxins of the 1970s. The problem of drug resistant human tumors continues to be a challenge, but is slowly yielding to intense scrutiny as we gain a better understanding of the molecular basis of drug resistance in human tumors. Human cell lines have been shown to be good model systems to predict cisplatin resistance in patient carcinomas. Cisplatin initiates a cascade in the cell to protect chromosomes from the toxic effects of the drug (i.e. resistance). Cells respond to cisplatin with a transient response of gene expression to protect the cellular DNA. Repeated challenges of cisplatin established a cell line with stable resistance which has elevated levels of gene expression associated with

398

K. J. SCANLONet al.

oncology: the interdependence of bench and bedside. Cancer Res. (Ippl). 51: 5060s-5064s. BAKER, S. J., MARKOWITZ, S., FEARON, E. R., WILI.SON, J. K. V. and VOGELSTEIN,B. (1990) Suppression of human coloreetal carcinoma cell growth by wild-type p53. Science 249: 912-915. BALDARI,C. T., MACCHIA,G., HEGUY, A., MELLI, M. and "I~LFOgD, J. L. (1991) Cyclosporin A blocks calciumdependent pathways of gene activation. J. biol. Chem. 266: 19103-19108. BASU,A., TEICr~R, B. A. and LAZO,J. S. (1990) Involvement of protein kinase C in phorbol ester-induced sensitization of HeLa cells to cis-diaminedichloroplatinum(II). Y. biol. Chem. 265: 8451-8457. BASU, A., KOZlKOWSKI,A. P., SATO, K. and LAZO, J. S. (1991) Cellular sensitization to cis-diaminedichloroplatinum(II) by novel analogs of the protein kinase C activator lyngbyatoxin A. Cancer Res. 51:2511-2514. BEDFORD, P., SHELLARD,S. A., WALKER,M. C., WI-IELAN, Acknowledgements--Research from the author's laboratory R. D. H., MASTERS,J. R. W. and HILL, B. T. (1987) cited in this review was supported by the National Cancer Differential expression of collateral sensitivity or resistInstitute (CA50618) and grants from the American Cancer ance to cisplatin in human bladder carcinoma cell lines Society (CH265G and CH483B). We would like to thank pre-exposed in vitro to either X-irradiation or cisplatin. Dr. B. Hill for her valuable discussions of the manuscript. Int. J. Cancer 40: 681-686. We would like to thank Ms Carol Polchow for preparing the BEDFORD, P., FICHTINGER-SCHEPMAN,A. M. J., SHELLARD, manuscript. S. A., WALKER,M. C., MASTERS,J. R. W. and HILL, B. T. (1988) Differential repair of platinum-DNA adducts in human bladder and testicular tumor continuous cell lines. Cancer Res. 40: 3019-3024. BEGG,A. C., VANDERKOLK,P. J., DEWIT,L. and BARTELINK, REFERENCES H. (1986) Radiosensitization by cisplatin of RIFI tumor cells in vitro. Int. J. Radiat. Biol. 50: 871-884. AaATE, C., LUK, D. and CURRAN, T. (1991) Transcriptional BEHRENS,B. C., HAMILTON,T. C., MASUDA,H., GROTZINGER, regulation by Fos and Jun in vitro: Interaction among K. R., WHANG-I~NG, J., LOUIE K. G., KNUTSEN, T., multiple activator and regulatory domains. Mol. Cell. McKoY, W. M., YOUNG, R. C. and OZOLS, R. F. (1987) Biol. 11: 3624-3632. Characterization of a cis-diaminedichloroplatinum(lI)ADUNYAH, S. E., UNLaP, T. M., WAGNER, F. and KRAFT, resistant human ovarian cancer cell line and its use A. S. (1991) Regulation of c-jun expression and AP-1 in evaluation of platinum analogs. Cancer Res. 47: enhancer activity by granulocyte-macrophage colony414-418. stimulating factor. J. biol. Chem. 266" 5670-5675. ALBAIN, K. S., SWINNEN, L. J., ERICKSON, L. C., STIFF, BERGERAT,J. P., DREWINKO,B., CORRY,P., BARLOGIE,B. and He, D. H. (1981) Synergistic lethal effect of cis-dichloroP. J. and FISHER, R. I. (1990) Cisplatin preceded by diamineplatinum and l-~-D-arabinofuranosylcytosine. concurrent cytarabine and hydroxyurea: A pilot study Cancer Res. 41: 25-30. based on an in vitro model. Cancer Chemother. Pharmac. 27: 33-40. BERNAL, S. D., SPEAK,J. A., BOEHEIM,K., DREYFUSS,A. I., AMAGASE,H., KAKIMOTO,M., HASHIMOTO,K., FUWA,T. and WRIGHT,J. E., TEICHER,B. A., RosowsKY, A., TSAO,S.-W. TSUKAGOSHI,S. (1989) Epidermal growth factor receptorand WONG, Y.-C. (1990) Reduced membrane protein mediated selective cytotoxicity of antitumor agents associated with resistance of human squamous carcinoma toward human xenografts and murine syngeneic solid cells to methotrexate and cis-platinum. Mol. Cell. tumors. Jpn. J. Cancer Rcs. 80: 670-678. Biochem. 95: 61-70. ANDREWS,P. A. and HOWELL,S. B. (1990) Cellular pharmaBERTINO,J. R. (1988) Review of mechanisms of cell resistcology of cisplatin: Perspectives on mechanisms of acance to therapy in lung cancer. Antibiotica Chemother. 41: quired resistance. Cancer Cells 2: 35-43. 29-34. ANDREWS,P. A. and JONES,J. A. (1991) Characterization of BERTRAND,R., KERRIGAN,D., SARANG,M. and POMMIER,Y. binding proteins from ovarian carcinoma and kidney (1991) Cell death induced by topoisomerase inhibitors. tubule cells that are specific for cisplatin modified DNA. Biochem. Pharmac. 42: 77-85. Cancer Commun. 3: 1-10. BHUYAN,B. K., FOLZ, S. J., DEZWAAN,J., NORTHCOTT,S. E., ANDREWS,P. A., MURPHY,M. P. and HOWELL,S. B. (1985) ALBERTS,D. S., GARCIA,D., WALLACE,T. L. and LI, L. H. Differential potentiation of alkylating and platinating (1991) Cytotoxicity of tetraplatin and cisplatin for human agent cytotoxicity in human ovarian carcinoma cells by and rodent cell lines cultured as monolayers and multicelglutathione depletion. Cancer Res. 45: 6250-6253. lular spheroids. Cancer Commun. 3: 53-59. ANDREWS,P. A., MURPHY,M. P. and HOWELL,S. B. (1987) BINf~TRUY, B., SMEAL, T. and KARIN, M. (1991) Ha-Ras Metallothionein-mediated cisplatin resistance in human augments c-Jun activity and stimulates phosphorylation ovarian carcinoma cells. Cancer Chemother. Pharmac. 19: of its activation domain. Nature 351: 122-127. 149-154. BOUGHATTAS,N. A., LEVI,F., FOURNIER,C., LEMAIGRE,G., ANDREWS,P. A., VELURY,S., MANN, S. C. and HOWELL, S. B. ROULON,A., HECQUET,B., MATHE,G. and REINBERG,A. (1988) cis-Diaminedichloroplatinum(II) accumulation in (1989) Circadian rhythm in toxicities and tissue uptake of sensitive and resistant human ovarian carcinoma cells. 1,2-diamino-cyclohexane(trans-1)oxalatoplatinum(II) in Cancer Res. 48: 68-73. mice. Cancer Res. 49: 3362-3368. ANDREWS, P. A., JONES,J. A., VARKi,N. M. and HOWELL, BOVEN, E., VAN DER VDGH, W. J. F., NAUTA, M. M., S. B. (1990) Rapid emergence of acquired cis-diamineSCHLUPER, H. M. M. and PINEOO,H. M. (1985) Comparadichloroplatinum(II) resistance in an in rive model of tive activity and distribution studies on five platinum human ovarian carcinoma. Cancer Commun. 2: 93-100. analogs in nude mice bearing human ovarian carcinoma ANTMAN, K. H. (1991) New developments in clinical xenografts. Cancer Res. 45: 86-90. D N A repair enzymes. The nuclear oncogene, fos, an early response gene, can be suppressed with an anti-fos ribozyme that reverts cisplatin-resistant cells to the sensitive phenotype. Clinically, cyclosporin, A Z T and other nucleoside analogs are showing promising results in reversing cisplatin resistance. Future clinical studies should resolve these new strategies. Early detection of drug resistance has moved in the last few years with several clinical studies toward confirming earlier studies on the effectiveness of the P C R assay. Thus, advances in understanding the molecular biology of cancer will hopefully produce diagnostic and therapeutic advances in the treatment of cancer in the 1990s.

Cisplatin resistance in human cancers BRAVO, R., BURCKHARDT,J., CURRAN,T. and MULLER, R. (1985) Stimulation and inhibition of growth by EGF in different A431 cell clones is accompanied by the rapid induction of c-fos and c-myc proto-oncogenes. EMBO J. 4: 1193-1197. BUNCO, M., FUJIWARA,Y., KASAHARA,K., NAKAGAWA,K., OHE, Y., SASAKI, Y., IRINO, S. and SAIJO, N. (1990) Decreased accumulation as a mechanism of resistance to cis-diaminedichloroplatinum(II) in human non-small cell lung cancer cell lines: Relation to DNA damage and repair. Cancer Res. 50: 2549-2553. BuscHr.g, M., R~HMSOORF,H. J., LITFIN,M., K~IN, M. and HERRLICH,P. (1988) Activation of the c-fos gene by u.v. and phorbol ester: different signal transduction pathways converge to the same enhance element. Oncogene 3: 301-311. CAUSSANEL, J.-P., LEVI, F., BRIENZA, S., MISSET, J.-L., ITZ~KI, M., ADAM, R , MILANO, G., HECOUET,B. and MATh'E, G. (1990) Phase I trial of 5-day continuous venous infusion of oxaliplatin at circadian rhythmmodulated rate compared with constant rate. J. nam. Cancer Inst. 82: 1046-1050. CECIl, T. R. and BASS,B. L. (1986) Biological catalysis by RNA. A. Rev. Biochem. 55: 599-629. CHAO, C. C.-K., LEE, Y.-L. and LIN-CrtAO, S. (1990) Phenotypic reversion of cisplatin resistance in human cells accompanies reduced host cell reactivation of damaged plasmid. Biochem. biophys. Res. Commun. 170: 851-859. CHARP,P. A. and REGAN,J. D. (1985) Inhibition of DNA repair by trifluoperazine. Biochim. biophys. Acta 824: 34-39. CHIN, K. V., UEDA, K., PASTAN,I. and GOTTESMAN,M. N. (1992) Modulation of activity of the promoter of the human MDR-I gene by Ras and p53. Science 255: 459-462. CHIU, R., BOYLE,W. J., MEEK, J., SMEAL,T., HUNTER, T. and KARIN, M. (1988) The c-fos protein interacts with c-jun/AP-1 to stimulate transcription of AP-1 responsive genes. Cell 54: 541-552. CHRISTEN, R. D., HOM, D. K., PORTER, D. C., ANDREWS., P. A., MACLEOD,C. L., HAFSTROM,L. and HOWELL,S. B. (1990) Epidermal growth factor regulates the in vitro sensitivity of human ovarian carcinoma cells to cisplatin. J. clin. Invest. 86: 1632-1640. CHU, G. and CHANG, E. (1988) Xeroderma pigmentosum group E cells lack a nuclear factor that binds to damaged DNA. Science 242: 564-567. CHu, G. and CHANG,E. (1990) Cisplatin-resistant cells express increased levels of a factor that recognizes damaged DNA. Proc. natn. Aead. Sci. U.S.A. 87: 3324-3327. COI-mN, D. R. and CURRAN,T. (1989) The structure and function of thefos proto-oncogene. Oncogenesis 1: 65-88. COLOMBANI,P. M., ROBa, A. and HEss, A. D. (1985) Cyclosporin A binding to calmodulin: A possible site of action on T lymphocytes. Science 228: 337-339.

DAMJANOVICH,S., ASZALOS,A., MULHERN,S., BALAZS,M. and MATYUS,L. (1986) Cytoplasmic membrane potential of mouse lymphocytes is decreased by cyclosporins. Mol. Immunol. 23: 175-180. DECKER,D. A., DRELICHMAN,A., JACOBS,J., HOSCHNER,J., KINZlE, J., LOH, J. J.-K., WEAVER,A. and A L - S ~ , F , M. (1983) Adjuvant chemotherapy with cts-diaminedichloroplatinum II and 120-hr infusion 5-fluorouraeil in stage III and IV squamous cell carcinoma of the head and neck. Cancer 51: 1353-1355, DE GRAEFFA., SLEBOS,R. J. C. and RODE~IS, S. (1988) Resistance to cisplatin and analogs: Mechanisms and potential clinical implications. Cancer Chemother. Pharmac. 22: 325-332. DEMPKE, W. C. M., HONKING,L. K. and HILL, B. T. (1991) Expression of collateral sensitivity to cisplatin, methotrexate and fluorouraeil in a human ovarian carcinoma

399

cell line following exposure to fractionated X-irradiation in vitro. Semin. Oncol., in press. DE POOTER, C. M. J., SCALLET, P. G., ELST, H. J., HUVnRECWrS, J. J., GrmUENS, E. E. O., VAN OOSTEROM, A. T., FICHTINGER-ScHEPMAN,A. M. J. and DE BRUIJN, E. A. (1991) Resistance patterns between cis-diaminedichloroplatinum(II) and ionizing radiation. Cancer Res. 51: 4523--4527. DEVITA, V. T., JR (1983) The relationship between tumor mass and resistance to chemotherapy. Cancer 51: 1209-1220. DEWlT, L. (1987) Combined treatment of radiation and cis-diaminedichloroplatinum (II): A review of experimental and clinical data. Int. J. Radiat. Oncol. Biol. Phys. 13: 403-426. DIJT, F. J., FICHTINGER-SCHEPMAN,A. M. J., BERENDS,F. and REEDUK, J. (1988) Formation and repair of cisplatin-induced adducts to DNA in cultured normal and repair-deficient human fibroblasts. Cancer Res. 48: 6058--6062. DONAHUE,B. A., AUGOT,M., BELLON,S. F., TREIBER,D. K., TONEY,J. H., LIPPARD,S. J. and ESSlGMANN,J. M. (1990) Characterization of a DNA damage-recognition protein from mammalian cells that binds specifically to intrastrand d(GpG) and d(ApG) DNA adducts of the anticancer drug cisplatin. Biochemistry 29: 5872-5880. DREWlNKO, B., GREEN, C. and Loo, T. L. (1976) Combination chemotherapy in vitro with cis-dichlorodiamineplatinum(II). Cancer Treat. Rep. 60: 1619-1625. DUCORE,J. M., ERICKSON,L. C., ZWELLING,L. A., LAURENT, G. and KOHN,K. W. (1982) Comparative studies of DNA cross-linking and cytotoxicity in Burkitt's lymphoma cell lines treated with cis-diaminodichloroplatinum(II) and L-phenylalanine mustard. Cancer Res. 42: 897-902. EASTMAN, A. (1986) Reevaluation of interaction of cis-dichloro(ethylenediamine)platinum(II) with DNA. Biochemistry 25: 3912-3915. EASTMAN, A. (1991) Analysis and quantitation of the DNA damage produced in cells by the cisplatin analog cis-[3H]dichloro(ethylenediamine)platinum(II). Anal. Biochem. 197: 311-315. EASTMAN, A. and SCHULTE, N, (1988) Enhanced DNA repair as a mechanism of resistance to cis-diaminedichloroplatinum(II). Biochemistry 27: 4730-3734. EDELMAN,A. M., BLUMENTHAL,D. K. and KREBS, E. G. (1987) Protein serine/threonine kinases. A. Rev. Biochem. 56: 567--613.

ENGEL,D. A., MULLER,U., GEDRICH,R. W., EUaANKS,J. S. and SFENK,T. (1991) Induction of c-los mRNA and AP-I DNA-binding activity by cAMP in cooperation with either the adenovirus 243- or the adenovirus 289-amino acid E1A protein. Proc. natn. Acad. Sci. U.S.A. 88: 3957-3961. ENGLANDERE. W., WIDEN,S. G. and WILSON,S. H. (1991) Mammalian/Lpolymerase promoter: phosphorylation of ATF/CRE-binding protein and regulation of DNA binding. Nucl. Acids Res. 19: 3369-3375.

ENSLEY, J. F., JACOBS, J. R., WEAVER, A., KINZIE, J., CRISSMAN,J., KISH,J. A., CUMMINGS,G. and AL-SARRAF, M. (1984) Correlation between response to cisplatinumcombination chemotherapy and subsequent radiotherapy in previously untreated patients with advanced squamous cell cancers of the head and neck. Cancer 54:811-814. ERLICH, H. A., GELFAND, D. and SNINSKY,J. J. (1991) Recent advances in the polymerase chain reaction. Science 252: 1643-1651. ERVlN, T. J., CLARK,J. R., WEICHSELBAUM,R. R., FALLON, B.G., MILLER,D., FABIAN,R. L., POSNV.g,M. g., NORRIS, C. M., JR, TUTTLE,S. A., SCHOENFELD,D. A., PRICE,K. N. and FREI, E., III. (1987) An analysis of induction and adjuvant chemotherapy in the multidisciplinary treatment of squamous-cell carcinoma of the head and neck. J. clin. Oncol. 5: 10-20.

400

K. J. SCANLON et al.

ETIENNE,M. C., BERNARD,S., FISCHEL,J. L., FORMENTO,P., GIOANNI,J., SANTINI,J., DEMARD,F., SCHNEIDER,i . and MILANO, G. (1991) Dose reduction without loss of efficacy for 5-fluorouracil and cisplatin combined with folinic acid. in vitro study on human head and neck carcinoma cell lines. Br. J. Cancer 63: 372-377. EVANS,W. K., SHEPHERD,F. A., FELD, R., OSOBA,D., DANG, P. and DEBOER,G. (1985) VP-16 and cisplatin as first-line therapy for small-cell lung cancer. J. clin. Oncol. 3: 1471-1477. EXTRA,J. M., ESPIE,M., CALVO,F., FERME,C., MIGNOT,L. and MARRY, M. (1990) Phase I study of oxaliplatin in patients with advanced cancer. Cancer Chemother. Pharmac. 25: 299-303. FARRELL,N., Qu, Y., FENG, L. and VAN HOUTEN,B. (1990) Comparison of chemical reactivity, cytotoxicity, interstrand cross-linking and DNA sequence specificity of bis(platinum) complexes containing monodentate or bidentate coordination spheres with their monomeric analogs. Biochemistry 29: 9522-9531. FICHTINGER-SCHEPMAN,A. M. J., VAN DER VEER,J. L., DEN HARTOG,J. H. J., LOH~AN,P. H. M. and REEDIJK,J. (1985) Adducts of the antitumor drug cis-diaminedichloroplatinum(II) with DNA: formation, identification and quantitation. Biochemistry 24: 707-713. FICHTINGER-SCHEPMAN,A. M. J., VAN OOSTEROM, A. T., LOHMAN, P. H. M. and BERENDS, F. (1987) Cisdiaminedichloroplatinum(II)-induced DNA adducts in peripheral leukocytes from seven cancer patients: Quantitative immunochemical detection of the adduct induction and removal after a single dose of cis-diaminedichloroplatinum(II). Cancer Res. 47: 3000-3004. FLANAGAN, W. M., CORTHESY, B., BRAM, R. J. and CRABTREE,G. R. (1991) Nuclear association of a T-cell transcription factor blocked by FK-506 and cyclosporin A. Nature 352: 803-807. FoJO, A., HAMILTON,T. C., YOUNG,R. C. and OZOLS, R. F. (1987) Multidrug resistance in ovarian cancer. Cancer 60: 2075-2080. FORSTER,A. C. and SYMONS,R. H. (1987) Self-cleavage of plus and minus RNAs of a virusoid and a structural model for the active sites. Cell 49:211-220. FOSTER, B. J., HARDING, B. J., WOLPERT-DEFILIPPES, M. K., RUmNSTEIN, L. Y., CLAGEYr-CARR, K. and LEYLAND-JONES,B. (1990) A strategy for the development of two clinically active cisplatin analogs: CBDCA and CHIP. Cancer Chemother. Pharmac. 25: 395-404. FOULKES, N. S., LAOIDE, B. M., SCHLOTa~R, F. and SASSON~-CORSI, P. (1991) Transcriptional antagonist cAMP-responsive element modulator (CREM) downregulates c-los cAMP-induced expression. Proc. Dam. Acad. Sci. U.S.A. 88: 5448-5452. Fox, M. and ROBERTS,J. J. (1987) Drug resistance and DNA repair. Cancer and Metastasis Reviews 6: 261-281. FRAM, R. J., ROBICHAUD,N., BISCHOV,S. D. and WILSON, J. M. (1987) Interactions of cis-diaminedichloroplatinum(II) with 1-/LD-arabinofuranosylcytosine in LoVo colon carcinoma cells. Cancer Res. 47: 3360-3365. FRAM,R. J., WODA,B. A., WILSON,J. M. and ROBICHAUD,N. (1990) Characterization of acquired resistance to cisdiaminedichloroplatinum(II) in BE human colon carcinoma cells. Cancer Res. 50: 72-77. FREI, E., III. (1985) Curative cancer chemotherapy. Cancer BeN. 45: 6523-6537. FKEI, E., III., Cuccm, C. A., ROSOWSKY,A., TANTRAVAHI, R., BERNAL, S., ERVIN, T. J., RUPRECHT, R. M. and HASELTINE, W. A. (1985) Alkylating agent resistance in vitro studies with human cell lines. Proc. hath. Acad. Sci. U.S.A. 82: 2158-2162. FRILING, R. S., BERGEL~N, S. and DANIEL,V. (1992) Two adjacent AP-l-like binding sites form the electrophileresponsive element of the murine glutathione S-trans-

ferase Ya subunit gene. Proc. hath. Acad. Sci, U.S.A. 89: 668-672. FUJIWARA, Y., KASAHARA,K., SUGIMOTO,Y., NISHIO, K., OHMORI, T. and SAUO, N. (1990) Detection of proteins that recognize platinum-modified DNA using gel mobility shift assay. Jpn. J. Cancer Res. 81: 1210-1213. FUNATO,T., YOSHIDA,E., JIAO,L., TOHE,T., KASHANI-SABET, M. and SCANLON,K. J. (1992) The utility of an anti-fos ribozyme in reversing cisplatin resistance in human carcinomas. Adv. Enz. Regul., 32, 195-209. GALLIGIONI, E., CANOBBIO, L., FIGOLI, F., FASSIO, T., FRUSTACI, S., CRIVELLARI,D., VACCHER,E., LO RE, O., GASPARINI,G., VERO~r~SI,A. and MONFARDINI,S. (1987) Cisplatin and 5-fluorouracil combination chemotherapy in advanced and/or metastatic colorectal carcinoma: A phase II study. Eur. J. Cancer din. Oncol. 23: 657--661. GILL, I., MUGGIA,F. M., TERHEG~EN,P. M. A. B., MICHAEL, C., PARKER,R. J., KORTES,V., GRUNBERG,S., CHRISTIAN, M. C., REED, E. and DEN ENGELSE, L. (1991) Doseescalation study of carboplatin (day 1) and cisplatin (day 3): Tolerance and relation to leukocyte and buccal cell platinum-DNA adducts. Ann. Oncol. 2:115-121. GILLILAND,G., PERRIN,S., BLANCHARD,K. and BUNN,H. F. (1990) Analysis of cytokine mRNA and DNA: Detection and quantitation by competitive polymerase chain reaction. Proc. hath. Acad. Sci. U.S.A. 87: 2725-2729. GINSBERG,D., MECHTA,F., YANIV,M. and ORES, M. (1991) Wildtype p53 can down-modulate the activity of various promoters. Proc. narD. Acad. Sci. U.S.A. 88: 9979-9983. GIOVANELLA, B. C., STEHLIN, J. S., WALL, M. E., WANI, M. C., NICHOLAS, A. W., LIU, L. F., SILBER, R. and POTMESIL, M. (1989) DNA Topoisomerase I-targeted chemotherapy of human colon cancer in xenografis. Science 246: 1046-1048. GREENBERG,M. E. and ZIFF,E. B. (1984) Stimulation of 3T3 cells induces transcription of the c - fON proto-oncogene. Nature 311: 433-438. GUNTER, K. C., IRVING,S. G., ZIPFEL,P. F., SIEBENLIST,U. and KELLY,K. (1989) Cyclosporin A-mediated inhibition of mitogen-induced gene transcription is specific for the mitogenic stimulus and cell type. J. Immanol. 142: 3286-329 I. GUYER,R. L. and KOSHLAND,D. E., JR (1989) The molecule of the year. Science 246: 1543-1546. HAl, T. and CURRAN,T. (1991) Cross-family dimerization of transcription factors Fos/Jun and ATF/CREB alters DNA specificity. Proc. hath. Acad. Sci. U.S.A. 88: 3720-3724. HAINSWORTH,J. D., WILLIAMS,S. D., EINHORN,L. H., BmCH, R. and GREED, F. A. (1985) Successful treatment of resistant germinal neoplasms with VP-16 and cisplatin: Results of a southeastern cancer study group trial. J. elin. Oncol. 3:666-671. HALL, D. J. and STILES,C. D. (1987) Platelet-derived growth factor-inducible genes respond differentially to at least two distinct intracellular second messengers. J. biol. Chem. 262: 15302-15308. HAMER,D. H. (1986) Metallothionein. A. Rev. Biochem. 55: 913-951. HAMILTON,T. C., WINKER,M. A., LOUIE,K. G., BATIST,G., BEHRENS,B. C., TSURUO,T., GROTZlNGF.~,K. R., McKoY, W. M., YOUNG,R. C. and OZOLS,R. F. (1985) Augmentation of adriamycin, melphalan and cisplatin cytotoxicity in drug-resistant and -sensitive human ovarian carcinoma cell lines by buthionine sulfoximine mediated glutathione depletion. Biochem. Pharmac. 34: 2583-2586. HANSSON,J. and WOOD, R. D. (1989) Repair synthesis by human cell extracts in DNA damaged by cis- and transdiaminedichloroplatinum(II). Nucleic Acids Res. 17: 8073-8091. HANSSON,J., KEYSE,S. M., LINDAHL,T. and WOOD, R. D. (1991) DNA excision repair in cell extracts from human

Cisplatin resistance in human cancers cell lines exhibiting hypersensitivity to DNA-damaging agents. Cancer Res. 51: 3384-3390. HARSTRICK,A., CASPER,J., GUBA, R., WILKE, H., POLIWODA, H. and SCHMOLL, H. J. (1989) Comparison of the antitumor activity of cisplatin, carboplatin and iproplatin against established human testicular cancer cell lines in vivo and in vitro. Cancer 63: 1079--1083. HARTIG, E., LONCAREVIC,I. F., BUSCHER, i . , HERRLICH, P. and RAHMSDORF, H. J. (1991) A new cAMP response element in the transcribed re#on of the human c-fos gene. Nucl. Acids Res. 19: 4153-4159. HASELOFF, J. and GERLACH, W. L. (1988) Simple RNA enzymes with new and highly specific endoribonuclease activities. Nature 334: 585-591. HEIGER-BERNAYS, W. J., ESSIGMANN, J. M. and LIPPARD, S. J. (1990) Effect of the antitumor drug cis-diaminedichloroplatinum(II) and related platinum complexes on eukaryotic DNA replication. Biochemistry 29: 8461-8466. HERMAN, T. S. and TEXCHER, B. A. (1988) Sequencing of trimodality therapy [cis-diaminedichloroplatinum(II)/ hyperthermia/radiation] as determined by tumor growth delay and tumor cell survival in the FSalIC fibrosarcoma. Cancer Res. 48: 2693-2697. HERMAN, T. S., TEICHER, B. A., CHAN, V., COLLINS, U S., KAUFMANN, M. E. and LOH, C. 0988) Effect of hyperthermia on the action of cis-diaminedichloroplatinum(II), rhodamie 1232[tetrachloroplatinum(II)], rhodamine 123 and potassium tetrachloroplatinate in vitro and in vivo. Cancer Res. 48: 2335-2341. HERMAN, T. S., TEICHER, B. A., HOLDEN, S. A., PFEFFER, M. R. and JONES, S. M. (1990) Addition of 2-nitroimidazole radiosensitizers to cis-diaminedichloroplatinum(II) with radiation and with or without hyperthermia in the murine FSalIC fibrosarcoma. Cancer Res. 50: 2734-2740. HIDAKA, H. and TANAKA, T. (1983) Naphthalenesulfonamides as calmodulin antagonists. Meth. Enzym. 102: 185-194. HILL, B. T. and BELLAMV,A. S. (1984) Establishement of an etoposide-resistant human epithelial tumor clel line in vitro: Characterization of patterns of cross-resistance and drug sensitivities. Int. J. Cancer 33: 599-608. HILL, B. T., WHELAN, R. D. H., GIBBY, E. M., SHEER, D., HOSKING, L. K., SHELLARO, S. A. and RUPNIAK, H. T. (1987) Establishment and characterization of three new human ovarian carcinoma cell lines and initial evaluation of their potential in experimental chemotherapy studies. Int. J. Cancer 39: 219-225. HILL, B. T., WHELAN,R. D. H., HOSKING, L. K., SHELLARD, S. A., BEDFORD, P. and LOCK, R. B. (1988) Interactions between antitumor drugs and radiation in mammalian tumor cell lines: Differential drug responses and mechanisms of resistance following fractionated X-irradiation or continuous drug exposure in vitro. NCI Monogr. 6: 177-181. HILL, B. T., SHELLARD, S. A., HOSKING, L. K., FICHTINGER-SCHEPMAN, A. M. J. and BEDFORD,P. (1990) Enhanced DNA repair and tolerance of DNA damage associated with resistance to cis-diaminedichloroplatinum (II) after in vitro exposure of a human teratoma cell line to fractionated X-irradiation. Int. d. Radiat. Oncol. Biol. Phys. 19: 75-83. HILLS, C. A., KELLAND, L. R., ABEL, G., SIRACKY, J., WmSON, A. P. and HARRAP, K. R. (1989) Biological properties of ten human ovarian carcinoma cell lines: Calibration in vitro against four platinum complexes. Br. J. Cancer 59: 527-534. HOEIJMAKERE, J. H. J. and BOOTSMA,D. (1990) Molecular genetics of eukaryotic DNA excision repair. Cancer Cells 2: 311-320. HOFMANN, J., DOPPLER,W., JAKOB,A., MALY,K., POSCH, L., UBERALL, F. and GRUNICKE, H. H. (1988) Enhancement of the antiproliferative effect of cis-diaminedichloro-

401

platinum(II) and nitrogen mustard by inhibitors of protein kinas¢ C. Int. J. Cancer 42: 382-388. HOLLANDER, M. C. and FORNACE,A. J., JR (1989) Induction of fos RNA by DNA-damaging agents. Cancer Res. 49:, 1687-1692. HOLLIS, L. S., SUNDQUIST, W. I., BURSTYN, J. N., HEIGER-BERNAYS, W. J., BELLON, S. F., ArIMED, K. J., AMUNDSEN,A. R., STERN, E. W. and LIPPARD, S. J. (1991) Mechanistic studies of a novel class of trisubstituted platinum(II) antitumor agents. Cancer Res. 51: 1866-1875. HOLT, J. T., GOPAL, T. V., MOULTON, A. D. and NIENHUIS, A. W. (1986) Inducible production of a c-fos antisense RNA inhibits 3T3 cell proliferation. Proc. hath. Acad. Sci. U.S.A. 83: 4794--4798. HONG, W.-S., SAIJO, N., SASAKI,Y., MINATO, K., NAKANO, H., NAKAGAWA, K., FUJIWARA, Y., NOMURA, K. and TWENTYMAN,P. R. (1988) Establishment and characterization of cisplatin-resistant sublines of human lung cancer cell lines. Int. J. Cancer 41: 462-467. HOPWOOD, L. E. and MOULDER, J. E. 0989) Radiation induction of drug resistance in RIF-1 tumors and tumor cells. Radiat. Res. 120: 251-266. HOPWOOD, L. E., DAVIES, B. M. and MOULDER, J. E. (1990) Drug resistance following irradiation of RIF-I tumors: influence of the interval between irradiation and drug treatment. Int. J. Radiat. Oncol. Biol. Phys. 19: 643-650. HOSPERS, G. A. P., MULDER, N. H., DE JONG, B., DE LEY, L., UGES, n . R. A., FICHTINGER-SCHEPMAN, A. i . J., SCl-mPER, R. J. and DE VRIES,E. G. E. (1988) Characterization of a human small cell lung carcinoma cell line with acquired resistance to cis-diaminedichloroplatinum(II) in vitro. Cancer Res. 48: 6803-6807. HOSPERS, G. A. P., DE VRIES, E. G. E. and MULDER, N. H. (1990a) The formation and removal of cisplatin (CDDP) induced DNA adducts in a CDDP sensitive and resistant human small cell lung carcinoma (HSCLC) cell line. Br. J. Cancer 61: 79-82. HOSPERS, G. A. P., MELIER, C., DE LEIJ, L., UGES, D. R. A., MULDER, N. H. and DEVries, E. G. E. (1990b) A study of human small-cell lung carcinoma (hSCLC) cell lines with different sensitivities to detect relevant mechanisms of cisplatin (CDDP) resistance. Int. J. Cancer 46: 138-144. HOWELL, S. B. and GILL, S. (1986) Lack of synergy between cisplatin and cytarabine against ovarian carcinoma in vitro. Cancer Treat. Rep. 70: 409-410. HUANG, T.-S., LEE, S.-C. and LIN, J.-K. (1991) Suppression of c-Jun/AP-I activation by an inhibitor of tumor promotion in mouse fibroblast cells. Proc. hath. Acad. Sci. U.S.A. 88: 5292-5296. HUNTER, T. (1991) Cooperation between oneogenes. Cell 64: 249-270. ISONISHI, S., ANDREWS, P. A. and HOWELL, S. B. (1990) Increased sensitivity to cis-diaminedichloroplatinum(II) in human ovarian carcinoma cells in response to treatment with 12-O-tetradecanoylphorbol 13-acetate. J. biol. Chem. 265: 3623-3627. ISONISHi, S., HOM, D. K., THIEBAUT, F. B., MANN, S. C., ANDREWS, P. A., BASU,A., LAZO, J. S., EASTMAN,A. and HOWELL, S. B. (1991) Expression of the c-Ha-ras oncogene in mouse NIH 3T3 cells induces resistance to cisplatin. Cancer Res. 51: 5903-5909. JONES, J. C., ZHEN, W., REED, E., PARKER, R. J., SANCAR,A. and BOHR, V. A. (1991) Gene-specific formation and repair of cisplatin intrastrand adducts and interstrand crossqinks in Chinese hamster ovary cells. J. biol. Chem. 266: 7101-7107. KASAHARA, K., FUJIWm,A, Y., NISHIO, K., OHMORI, T., SUGIMOTO, Y., KOMIYA, K., MATSUDA, T. and SAIJO, N. (1991) Metallothionein content correlates with the sensitivity of human small cell lung cancer cell lines to cisplatin. Cancer Res. 51: 3237-3242.

402

K. J. SCANLONet al.

KASHANI-SABET,M. and SCANLON,K. J. (1989) Drug selectivity and membrane transport properties of normal human lymphocytes. Chemotherapy 35: 363-372. KASHANI-SABET,M., ROSSl,J. J., LU, Y., MA., J. X., CHEN,J., M1YACHI,H. and SCANLON,K. J. (1988) Detection of drug resistance in human tumors by in vitro enzymatic amplification. Cancer Res. 48: 5775-5778. KASHANI-SABET,M., LU, Y., LEONG, L., HAEDICKE,K. and SCANLON,K. J. (1990a) Differential oncogene amplification in tumor cells from a patient treated with cisplatin and 5-fluorouracil. Fur. J. Cancer 26: 383-390. KASHANI-SABET,M., WANG,W. and SCANLON,K. J. (1990b) Cyclosporin A suppresses cisplatin-induced c-fos gene expression in ovarian carcinoma cells. J. biol. Chem. 265: 11285-11288. KASHANI-SABET,M., FUNATO,T., TONE,T., JIAO, L., WANG, W., YOSHIDA,E., WE, A. M., MORENO,J. G., TRAWEEK, S. T., AHLERING,T. E. and SCANLON,K. J. (1992) Reversal of the malignant phenotype by an anti-ras ribozyme. Antisense Res. Develop., 2: 3-15. KATZ, E. J., ANDREWS,P. A. and HOWELL,S. B. (1990) The effect of DNA polymerase inhibitors on the cytotoxicity of cisplatin in human ovarian carcinoma cells. Cancer Commun. 2: 159-164. KAWAI,K., KAMATANI,N., GEORGES,E. and LING, V. (1990) Identification of a membrane glycoprotein overexpressed in murine lymphoma sublines resistant to cis-diaminedichloroplatinum(II). J. biol. Chem. 265: 13137-13142. KEDAR, P. S., LOWY, n. R., WIDEN, S. G. and WILSON,S. H. (1990) Transfected human /~-polymerase promoter contains a ras-responsive element. Mol. Cell. Biol. 10: 3852-3856. KEDAR, P. S., WIDEN, S. G., ENGLANDER,E. W., FORNACE, A. J., JR and WILSON, S. H. (1991) The ATF/CREB transcription factor-binding site in the polymerase fl promoter mediates the positive effect of N-methyl-N'nitro-N-nitrosoguanidine on transcription. Proc. HatH. Acad. Sci. U.S.A. 88: 3729-3733. KELLEY, S. L. and ROZENCWEIG,M. (1989) Resistance to platinum compounds: Mechanisms and beyond. Eur. J. Cancer clin. Oncol. 25: 1135-1140. KELLEY, S. L., BASU, A., TEICHER, B. A., HACKER, M. P., HAMER, D. H. and LAZO,J. S. (1988) Overexpression of metallothionein confers resistance to anticancer drugs. Science 241: 1813-1815. KEMENY, N., ISRAEL,K., NIEDZWIECKI,D., CHAPMAN,D., BOTET,J., MINSKY,B., VINCIGUERRA,V., ROSENBLUTH,R., BOSSELLI,B., COCHRAN,C. and SHEEHAN,K. (1990) Randomized study of continuous infusion fluorouracil versus fluorouracil plus cisplatin in patients with metastatic colorectal cancer. J. clin. Oncol. 8: 313-318. KERN, D. H., MORGAN,C. R. and HILDEBRAND-ZANKI,S. U. (1988) In vitro pharmacodynamics of l-/~-D-arabinofuranosylcytosine: Synergy of antitumor activity with cisdiaminedichloroplatinum(II). Cancer Res. 48: 117-121. KHANSUR, T. and KENNEDY, A. (1991) Cisplatin and 5fluorouracil for advanced locoregional and metastatic squamous cell carcinoma of the skin. Cancer 67: 2030-2032. KIKUCHI, Y., MIYAUCHI,M., KIZAWA, I., OOMORI,K. and KATO, K. (1986) Establishment of a cisplatin-resistant human ovarian cancer cell line. J. natn. Cancer Inst. 77: 1181-1185. KIKUCHI,Y., OOMORI,K., KIZAWA,I., HIRATA,J., KITA, T., MIYAUCHI, M. and KATO, K. (1987) Enhancement of antineoplastic effects of cisplatin by calmodulin antagonists in nude mice bearing human ovarian carcinoma. Cancer Res. 47: 6459-6461. KIKUCHI, Y., IWANO, I., MIYAUCHI,M., KITA, T., SUGITA, M., TENJIN, Y. and NAGATA,I. (1990) Possible mechanisms of resistance to cis-diaminedichloroplatinum(II) of human ovarian cancer cells, dpn. J. Cancer Res. 81: 701-706.

KINGSTON, R. E., SEViN, B.-U., RAMOS, R., SAKS, M., DONATO, D., JARRELL, M. A. and AVES£TTE, H. E. (1989) Synergistic effects of cis-platinum and cystosine arabinoside on ovarian carcinoma cell lines, demonstrated by dual=parameter flow cytometry. Gyn. Oncol. 32: 282-287. KOUZARIDES,T. and ZIFF, E. (1989) Behind the Fos and Jun leucine zipper. Cancer Cells 1: 71-76. KUNZE,N., KLEIN,M., RICHTER,A. and KNIPPEgS,R. (1990) Structural characterization of the human DNA topoisomerase I gene promoter. Fur. £. Biochem. 194: 323-330. KUPPEN, P. J. K., SCHUITEMAKER,H., VAN'T VEER, L. J., DEBRUIJN,E. A., VANOOSTEROM,A. T. and SCHRIER,P. I. (1988) cis-Diaminedichloroplatinum(II)-resistant sublines derived from two human ovarian tumor cell lines. Cancer Res. 48: 3355-3359. LAI, G.-M., OZOLS, R. F., SMYTH,J. F., YOUNG, R. C. and HAMILTON, T. C. (1988) Enhanced DNA repair and resistance to cisplatin in human ovarian cancer. Biochem. Pharmac. 37: 4597-4600. LAX, G.-M., OZOLS, R. F., YOUNG, R. C. and HAMILTON, T. C. (1989) Effect of glutathione on DNA repair in cisplatin-resistant human ovarian cancer cell lines. J. natn. Cancer Inst. 81: 535-539. LAZO, J. S. and BASU,A. (1991) Metallothionein expression and transient resistance to electrophilic antineoplastic drugs. Cancer Biol. 2: 267-271. LEDWITH, B. J., MANAM, S., KRAYNAK,A. R., NICHOLS, W. W. and BRADLEY,M. O. (1990) Antisense-fos RNA causes partial reversion of the transformed phenotypes induced by the c-Ha-ras oncogene. Mol. Cel. Biol. 10: 1545-1555. LEE, W., HASLINGER,A., KARIN, M. and TJIAN, R. (1987a) Activation of transcription by two factors that bind promoter and enhancer sequences of the human metallothionein gene and SV40. Nature 325: 368-372. LEE, W., MITCHELL, P. and THAN, R. (1987b) Purified transcription factor AP-I interacts with TPA-inducible enhancer elements. Cell 49: 741-752. LEMAIRE, M.-A., SCHWARTZ, A., RAHMOUNI, A. R. and LENG,M. (1991) Interstrand cross-links are preferentially formed at the d(GC) sites in the reaction between cisdiaminedichloroplatinum(II) and DNA. Proc. natn. Acad. Sci. U.S.A. 88: 1982-1985. LEVI, F., MISSET,J. L., BRIENZA,S., ADAM,R., GASTIABURU, J., LE COUTURIER,S., BOURDOULOUS,METZGER,G. and BISMUTH,H. (1991) Circadian-rhythm modulated chemotherapy against metastatic colorectal cancer: Results of an extended Phase II chromotherapy trial with 5-fluorouracil, folinic acid and oxalato-platinum using an ambulatory multichannel programmable in the time pump. In: Platinum and Other Metal Coordination Compounds in Cancer Chemotherapy, Howell, S. (ed.) Plenum Press,

N.Y., in press. LEVlNE,A. J., MOMAND,J. and FINLAY,C. A. (1991) The p53 tumor suppressor gene. Nature 351: 453-456. LOEHRER,P. J. and EINHORN,L. H. (1984) Cisplatin. Ann. intern. Med. 100: 704-713. Loum, K. G., BEHRENS,B. C., KINSELLA,T. J., HAMILTON, T. C., GROTZINGER, K. R., McKoY, W. M., WINKER, M. A. and OZOLS, R. F. (1985) Radiation survival parameters of antineoplastic drug=sensitive and -resistant human ovarian cancer cell lines and their modification by buthionine sulfoximine. Cancer Res. 45: 2110-2115.

Loum, K. G., OZOLS, R. F., MYERS, C. E., OSTCHEGA,Y., JENKINS,J., HOWSER,D. and YOUNO,R. C. (1986) Longterm results of a cisplatin-containing combination chemotherapy regimen for the treatment of advanced ovarian carcinoma. J. din. Oncol. 4: 1579-1585. Lu, Y., HAN, J. and SCANLON,K. J. (1988) Biochemical and molecular properties of cisplatin-resistant A2780 ceils grown in folinic acid. J. biol. Chem. 263: 4891--4894.

Cisplatin resistance in human cancers

4O3

MANN, S. C., ANDREWS,P. A. and HOWELL, S. B. (1991) protooncogenes c-jun and c-fos as regulators of DNA replication. Proc. natn. Acad. Sci. U.S.A. 88: 3947-3951. Modulation of cis-diaminedichloroplatinum(II) accumuMURPHY, D., McGowN, A. T., CROWTHER,D., MANDER,A. lation and sensitivity by forskolin and 3-isobutyl-1and Fox, B. W. (1991) Metallothionein levels in ovarian methylxanthine in sensitive and resistant human ovarian carcinoma cells. Int. J. Cancer 48: 866-872. tumors before and after chemotherapy. Br. J. Cancer 63: MANSOURLA., HENLE,K. J., BENSON,A. M., MOSS,A. J. and 711-714. NAGLE, W. A. (1989) Characterization of a cisplatin- NEGRIN, R. S. and BLUME,K. G. (1991) The use of the polymerase chain reaction for the detection of minimal resistant subline of routine RIF-I cells and reversal of drug resistance by hyperthermia. Cancer Res. 49: residual malignant disease. Blood 78: 255-258. 2674-2678. NEWMAN, E. M., LU, Y., KASHANI-SABET,M., K.ESAVAN,V. MASUDA, H., OZOLS, R. F., LAI, G.-M., FOJO, A., and SCANLON, K. J. (1988) Mechanisms of crossROTHENBERG,M. and HAMILTON,T. C. (1988) Increased resistance to methotrexate and 5-fluorouracil in an A2780 DNA repair as a mechanism of acquired resistance to human ovarian carcinoma cell subline resistant to cisc/s-diaminedichloroplatinum (II) in human ovarian platin. Biochem. Pharmac. 37: 443-447. cancer cell lines. Cancer Res. 48: 5713-5716. NnMI, S., NAKAGAWA,K., YOKOTA, J., TSUNOKAWA,Y., MASUDA, H., TANAKA,T., MATSUDA,H. and KUSABA, I. NISHIO,K., TERASHIMA,Y.~ SHIBUYA,M., TERADA,M. and (1990) Increased removal of DNA-bound platinum in a SAIJO, N. (1991) Resistance to anticancer drugs in human ovarian cancer cell line resistant to cis-diamineNIH3T3 cells transfected with c-myc and/or c-H-ras dichioroplatinum(II). Cancer Res. 50: 1863-1866. genes. Br. J. Cancer 63: 237-241. MATTILA,P. S., ULLMAN,K. S., FIERING,S., EMMEL,E. A., NISHIKURA,K. and MURRAY,J. M. (1987) Antisense RNA McCuTctmON, M., CRABTREE,G. R. and HERZENBERG, of proto-oncogene c-los blocks renewed growth of quiesL. A. (1990) The actions of cyclosporin A and FK506 cent 3T3 cells. Mol. Cel. Biol. 7: 639-649. suggest a novel step in the activation of T lymphocytes. NOONAN, K. E., BECK, C., HOLZMAYER,T. A., CmN, J. E., EMBO J. 9: 4425-4433. WUNDER,J. S., ANDRULIS,I. L., GAZDAR,A. F., WILLMAN, McDoNNELL, S. E., KERR, L. D. and MATRISIAN,L. M. C. L., GRIFFITH,B., VON HOFF, D. D. and RONINSON,I. B. (1990) Epidermal growth factor stimulation of (1990) Quantitative analysis of MDR1 (Multidrug resiststromelysin mRNA in rat fibroblasts requires induction ance) gene expression in human tumors by polymerase of proto-oncogenes c-fos and c-jun and activation of chain reaction. Proc. hath. Acad. Sci. U.S.A. 87: protein kinase C. Mol. Cell. Biol. 10: 4284-4293. 7160-7164. MCKEAGE, M. J., HIC,GINS,J. D. and KELLAND,L. R. (1991) OESTERREICH,S., SCHUNCK,H., BENNDORF,R. and BmLKA, Platinum and other metal coordination compounds in H. (1991) Cisplatin induces the small heat shock protein cancer chemotherapy. A commentary on the sixth HSP25 and thermotolerance in Ehrlich ascites tumor international symposium: San Diego, California, 23-26 cells. Biochem. biophys. Res. Commun. 180: 243-248. January 1991. Br. J. Cancer 64: 788-792. ONODA,J. M., JACOaS,J. R., TAYLOR,J. D., SLOANE,B. F. and McLAUGHLIN,K., STEPHENS,I., MCMAHON,N. and BROWN, HONN, K. V. (1986) Cisplatin and nifedipine: synergistic R. (1991) Single step selection of cis-diaminedichlorocytotoxicity against murine solid tumors and their metasplatinum(II) resistant mutants from a human ovarian tases. Cancer Lett. 30:. 181-188. carcinoma cell line. Cancer Res. 51: 2242-2245. OSIEKA, R., SEEBER, S., PANNENBACKER,R., SOLL, D., MEIJER, C., MULDER, N. H., HOSPERS, G. A. P., UGES, GLATTE, P. and SCHMIDT,C. G. (1986) Enhancement of D. R. A. and DE Vries, E. G. E. (1990) The role of etoposide-induced cytotoxicity by cyclosporin A. Cancer glutathione in resistance to cisplatin in a human small cell Chemoth. Pharmac. 18: 198-202. lung cancer cell line. Br. J. Cancer 62: 72-77. OZAWA, S., SUGIYAMA,Y., MITSUHASHI,J. and INABA,M. MEISTER, A. and ANDERSON,M. E. (1983) Glutathione. A. (1989) Kinetic analysis of cell killing effect induced by Rev. Biochem. 52:711-760. cytosine arabinoside and cisplatin in relation to cell cycle METCALFE,S. A., CAIN, K. and HILL, B. T. (1986) Possible phase specificity in human colon cancer and Chinese mechanism for differences in sensitivity to cis-platinum in hamster cells. Cancer Res. 49: 3823-3828. human prostate tumor cell lines. Cancer Lett. 31: OZOLS, R. F. and YOUNG, R. C. (1984) Chemotherapy of 163-169. ovarian cancer. Semin. Oncol. l h 251-263. MISTRY, P., KELLAND, L. R., ABEL, G., SIDHAK, S. OZOLS, R. F., LOUIE,K. G., PLOWMAN,J., BEHRENS,B. C., and HARRAP, K. R. (1991) The relationships between FINE, R. L., DYKES, D. and HAmLTON, T. C. (1987) glutathione, glutathione-S-transferase and cytotoxicity Enhanced melphalan cytotoxicity in human ovarian of platinum drugs and melphalan in eight human cancer in vitro and in tumor-bearing nude mice by ovarian carcinoma cell lines. Br. J. Cancer 64: buthionine sulfoximine depletion of glutathione. 215-220. Biochem. Pharmac. 36: 147-153. MIVECHI, N. F., MIYACHI, H. and SCANLON,K. J. (1990) PAPATHANASIOU,M. A., K~R, N. C. K., ROemNS, J. H., Heat radiosensitization and the level of DNA polyMcBRIDE,O. W., ALAMO,I., JR, BARRETT,S. F., HICKSON, merases c~and ~ of human colony-forming unit-granuloI. D. and FORNACE,A. J., JR (1991) Induction by ionizing cyte-macrophage and myeloid leukemias sensitive and radiation of the gadd45 gene in cultured human cells: resistant to chemotherapeutic agents. Cancer Res. 50: Lack of mediation by protein kinase C. Mol. Ceil. Biol. 2044-2048. 11: 1009-1016. MI'~'Acm, H., HAN, H. and SCANLON,K. J. (1991) Dihydro- PARDr~, A. B. (1987) Molecules involved in proliferation of folate reductase gene expression characterized by the normal and cancer cells: Presidential address. Cancer Res. PCR assay in human leukemia cells. In Vivo 5: 7-12. 47: 1488-1491. MUC~IA, F. M (1991) Introduction: Cisplatin update. PAI~DES, J,, HONO, W. K., FELDER, T. B., DIM~RY, I. W., Semin. Oncol. 18: 1-4. CHOKSI, A. J., NEWMAN, R. A., CASTELLANO$,A. M., MULLER, R., GUST, R., BERNHARDT, G., KELLER, C., ROm31NS,K. T., M c C o y , K., ATKINSON,N,, g a t A ~ , SCHONENBERGER,H., SEEBVm,S., OSmKA,R., EASTMAN,A. A. M., HEgstt, E. M. and GOEPEmT,H. (1988) Prospective and JENmmWEIN, M. (1990) [DL-l,2-Bis(2-hydroxyrandomized trial of high-dose cisplatin and fluorouraeil phenyl)ethylene diamine]dichloroplatinum(II), a new infusion with or without sodium diethyldithiocarbamate compound for the therapy of ovarian cancer. Fur. in recurrent and/or metastatic squamous cell carcinoma J. Cancer clin. Oncol. 116: 237-244. of the head and neck. J. clin. Oncol. 6: 955--962. MtmAKAm, Y., SAXAKE,M., YAMAGUCm-IwAi,Y., SAK~a, PARKER, R. J., GILL, I., T~O~m, R., VIom,mST, J. A., M., MU~MATSU, M. and ITO, Y. (1991) The nuclear GRtmeERO, S., Muoom, F. M., R.ImD, E. (1991) JPT S2/~J

404

K. J. SCANLONet al.

Platinum-DNA damage in leukocyte DNA of patients receiving carboplatin and cisplatin chemotherapy, measured by atomic absorption spectrometry. Carcinogenesis 12: 1253-1258. PE~Z, R. P., HAMILTON,T. C. and OZOLS, R. F. (1990) Resistance to alkylating agents and cisplatin: Insights from ovarian carcinoma model systems. Pharmac. Ther. 48: 19-27. PEREZ, R. P., GODWIN,A. K., HAMILTON,T. C. and OZOLS, R. F. (1991a) Ovarian cancer biology. Semin. Oncol. lg: 186-204. PEREZ, R. P., O'DwYER, P. J., HANDEL,L. M., OZOLS,R. F. and HAMILTON,T. C. (1991b) Comparative cytotoxicity of CI-973, cisplatin, carboplatin and tetraplatin in human ovarian carcinoma cell lines. Int. J. Cancer 48: 265-269. POMMIER,Y., KERRIGAN,O., HARTMAN,K. D. and GLAZER, R. I. (1990) Phosphorylation of mammalian DNA topoisomerase I and activation by protein kinase C. J. biol. Chem. 265: 9418-9422. RAUSCHER,F. J., SAMBUCETTI,L. C., CURRAN,T., DISTEL, R. J. and SPIEGELMAN,B. M. (1988) Common DNA binding site for Fos protein complexes and transcription factor AP-I. Cell 52: 471-480. RECONDO,G., BENHAMEDM., CVlTKOVIC,E., MARANDAS,P. and ARMAND,J. P. (1989) 96-Hour continuous infusion of cis-platinum, 5-fluorouracil and bleomycin in recurrent or metastatic head and neck squamous cell carcinoma, unexpected anemia. Eur. J. Cancer clin. Oncol. 25: 15291533. REDDY, G. P. V. and PARDEE,A. B. (1980) Multienzyme complex for metabolic channeling in mammalian DNA replication. Proc. nam. Acad. Sci. U.S.A. 77: 3312-3316. REED, E. and KOHN, K. W. (1990) Platinum analogs, In: Cancer Chemotherapy: Principles and Practice Cisplatin and Platinum Analogs, pp. 465-490, CHABNER,B. A. and

COLLINS,J. (eds) J. B. Lippincott, Philadelphia. REED, E., YUSPA,S. H., ZWELLING,L. A., OZOLS,R. F. and POIRIER, M. C. (1986) Quantitation of cis-diaminedichloroplatinumlI (cisplatin)-DNA intrastrand adducts in testicular and ovarian cancer patients receiving cisplatin chemotherapy. J. clin. Invest. 77: 545-550. REED, E., OZOLS, R. F., TARONE, R., YUSPA, S. H. and POIRIER,M. C. (1987) Platinum-DNA adducts in leukocyte DNA correlate with disease response in ovarian cancer patients receiving platinum-based chemotherapy. Proc. nam. Acad. Sci. U.S.A. 84: 5024--5028. REED, E., SAUERrlOFF,S. and POIRmR, M. C. (1988) Quantitation of platinum-DNA binding after therapeutic levels of drug exposure--A novel use of graphite furnace spectometry. Atomic Spectrosc. 9: 93-95. REED, E., OSTCHEGA,Y., STEINBERG,S. M., YUSPA, S. H., YOUNG, R. C., OZOLS, R. F. and POIRIER,M. C. 0990) Evaluation of platinum-DNA adduct levels relative to known prognostic variables in a cohort of ovarian cancer patients. Cancer Res. 50: 2256-2260. R1ABOWOL,K. T., VOSATKA,R. J., ZIFF, E. B., LAMB,N. J. and FERAMISCO,J. R. (1988) Microinjection offos-specific antibodies blocks DNA synthesis in fibroblast cells. Mol. Cell. Biol. 8: 1670-1676. ROBERTS,J. and FRIEDLAS,F. (1987) Quantitative estimation of cisplatin-induced DNA interstrand cross-links and their repair in mammalian cells: Relationship to toxicity. Pharmac. Ther. 34: 215-246. ROBERTS, J. D., VAN HOUTEN, B., Qu, Y. and FARRELL, N. P. (1989) Interaction of novel bis(platinum) complexes with DNA. Nucleic. Acids Res. 17(23): 9719-9733. ROSE, W. C. and BASLER,G. A. (1990) In vivo model development of cisplatin-resistant and -sensitive A2780 human ovarian carcinomas. In Vivo 4: 391-396. ROSENBERG,B. (1985) Fundamental studies with cisplatin. Cancer 55: 2303-2316.

ROSENTHAL,S. A. and HALT,W. N. (1988) Potentiation of

DNA damage and cytotoxicity by calmodulin antagonists. Yale J. biol. Med. 61: 39-49. ROSOWSKY,A., WRIGHT,J. E., CUCCHI,C. A., FLATOW,J. L., TRITES, D. H., TEiCHER, B. A. and FREI, III, E. 0987) Collateral methotrexate resistance in cultured human head and neck carcinoma cells selected for resistance to cis-diaminedichloroplatinum(II). Cancer Res. 47: 5913-5918. ROWLAND, K. M. JR, TAYLOR, S. G., SPIERS, A. S. D., DECONTI, R. C., O'DONNELL,M. R., SHOWEL,J., STOTT, P. B., MILNER,L. M. and MARSH,J. C. (1986) Cisplatin and 5-FU infusion chemotherapy in advanced, recurrent cancer of the head and neck: An eastern cooperative oncology group pilot study. Cancer Treat. Rep. 70: 461-464. SASSONE-CORSl,P., SISSON,J. C. and VERMA,I. M. (1988) Transcriptional autoregulation of the proto-oncogene los. Nature 334: 314-319. SCANLON, K. J. and KASHANI-SABET,M. (1988) Elevated expression of thymidylate synthase cycle genes in cisplatin-resistant human ovarian carcinoma A2780 cells. Proc. nam. Acad. Sci. U.S.A. 85: 650-653. SCANLON,K. J. and KASHANI-SABET,M. (1989) Utility of the polymerase chain reaction in detection of gene expression in drug-resistant human tumors. J. clin. Lab. Anal. 3: 323-329. SCANLON,K. L, NEWMAN,E. M., Lu, Y. and PRIEST,D. G. (1986) Biochemical basis for cisplatin and 5-fluorouracil synergism in human ovarian carcinoma cells. Proc. natn. Acad. Sci. U.S.A. 83: 8923-8925. SCANt.ON, K. J., KASHANI-SABET,M., CASHMORE,A. R., PALLAI,M., MOROSON,B. A. and SAKETOS,M. (1987) The role of methionine in methotrexate-sensitive and methotrexate-resistant mouse leukemia LI210 cells. Cancer Chemother. Pharmac. 19: 25-29. SCANLON,K. J., Lu, Y., KASHANI-SAaET,M., MA. J. X. and NEWMAN,E. (1988) Mechanisms for cisplatin-FUra synergism and cisplatin resistance in human ovarian carcinoma cells both in vitro and in vivo. Adv. Exp. Med. Biol. 244: 127-135. SCANLON, K. J., KASHANI-SABET,M. and MIYACHI, H. (1989a) Differential gene expression in human cancer cells resistant to cisplatin. Cancer Invest. 7: 581-587. SCANLON,K. J., KASHANI-SABET,M., MIYACHI,S., SOWERS, L. C. and RoSSl, J. J. (1989b) Molecular basis of cisplatin resistance in human carcinomas: Model systems and patients. Anticancer Res. 9: 1301-1312. SCANLON, K. J., KASHANI-SABET,i . and SOWERS, L. C. (1989c) Overexpression of DNA replication and repair enzymes in cisplatin-resistant human colon carcinoma HCT8 cells and circumvention by azidothymidine. Cancer Commun. 1: 269-275. SCANLON, K. J., FUNATO, T., PEZESHKI,B., TONE, T. and SOWERS, L. C. (1990a) Potentiation of azidothymidine cytotoxicity in cisplatin-resistant human ovarian carcinoma cells. Cancer Commun. 2: 339-343. SCANLON,K. J., WANG,W. and HAN, H. (I 990b) Cyclosporin A suppresses cisplatin-induced oncogene expression in human cancer cells. Cancer Treat. Rev. 17: 27-35. SCANLON,K. J., J1AO,L., FUNATO,T., WANG,W., TONE, T., ROSSl, J. J. and KASrO,NI-SABET, M. (1991) Ribozymemediated cleavage of c-fos mRNA reduces gene expression of DNA synthesis enzymes and metallothionein. Proc. hath. Acad. Sci. U.S.A. 88: 10591-10595. SCHABEL,F. M., TRADER,M. W., LASTER,W. R., CORBETT, T. H. and GRISWOLD, D. P., JR (1979) Cis-dichlorodiamineplatinum(II): Combination chemotherapy and cross-resistance studies with tumors of mice. Cancer Treat. Rep. 63: 1459-1473. SCHILDER, R. J., HALL, L., MONKS, A., HANDEL, L. M., FORNACE,A. J., OZOLS,R. F., FOJO,A. T. and HAMILTON, T. C. (1990) Metallothionein gene expression and

Cisplatin resistance in human cancers

405

SWaNNEN,U J., B~mr,ms, D. M., FIS~.R, S. G., ALnAm, K. S., resistance to cisplatin in human ovarian cancer. Int. FISHER, R. I. and ERICKSON,L. C. (1989) 1-~-D-arabinoJ. Cancer 45: 416-422. furanosylcytosine and hydroxyurea production of cytoSCrnLSKY, R. L., CHOl, K. E., GRAYHACK,J., GRIMMER,D., toxic synergy with c/s-diaminedichloroplatinum(II) and GUARNIERI,C. and FULLEM,U (1990) Phase I clinical and modification platinum-induced DNA interstrand crosspharmacologic study of intraperitoneal cisplatin and linking. Cancer Res. 49:. 1383-1389. fluorouracil in patients with advanced intraabdominal SWINNEN, L. J., ELLIS, N. K. and ERICKSON,L. C. (1991) cancer. J. clin. OncoL 8: 2054-2061. Inhibition of c/s-diamine-l, l-cyclobutane dicarboxylatoSCtmlDT, C. J. and HAMER,D. H. (1986) Cell specificity and platinum(II)-induced DNA interstrand cross-link rean effect of ras on human metallothionein gene exmoval and potentiation of c/s-diamine-l,l-cyclobutane pression. Proc. ham. Acad. Sci. U.S.A. 83: 3346-3350. dicarboxylatoplatinum(II) cytotoxicity by hydroxyurea SCHWARTZ,A., MARROT, L. and LENG, M. (1989) Conforand l-//-D-arabinofuranosylcytosine. Cancer Res. 51: mation of DNA modified at a d(GG) or a d(AG) site 1984-1989. by the antitumor drug c/s-diaminedichloroplatinum(II). TAKEISHI, K., KANEDA, S., AYUSAWA, D., SmMlZU, K., Biochemistry 28: 7975-7984. GOTOH,O. and SENO,T. (1989) Human thymidylate synSCHWARTZ, J. L., ROTMENSCH,J., "BECKET'r,M. A., JAFFE, thase gene: Isolation of phage clones which cover a D. R., TOOHILL, M., GIOVANAZZI,S. M., MCINTOSH, functionally active gene and structural analysis of the J. and WEICHSELBAUM,R. R. (1988) X-ray and cisregion upstream from the translation initiation codon. diaminedichloroplatinum(II) cross-resistance in human J. Biochem. 106: 575-583. tumor cell lines. Cancer Res. 48: 5133-5135. SSARMA,R. C. and ScmMKE, R. T. (1989) Enhancement of TAKIMOTO,M., QUINN,J. P., FARINA,A. R., STAUDT,L. M. and LEVENS, D. (1989) Fos/jun and octamer-binding the frequency of methotrexate resistance by gammaprotein interact with a common site in a negative radiation in chinese hamster ovary and mouse 3T6 cells, Cancer Res. 49: 3861-3866. element of the human c-myc gene. J. biol. Chem. 264: SHEAFF,R., ILSLEY,D. and KUCHTA,R. (1991) Mechanism 8992-8999. of DNA polymerase ~ inhibition by aphidicolin. TEEUNG, M. and CARNEY,D. N. (1990) Phase II study of carboplatin and cyclophosphamide combination chemoBiochemistry 30: 8590-8597. SHEmANL N. and EASTMAN,A, (1990) Analysis of various therapy for the treatment of advanced ovarian cancer. Eur. J. Gynaecol. Oncol. 11: 219-224. mRNA potentially involved in cisplatin resistance of murine leukemia L1210 cells. Cancer Lett. 52: 179-185. TEETER, L. D., ECKERSaERG,T., TSAI, Y. and Kuo M. T. SHEIBANI,N., JENNERWEIN,M. M. and EASTMAN,A. (1989) (1991) Analysis of the chinese hamster P-glycoprotein/ DNA repair in ceils sensitive and resistant to cismulti-drug resistance gene pgpl reveals that the AP-I site diaminedichloroplatinum(II): Host cell reactivation of is essential for full promoter activity. Cell Grow. D/ft. 2: damaged plasmid DNA. Biochemistry 28: 3120-3124. 429-437. SHELLARD, S. A., HOSKING, L. K. and HILL, B. T. (1991) TEICHER, B. A., HOLDEN,S. A., KELLEY,M. J., SHEA,T. C., Anomalous relationship between cisplatin sensitivity and CUCCHI, C. A., ROSOWSKY,A., HENNER,W. D. and FREI, the formation and removal of platinum-DNA adducts in Ill., E. (1987) Characterization of a human squamous two human ovarian carcinoma cell lines in vitro. Cancer carcinoma cell line resistant to cis-diaminedichloroRes. 51: 4557-4564. platinum(II). Cancer Res. 47: 388-393. SHERMAN, M. L., DATTA, R., HALLAHAN, D. E., TEICHER, B. A., HERMAN,T. S., HOLDEN,S. A., WANG, Y., WEICHSELBAUM,R. R. and KUFE, D. W. (1990) Ionizing PFEFFER, M. R., CRAWFORD,J. W. and FREI, E. (1990) radiation regulates expression of the c-jun protoTumor resistance to alkylating agents conferred by oncogene. Proc. natn. Acad. Sci. U.S.A. 87: 5663-5666. mechanisms operative only in vivo. Science 247: SHERMAN,S. E. and LIPPARD,S. J. (1987) Structural aspects 1457-1461. of platinum anticancer drug interactions with DNA. TEICHER, B. A., HERMAN,T. S, SHULMAN,L., BUBLEY,G., Chem. Rev. 87: 1153-1181. COLEMAN,C. N. and FREI,III, E. (1991a) Combination of SHmANUMA,M., KUROKLT. and NosE, K. (1987) Inhibition etanidazole with cyciophosphamide and platinum comof proto-oncogene c-fos transcription by inhibitors of plexes. Cancer Chemother. Pharmac. 28: 153-158. protein kinase C and ion transport. Fur. d. Biochem. 164: TEICHER,B. A., HOLOEN,S. k., HERMAN,T. S., SOTOMAYOR, 15-19. E. A., KHANDEKAR,V., ROSBE, K. W., BRANN, T. W., SHIONOYA,S., Lu, Y. and SCANLON,K. J. (1986) Properties KORBUT,T. T, and FRELE. (1991b) Characteristics of five of amino acid transport systems in K562 cells sensitive human tumor cell lines and sublines resistant to cisand resistant to cis-diaminedichloroplatinum(II). Cancer diaminedichloroplatinum(II). Int. J. Cancer 47: 252-260. Res. 46: 3445-3448. TERHEGGEN,P. M. A. B., EMONDT,J. Y., FLOOT, B. G. J., DIJKMAN, R., SCHRIER,P. I., DEN ENGELSE,L. and BEGG, SKLAR, M. D. (1988) Increased resistance to cis-diaminedichloroplatinum(II) in NIH 3T3 cells transformed by ras A. C. (1990) Correlation between cell killing by cisoncogenes. Cancer Res. 48: 793-797. diaminedichloroplatinum(II) in six mammalian cell lines SORSCHER, D. H. and CORDEmO-STo~m, M. (1991) Gene and binding of a cis-diaminedichloroplatinum(II)-DNA replication in the presence of aphicidolin. Biochemistry antiserum. Cancer Res..50: 3556-3561. 30: 1086-1090. TIMMER-BOSSCHA, H., HOSPERS, G. A. P., MEIJER, C,, SPEIZER,L. A., WATSON,M. J., KANTER,J. R. and BRUNTON, MULDER,N. H., MUSK1ET,F. A. J., MARXam,I. A., UGES, D. R. A. and DE Vries, E. G. E. (1989) Influence of L. L. (1989) Inhibition of phorbol ester binding and protein kinase C activity by heavy metals. J. biol. Chem. docosahexaenoic acid on cisplatin resistance in a human 264: 5581-5585. small cell lung carcinoma cell line. J. natn. Cancer Inst. STEIN, B., RAHMSDOgF,H. J., STEF~N, A., LtTFIN, M. and 81: 1069-1075. HEgRLICH, P. (1989) U.V.-induced DNA damage is an TOKUI, T., INAGAKI, M., NISHIZAWA, K., YATANI, R., intermediate step in U.V.-induced expression of human KUSAGAWA, M., AJIRO, K., NISmMo~, Y., DATE, T. and MATSUKAGE, A. (199 I) Inactivation of D N A polymerase immunodeficiency virus type 1, collagenase, e-fos and // by /n vitro phosphorylation with protein kinase C. metallothionein. Mol. Cell. Biol. 9: 5169-5181. J. bioL Chem. 266: 10820-10824. SWAT, P., CHAN, P. K. and StATER, L. M. (1989) CycloTONEY,J. H., DONAmm, B. A., KELLErT,P. J., BRUHN,S. L., sporin A and verapamil enhancement of daunorubicinESSlGMANN,J. M. and LIPI'ARD,S. J. (1989) Isolation of produced nucleolar protein B23 transloeation in cDNAs encoding a human protein that binds selectively daunorubicin-resistant and -sensitive human and routine to DNA modified by the anticancer drug c/s-diaminetumor ceils. Cancer Res. 49: 677--680.

406

K. J. SCANLONet al.

dichloroplatinum(II). Proc. natn. Acad. $ci. U.S.A. 86: 8328-8332. Tmyro~, T. R. and H I C ~ N , J. A. (1990) How to kill cancer cells: Membranes and cell signaling as targets in cancer chemotherapy. Cancer Cells 2: 95-105. TSAI, C.-M., GAZDAR, A. F., VENZON, D. J., STmNaFJtO, S. M., DEDRICK, R. L., MULSmNE, J. L. and KRAMER, B. S. (1989) Lack of #I vitro synergy between etoposide and c/s-diaminedichloroplatin(II). Cancer Res. 49: 2390-2397. TWENTYMAN, P. R., FOX, N. E. and WHITE, D. J. G. (1987) Cyclosporin A and its analogs as modifiers of adriamycin and vincristine resistance in a multi-drug resistant human lung cancer cell line. Br. J. Cancer 56: 55-57. VANHAMME,L. and SZPIRER,C. (1987) Methionine metabolism defect in cells transfected with an activated HRAS1 oncogene. Exp. Cell. Res. 169: 120-126. VAZQUEZ-PADUA, M. A., STARNES, M. C. and O-raNG, Y.-C. (1990) Incorporation of Y-azido-Y-deoxythymidine into cellular DNA and its removal in a human leukemic cell line. Cancer Commun. 2: 55-62. VEELKEN, H., TYCKO, B. and SKLAR, J. (1991) Sensitive detection of clonal antigen receptor gene rearrangements for the diagnosis and monitoring of lymphoid neoplasms by a polymerase chain reactionmediated ribonuciease protection assay. Blood 78: 1318-1326.

WALKER, M. C., POVEY, S., PAR~NOTON, J. M., RIDDLE, P. N., KNUECHEL, R. and MAST, S, J. R. W. (1990) Development and characterization of cisplatin-resistant human testicular and bladder tumor cell lines. Eur. J. Cancer 26: 742-747. WALL~r.R, K. E., D~GR~oRIO, M. W. and LI, O. C. 0986) Hyperthermic potentiation of c/s-diaminedichloroplatinum(II) cytotoxicity in chinese hamster ovary cells resistant to the drug. Cancer Res. 46: 6242-6245. WANO, T. S.-F. (1991) Eukaryotic DNA polymerases. A. Rev. Biochem. 60: 513-552. WEINBERG,R. A. (1989) Oncogenes, antioncogenes and the molecular basis of multistep carcinogenesis. Cancer Res. 49: 3713-3721. WIDEN,S. G. and WILSON,S. H. (1991) Mammalian fl-polymerase promoter: Large-scale purification and properties of ATF/CREB palindrome binding protein from bovine testes. Biochemistry 30: 6296-6305. WOLF, C. R., HAYWARD, I. P., LAWRIE, S. S., BUCKTON, K., MCINTYI~, M. A., ADAMS, D. J., LEWIS, A. D., SCOTT, A. R. R. and SMYTH, J. F. (1987) Cellular heterogeneity and drug resistance in two ovarian adenocarcinoma cell lines derived from a single patient. Int. J. Cancer 39: 695-702. ZITTOUN, J., MARQImT, J., DAVID, J. C., MANmY, D. and ZXTTOUN, R. (1989) A study of the mechanisms of cytotoxicity of Ara-C on three human leukemic cell lines. Cancer Chemother. Pharmac. 24: 251-255.

Cisplatin resistance in human cancers.

Cancer chemotherapeutic agents primarily act by damaging cellular DNA directly or indirectly. Tumor cells, in contrast to normal cells, respond to cis...
2MB Sizes 0 Downloads 0 Views