<<

[CANCERRESEARCH56, 5533-5546, December 15, 1996] Special Lecture

Beyond DNA Cross-Linking: History and Prospects of DNA-targeted Cancer Treatment—Fifteenth Bruce F. Cain Memorial Award Lecture'

Kurt W. Kohn2 Laboratory of Molecular , Division of Basic Sciences, National Cancer Institute, Bethesda, Maryland 20892

Abstract tional tri(2-chloroethyl)amine revealed an unusually selective inhibition of reproduction in yeast (relative to a lesser The origin of cancer can be traced to the wartime potency of inhibition of respiration), and this selective inhibition of discovery ofthe lymphotoxic action ofnitrogen mustards These and other reproduction was lost upon hydrolysis of the mustard (10). The bifunctional agents were later found to produce various types of DNA cross-links, and some of these agents continue to be mainstays of current significance of this clue to an alkylation chemistry, however, was not therapy. The cellular pharmacologyofthese wasStUdiedextensively yet appreciated. During this time period, was applied during the @970sand 1980s by means of DNA filter elution methodology topically to tumors or injected intratumorally, mainly in animal ex In the course ofthese investigations, DNA were discovered periments, but a few clinical cases were included, and some local to be targets of and several other classes of anticancer tumors were eradicated (1 l—l3).@ Intravenous mustard gas inhibited drugs. DNA cross-linkers and topolsomerase blockers have generally the growth, and even caused regression, of implanted tumors in rats similar cytotoxic mechanisms, which depend on DNA damage detection, (13). DNA repair, arrest, and cell death by . The molecular Clinical antitumor trials with nitrogen mustard, however, were control of these processes, involving oncogenes and tumor suppressor genes, is being revealed by current research. Cancer cells often have begun only after a later military disaster during World War II, when defects within these control systems, and these defects may confer selective evidence pointedto a selective lymphotoxic action by mustardgas in sensitivity to appropriately designed therapy. Panels of human tu humans (14). The wartime studies of mustard gas and nitrogen mus mor cell lines may serve to link the molecular defects with specific drug turd were shrouded in secrecy and were reported in summary after the sensitivities. Such correlations could guide the selection of drugs for war (15—21). Gilman and Philips, in their landmark paper in 1946 therapy based on molecular diagnosis of individual tumors. (15), surmised correctly the alkylation chemistry of the sulfur and nitrogenmustardsand summarizedthe biological findings in experi Introduction mental animals and humans. They cited the work of many researchers The era of cytotoxic anticancer treatment is now just 50 years old, in the United States and Great Britain whose wartime reports were having begun with the discovery of the antitumoraction of bifunc still classified and unavailable in the open literature. Evidence for tional alkylating agents, some of which remain to this day among the actions on chromosomes was reportedin 1943 in documents of the most useful anticancerdrugs. The key mechanistic feature of these Chemical Board,Ministryof Supply [London;cited by Elmoreet a!. and related drugs is their ability to cross-link DNA. This somewhat (20)].Studiesofthereactionmechanismsofnitrogenmustardwere personalreview will trace our understandingof DNA cross-linking reported in 1946 by Joseph S. Fruton, Max Bergmann, Calvin Golum agentsfromtheirdiscovery50 yearsago to theirrole in the currentera bic, and their coworkers in a series ofeight papers (J. Org. Chem., 11: of molecular biology and will consider also the cell biologically 518—591),andthe alkylation mechanismwas defined (22). related DNA blockers. Other historical details were The clue that led to the first extensive therapeutictrial against included in another recent review (1). was the lymphotoxic effect noted in victims of the World War II tragedy at the Italian port of Ban. The events at Ban were kept History hushed up for many years but were eventually described in detail by Infield (14). The exposure to mustard gas on that occasion differed It is one of those paradoxes of our chaotic world that the origin of from the usual World War I experience in that many of the victims cancer chemotherapy traces back to the use of poison gas in war (Fig. received prolonged exposure over much of the body surface to low 1A). As early as 1917, mustard gas was noted to suppress the bone concentrations of the lipid-soluble alkylating agent dissolved in fuel marrow(2), although its cytotoxic action was attributedto the pro oil. It was the aftermath of a surprise German air attack that destroyed duction of hydrochloricacid in hypotheticalintracellularregionsthat a large number of ships in the crowded harbor, including among them were presumed to be hypersensitive to acidity (3—5).Thebone mar an American ship that was carrying mustard gas bombs. (The Allies row in human casualties showed suppression of granulocytic compo nents, involving not merely local necrosis, but inhibited regeneration evidently brought in poison gas on the possibility that the other side, (6, 7). Additionally,thrombocytopeniawas noted in these early stud in desperation,mightuse such weapons.) Severalfactorscombinedto ies. Intravenous administration of small doses of mustard gas in yield evidence of a specific lymphotoxic action of mustard gas that animals produced leukopenia and impaired antibody response, and was the basis for the clinical trials: the unusual nature of the exposure; these effects were thoughtto resemblethe effects of ionizing radiation the improved pathology techniques; and, very significantly, the in sight and devotion of the medical and research staff who investigated (8, 9). Between the world wars, studies of mustard gas and of the trifunc the event at the time (14). Nitrogen mustards were easier to handle than mustard gas because Received 9/16196;accepted 10/15/96. I Presented at the 87th Annual Meeting of the American Association for Cancer 3 Adair and Bagg (1 1) state that the antitumor testing of mustard gas was suggested to Research, Apnl 24, 1996, Washington, D.C. themin 1927 by Dr.JamesEwing. who hadbeen impressedby the peculiarand specific 2 To whom requests for reprints should be addressed. Phone: (301) 496-2769; Fax: nature of mustard gas burns while serving with the United States Army Medical Museum (301) 402-0752; E-mail: [email protected]. duringthe war. 5533

Downloaded from cancerres.aacrjournals.org on September 27, 2021. © 1996 American Association for Cancer Research. HISTORYAND PROSPECTSOF DNA-TARGETEDCANCERTREATMENT A B

I

r@

C L@@C I C CI

CH3 Fig. 1.A, gas shell explosion somewhere in France, 1918 (201). B, the first published radiograms of tumor regression following systemic chemotherapy: Hodgkin's disease treated with nitrogenmustard(17). Left,before treatment;right,9 weeks afterstartof treatment.Structuresof mustardgas (A) and nitrogenmustard(B) are also shown.

they could be stored as stable crystalline hydrochiorides; upon neu activity in injured or hormonally stimulated tissues. It thus became the tralization, the nitrogen mustard free base had similar chemical and first known chemical that, like ionizing radiation, produced mutations biological properties to mustard gas (17, 24). Nitrogen mustards, and chromosome damage (19). Mustard gas was presumed to damage therefore, were used in the first trials of systemic therapy. Immedi the genetic material, although the chemical nature of the genes was ately after the war, the first results with nitrogen mustards in the not yet established. The mutagenic properties of mustard gas were treatment of neoplastic disease (chiefly , , and discovered in 1942 but were not made public until after the war (19). allied conditions) were summarized in an Official Statement to the Auerbach et al. (19) in 1947 inferred that the mustards “might medical profession by Cornelius P. Rhoads as Chairman of a Com combine with the materials composing the gene.― mittee of the National Research Council in New York (18). The full Some of the early work yielded evidence that remains interesting published reports of the first treated patients soon followed (16, 17). even now in the modern light of DNA damage and repair mechanisms. Although tumors sometimes regressed (Fig. 1B), the regression was Mustard gas, unlike ionizing radiation, produced delayed mutations, only temporary, and treatment was limited by hematological toxicity. as evidenced by mutant Drosophi!a males that produced mosaic Only hematological tumors responded. It was not clear whether ni offspring (19). The chromosome breaks produced by nitrogen mustard trogen mustards were superior to radiation therapy, although radiore may occur preferentially in regions of heterochromatin (25), suggest sistant tumors sometimes responded to the new chemotherapy (17). It ing a selectivity of action that might still be profitably pursued. is remarkable how many of the essential features of these drugs were The concept of cross-link formation (although not specifically of discerned correctly at this early date. DNA) emerged from the observation that the characteristic biological Many of the investigators who were involved in the nitrogen actions of the mustards required the presence of at least two alkylating mustard studies during and immediately following World War H went groups in the molecule (23, 25, 26). In 1949, Goldacre et aL (23) on to become leaders in cancer chemotherapy, hematology, and phar envisioned that “thetwo[haloalkyll groups are required to permit the macology: Alfred Gilman, Frederick S. Philips, Louis S. Goodman, molecule to react at two distinct points, lying either on a single surface Maxwell M. Wintrobe, William Dameshek, Cornelius P. Rhoads, or fibre or, more especially, on two contiguous fibres.―Manynitrogen Joseph H. Burchenal, C. Chester Stock, and David A. Karnofsky. mustard derivatives were tested for their ability to prolong the survival Since nitrogen mustard is a simple chemical structure, easily ame of mice bearing transplanted leukemia, but only those that had at least nable to variation, hundreds of derivatives were tested, but the hope of two alkylating groups were active (27). The double-stranded structure curative therapy was unrealized. Nevertheless, some pharmacological of DNA and its way of replicating, as described by Watson and Crick improvements were obtained in the form of , mel 3 years later, made DNA a natural candidate for cross-linking target. phalan, and , which remain prominent in modem cancer An experimental test of the notion of DNA interstrand cross-linking chemotherapy. Attempts to progress by way of understanding of was made possible by the concept of DNA helix-coil transition, mechanisms were limited by lack of basic knowledge and lack of developed by Marmur and Doty (28) and Doty et a!. (29) in 1959. The powerful methodology. helix-coil transition (also called “DNAmelting―)was found to be rapidly reversible, but only up to the end of the transition, at which Cross-Links between cDNA Strands point the complementary strands were presumed to have separated In the early search for chemical mutagens, mustard gas was tested completely and could no longer reassociate rapidly. Separated strands because of observations that pointed to its interference with mitotic could, however, reassociate slowly with the concentration dependence 5534

Downloaded from cancerres.aacrjournals.org on September 27, 2021. © 1996 American Association for Cancer Research. HISTORY AND PROSPECTS OF DNA-TARGETED CANCER TREATMENT

kinetics as a function of time exhibited a lag, consistent with the 6@iMHN2 requirement for a sequence of two reactions, as would be expected in the formation of a cross-link. Similar methodology revealed DNA interstrand cross-link formation and repair in bacterial cells (31). The site on DNA that is most readily alkylated was determined by 20' Brookes and Lawley to be the guanine-N7 position (32, 33). These investigators also detected nitrogen mustard-linked guanine dimers that could have come from interstrand or intrastrand cross-links. 30' Several other classes of anticancer drugs also proved capable of DNA cross-linking, including mitomycin (34), (35), and chloro 40' ethylnitrosoureas (36). 50' DNA Damage in Mammalian Cells The fact that the reaction of a single nitrogen mustard molecule could prevent the denaturation of an entire DNA molecule suggested 60' a way to measure interstrand cross-links in biological experiments. In cells treated with pharmacologically relevant drug doses, however, the 2K extent of interstrand cross-linking was only on the order of 1 per million bp, and DNA of this size could not be isolated by the usual D N methods. To carry out DNA cross-linking studies in cells and tissues, Fig. 2. Evidenceof DNA interstrandcross.linking.EquilibriumsedimentationinCsC1 therefore, novel methodology was needed. of bacterial DNA treated with various concentrations of nitrogen mustard (HN2) and then A solution to this problem came out of an unexpected observation briefly exposed to 0.02 NNaOH followed by neutralization. Native (N) and denatured (D) in 1971 while we were studying newly replicated DNA that we DNAfonneddistinctbandsthatcouldbequantitated.HN2wasseentoinhibitthestrand separation of a dose-dependent fraction of DNA molecules. thought might be associated with structural components of the nucleus that could be separated by filtration (37). When nuclear lysates were subjected to mild controlled shearing and then dripped through mem of bimolecular kinetics. The anticipated effect of an interstrand cross brane filters, we found newly replicated DNA to be selectively re link was that the linked strands would reassociate rapidly with uni mined. The ratio of newly replicated relative to bulk DNA was molecular kinetics. In 1961, I had the opportunity to work in Paul consistently elevated in the filter-retained fraction and gradually re Doty's laboratory and to test this idea using the cutting-edge technol turned to equivalency during a chase period. Thus, newly replicated ogy of the day: the analytical ultracentrifuge with UV optics. Brief DNA appeared to be part of a detergent-resistant complex that could exposure to 0.02 N NaOH normally denaturedall of the DNA (i.e., be retained on filters. The big surprise came when we used NaOH as converted all of the DNA to single strands). However, if the DNA was the lysis solution. Contrary to what we had seen with various deter pretreated with nitrogen mustard, a fraction of the DNA molecules did gent recipes, newly replicated DNA was now enriched in the flow not denature (Fig. 2). The conversion of denaturable to undenaturable through fraction of the lysate. We surmised that short nascent DNA DNA exhibited first-order kinetics with respect to the amount of single strands, released by alkali, could pass through filters, whereas nitrogen mustard bound to DNA, indicating that the reaction of a intact mature DNA was retained. Using X-rays to generate single single molecule of nitrogen mustard could prevent the strand separa strand breaks, we confirmed that shortened DNA single-strand seg tion of an entire DNA molecule of up to 30 kbp (30). Moreover, the ments passed through filters selectively in alkali. Using a peristaltic

A

0

Fig. 3. Kinetics of DNA alkaline elution (48). A, cells labeled with ‘4C.thymi&newereexposed to various doses of X-ray at 0'C then lysed with detergent on filters, and the DNA was eluted at pHl2.0 (open symbols) or pHl2.8 (filled symbols). I B, dependence of apparent elution rate constant on X-ray dose. Different symbols in B refer to 2 [ode pendent sets of experiments.

10 HOURS OF ELUTION X-ray Dose (R)

5535

Downloaded from cancerres.aacrjournals.org on September 27, 2021. © 1996 American Association for Cancer Research. HISTORY AND PROSPECTS OF DNA-TARGETED CANCER TREATMENT

pump to control the flow of alkaline solution through the filter, we 1.0 found that the system could be made surprisingly sensitive and robust t@ with respect to the size of the DNA single-strand segments. , ‘Q

It was possible to adapt the DNA filter elution technique to a @‘ @ . variety of DNA lesions (38—42). Assays were developed for single ‘t@ b@ strand breaks, double-strand breaks (43), base-free sites (44), inter . b @ strand cross-links, DNA-protein cross-links (45), and protein-associ Normal Normal @Jy ated DNA strand breaks (46, 47). 0.5 Alkaline elution of DNA single strands exhibited a simple log linear behavior with an apparent rate constant that was proportional to Lu

single-strand break frequency (Fig. 3; Ref. 48). Moreover, the elution U.. rate was independent of the number of cells loaded (up to the point of z 0 A C clogging the filter) and nearly independent of flow rate. This was 0.3 surprising in view of the high and variable viscosity in the filter pores Lu 1.0 through which the DNA was passing. Presumably, there are mutually @.1 opposing factors that tend to cancel each other out and thus give the z

observed simple and robust behavior. After optimizing the parameters U- 0 of the system, we soon had in our hands a method that could quantitate z biologically relevant levels of DNA single-strand breaks in cells and 0 I- tissues. C-) From the dependence of alkaline elution kinetics on time and X-ray dose and the fact that X-rays produce strand breaks that are distributed U- randomly, we deduced the behavior of a homogeneous size population of DNA single strands (42). The result seemed at first surprising, because the theoretical elution rate for a population of equal-size strands was a step function: the probability of release from the filter is constant up to a critical time, when all the strands have been released (Fig. 4). If the strands are shorter, it takes less time to reach the total elution point. Physically it would be as if the strands are gliding at a constant rate over a fulcrum, and the time at which a given strand is released depends on how that strand happened to be placed 5 10 15 5 10 15 relative to a fulcrum when the strand was initially deposited on the HOURSOFELUT1ON filter (Fig. 4, inset A). Actually the situation must be more complex, Fig. 5. DNA single-strand breaks during nucleotide excision repair in Uv-irradiated as suggested by Fig. 4, inset B. human fibroblasts, as revealed by DNA alkaline elution. Open symbols, internal controls One of our first applications of the alkaline elution method was showing the elution of DNA from cells irradiated with a standard dose of X-rays. Filled symbols:A, normal human fibroblasts, untreated; B, normal human fibroblasts 4 mm after to nucleotide excision repair. Single-strand break formation was an Uv irradiation;C,xerodermapigmentosumcells(complementationgroupA),untreated; expected part of the repair process, but it had not been clearly D, xeroderma pigmentosum cells 4 mm after UV irradiation. The result showed that demonstrated in mammalian cells. The high sensitivity of the normal cells but not xeroderma pigmentosum cells produce transient DNA single-strand breaksduringnucleotideexcisionrepair.Thisexperimentwascarriedout by AlbertJ. alkaline elution method allowed this type of observation to be Furnace, Jr., in our laboratory in 1975 (49).

A B made clearly (Fig. 5; Ref. 49). After irradiation with UV light, normal human fibroblasts exhibited a wave of single-strand breaks, n=1 k which peaked a few minutes following radiation and then slowly subsided. Xeroderma pigmentosum cells, which were thought in

C capable of nucleotide excision repair, exhibited no UV-induced w DNA strand-break signal. Lu @n=2 DNA-Protein Cross-Links z k/2 --- - 0 The UV repair study led in an unexpected direction. When we checked whether the sensitivity of our strand break assay might have -J I n4 Lu changed during the repair process, we were surprised to find that the k/8 --@-H----f---@---@increased elution caused by a standard dose of X-ray was reduced during the repair period (50). We suspected that this might be due to 1/k 2/k 4/k DNA-protein cross-links, because UV was known to produce such ELUTIONTIME(t) cross-links, and cell proteins might adsorb to the filters and prevent the elution of protein-linked DNA strands. We found that this was Fig. 4. Theoretical alkaline elution behavior of homogeneous populations of DNA indeed what was happening and, furthermore, that xeroderma cells strands of a given size. k is the elution rate of strands of unit length (n 1). Strands of twice that length would elute at one-half the rate k, and so forth. In each case, the elution were incapable of repairing the UV-induced DNA-protein cross-links rate would be constant up to the time when all the DNA has eluted, at which point the (50). This indicated that nucleotide excision repair acted not only on elution rate would fall to zero (42). Inset A, simple model of a DNA strand gliding at a constant rate over the boundary between adjacent filter pores; inset B, a more realistic base damage but also on DNA-protein cross-links. model of a DNA strand gliding between several filter pores. We were able then to devise a very sensitive andquantitativeassay 5536

Downloaded from cancerres.aacrjournals.org on September 27, 2021. © 1996 American Association for Cancer Research. HISTORYAND PROSPECTSOF DNA-TARGETEDCANCERTREATMENT

2@ I I

I .@

A 0.5 HN2 Concentration 4@M)

Fig. 6. DNA interstrand cross-link assay. The alkaline elution assay was modified by using the less adsorptive polycarbonate filters, treating the cell lysate on the filter with proteinase K, and adding dodecyl sulfate to the eluting solution. A, increased elution rates at various doses of X-rays. B, reduced elution rates due to treatments with various doses of nitrogen mustard The cells were irradiated with 300 rad at OC before elution. C, linearization of the cross-linking effect with respect to drug concentration. Cross-link/break parameter = ‘@/@l—R0)/(l —R)l —I, where R,, and R are the fractions of DNA retained by control and drug-treated cells, respectively (39). Different symbols in C refer to independentsetsof experiments.

for DNA-protein cross-links (45), and this was to become important in increased DNA alkaline elution with an agent that had been found to an unexpected way. produce DNA strand breaks by other techniques. We knew that DNA-protein cross-links could obscure the detection DNA Interstrand Cross-Links in Mammalian Cells of DNA strand breaks, but we had no reason to believe that doxoru The alkaline elution rate of DNA single-strand segments from bicin would produce DNA-protein cross-links. By that time, however, X-irradiated cells could be reduced by either DNA-protein cross-links we had worked out the conditions to remove the effects of DNA or DNA interstrand cross-links. Because bifunctional alkylating protein cross-links in the assay. When we applied deproteinizing agents generally produce both types of lesions, a specific assay for conditions to , we were astounded by the results (46, 47). interstrand cross-links required a way to eliminate the effects of the We now saw a classic alkaline elution profile indicative of randomly DNA-protein cross-links. We soon learned how to prevent protein distributed stand breaks; the pattern looked as if the cells had been adsorptionto the filters, andthis allowed us to devise a specific assay X-rayed. The DNA elution was log-linear, and the apparent rate for interstrand cross-links suitable for pharmacological studies in constant of elution was proportional to drug concentration over a mammalian cells (Fig. 6; Ref. 39). pharmacologically relevant dose range (Fig. 7). We applied this methodology to cell pharmacology studies of a That meant that doxorubicin did produce DNA strand breaks, but it variety of DNA cross-linkers, including nitrogen mustards (51—56), also produced DNA-protein cross-links, and the latter completely platinumcomplexes (53, 55—62),(63), andchloroethylmtro obscured the former in the nondeproteinized assay. For randomly soureas (51, 64). The latter compounds were particularly interesting distributed DNA-protein cross-links to completely obscure the DNA because of their manner of producing cross-links slowly after alkyla strand break signal, the DNA-protein cross-links would have to be tion at guathne-O@ positions and because of the inverse relationship present in large excess. Because our assays were quantitative, we were we found between cross-linking and the activity of the guanine-06- able to determine the relative frequencies of the two types of DNA alkyltransferase repair enzyme in different cell types (65). Drug lesion. We found that the frequencies of the single-strand breaks and sensitivity was dependent, in part, on the extent of cross-link forma DNA-protein cross-links were equal within experimental error. We tion (65—67).Further clinical development of new chloroethylating concluded, therefore, that the locations of the two types of lesion agents was proposed (68—70). could not be random (Fig. 7, top left). Rather, there must be one DNA-protein cross-link within each DNA single-strand segment. Protein-associated DNA Strand Breaks: An Effect of There was only one plausible way for this to occur, which was that the Topolsomerase-targeted Drugs proteins were linked consistently either to the 3' or the 5' termini of An unanticipated development occurred in 1978 in the course of the breaks (Fig. 7, top right). This suggested that the linked protein alkaline elution studies of doxorubicin. Because doxorubicin had been might be an enzyme that produced the break. There was at the time reported to produce DNA strand breaks in alkaline sucrose gradient precedent for this type of mechanism in J. C. Wang's nicking-closing studies, we anticipated no difficulty in demonstrating these breaks by activity ofE. coli w protein (71), Blair and Helinski's DNA relaxation alkaline elution. However, using treatments that should have produced enzyme associated with colicinogenic plasmids (72), and Champoux DNA strandbreaks,we saw no increasein DNA elution whatsoever. and Dulbecco's DNA untwisting enzyme of mammalian cells (73, This was puzzling because it was the first time we failed to find 74), enzymes that we now call DNA topoisomerases. Several types of 5537

Downloaded from cancerres.aacrjournals.org on September 27, 2021. © 1996 American Association for Cancer Research. HISTORY AND PROSPECIS OF DNA-TARGETED CANCER TREATMENT

5,-.. __...3' 3,--- __..5' 3,..- _

Fbndom model Bound-to-one-terminus model

Fig. 7. Protein-associated DNA strand breaks pro duced in mouse leukemia cells by doxorubicin (@i concentration as indicated; treatment time, I h). Depro teinizing alkaline elution conditions were used. With out deproteinization, no increase in elution occurred. The DNA of treated cells was prelabeled with ‘4C- thymidine. An intemal standard of I3Hlthymidine-la 0 beled cells irradiated with 300 rad was added, and this w served as a normalized time scale on the horizontal axis. This experiment was carried out by Warren E. Ross in our laboratory in 1978(47). Top left, randomly distributed single-strand breaks and DNA-protein cross-links. Top right, proteins linked consistently to one terminus at each single-strand break.

DNA intercalating drugs were found to produce protein-associated enzyme induced came from quantitative studies in cells. Several DNA strand breaks, as we then called these lesions, and were there observations were consistent with an enzymatic origin (75): (a) after fore hypothesized to be topoisomerase blockers (47, 75—78). addition of drug, the lesion frequencies climbed to a steady-state level Support for the idea that protein-associated DNA strand breaks are that was dependent on drug concentration (Fig. 8); (b) the level of lesions showed saturation kinetics with increasing drug concentration, indicating that there was a limiting factor governing the number of lesions that were formed; (c) the lesions reversed rapidly upon drug >- U removal; and (cOboth the formation and the reversal of lesions were z strongly temperature dependent. In all of these experiments, the strand 0 LU breaks and DNA-protein cross-links appeared and disappeared to U. gether, supporting the idea that they were different aspects of the same

LU lesions. Even more telling, the formation and reversal of breaks was demonstrable in isolated cell nuclei without AlP, showing that the 3 Z?r reversal was not a DNA repair process (79—82). 4 In attempting to identify the responsible enzyme, we found that the activity could be extracted from isolated nuclei with 0.35 MNaC1 and that the activity could be restored by adding the extract back to the isolated nuclei (83). Using this assay, the extract was fractionated to isolate the active component. Specific identification of the enzyme, however, had to await the characterization of topoisomerase II by Leroy Liu et a!. (84, 85), and it was in Liu's laboratory that drug effects on the purified enzyme were first demonstrated (86—88).We verified that the component responsible for the drug effects in the filter assays was in fact topoisomeraseH (89). TIME FOLLOWINGDRUG ADDITION (minutes) In addition to DNA intercalators, the epipodophyllotoxin deriva Fig. 8. Kinetics of protein-associatedbreak formation and reversal in mammaliancells. tives etopside (VP16) and (VM26), which are not strong ThisexperimentwascarriedoutbyLeonardA.Zwellinginourlaboratoryin1981(75)and DNA binders,were found to produceprotein-associatedDNA strand used the dnig m-AMSA. also known as , which gave very clear and specific results. [m-AMSA was developed by Bnice F. Cain, who is commemoratedby these lectures(202, breaks and were shown to be topoisomerase-targeted drugs (90—94). 203).]Cellswereexposedtotheindicatedconcentrationsofdnigfor60miii,atwhichtime It is instructive that these chemical modifications of the natural the dnig was washed away and the cells were incubated further in the absence of dnig. At various times, aliquots of cells were subjected to alkaline elution under deproteinizing product, , retain the cytotoxic potency of the parent conditionsto determinesingle-strandbreak frequencies.Under nondeproieinizingconditions, but change the target from tubulin to topoisomerase H (95—97);nature there was no increase in DNA alkaline elution (data not shown) because of the presence of covalent DNA-protein complexes. Assays for DNA-protein cross-links (data not shown) sometimes exhibits coherence in unexpected ways. indicated a nearly 1:1 ratio of DNA-protein cross-links and single-strand breaks. Whereas topoisomerase H is inhibited by a wide variety of corn 5538

Downloaded from cancerres.aacrjournals.org on September 27, 2021. © 1996 American Association for Cancer Research. HISTORY AND PROSPECTS OF DNA-TARGETEDCANCER TREATMENT

Fig. 9. Stacking model of drug-stabilized topoisomerase H-DNAcomplexes.Basesequencepreferenceforthe location of the siteson the DNAdiffersfordifferentdrugclasses.For doxorubicin and related anthracycines, the strongest base pref erence is for an A at the immediately upstream flank of the DNA strand break, as shown.Ellipticinesprefera T, and prefer a C at this position. Amsacrine prefers an A at theimmediatelydownstreamflankofthebreak.Topoisomerase H is a homodimer conferring diadic symmetry to the complex in which there can be a pair of single-strand breaks separated by a 5' overlap of 4 bp. To simplify the diagram, the two mono mers of a topoisomerase homodimer are shown as separate circles. Drug molecules are represented by filled rectangles.

Ellipt icine VM26

pounds, topoisomerase I thus far has a relatively narrow range of by Yi Fan et a!. in our laboratory4 has indicated recently that a specific inhibitors (98, 99). Chief among these is , a structure with camptothecin stacked against a base pair at a topoi fascinating natural product the discovery, chemical structure, and somerase I cleavage site is in accord with a variety of experimental structure-activity relationships of which were described by Monroe evidence regarding structure-activity relationships (125), direct Wall and Mansukh Wani in the Cain Memorial Award Lecture in interaction between camptothecin derivatives and DNA purines, 1994 (100). The first observations on the effects of camptothecin on especially at the cleavage site (126, 127), and drug-resistant to DNA were actuallymade by SusanHorwitz(101), who presentedthe poisomerase I mutations (1 12). Cain Memorial Award Lecture in 1992. Camptothecin was shown to block purified topoisomerase I specifically (102, 103) and to produce Toward Better Drugs through MOdified Mechanisms protein-associated DNA strand breaks in cells (104). The role of topoisomerases as drug targets in cells was confirmed by the isolation After 50 years, some of the most useful anticancer drugs still and characterization of drug-resistant mutant forms of topoisomerase include DNA cross-linking agents, such as cyclophosphamide and I or H from drug-resistantcells (105—112). cisplatin, and topoisomerase blockers, such as doxorubicin and camp The different types of topoisomerase H blockers were found to tothecin derivatives. One path toward improved drugs would be stabilize DNA-cleavable complexes at different sites in DNA, depend through chemical structure modification based on knowledge of ent in parton DNA sequence (113, 114). The location of the blocked chemical mechanisms. It may be asked, for example, whether the complexes on DNA in reconstituted chromatin depended in part on drugs react in a DNA sequence-dependent manner, whether drugs of nucleosorne positioning (115). In cells, topoisomerase H DNA cleav a given class differ in their preferred DNA reaction sites, and, if so, age sites were found to occur frequently in regions of DNA that are whether such differences have consequences relevant to therapy. nuclear matrix association regions (1 16) and in the c-myc gene when When modified drugs show major alterations in mechanism in the it is actively transcribed (117, 118). Camptothecin-trapped topoi absence of clear information on the biological consequences of such somerase I complexes in vivo tend to occur in transcriptionally active alterations, it may be justified to consider the modified drug as if it regions (119) and in intemucleosomal linkerregions (120). were a new chemical structure and to move toward of the Analysis of the nucleotide sequence contexts at sites of drug modified drug in comparison with its parent. induced topoisomerase H cleavage complexes in purified DNA sug Nitrogen mustards have characteristic nucleotide sequence prefer gested a structural model for the complexes (121). The base pairs that ences for DNA alkylation (128—131),and topoisomerase blockers, as most influenced site selection were those that were situated on the already discussed, also vary in DNA site preferences. However, we immediate flank of a cleavage site. Depending on the drug, this base know very little about how such differences translate into effects on pair could be on the 5' or the 3' flank, and the base preference differed cells. Alkylating agents react with many biomolecules other than for different drug classes (121 , 122). This suggested a model of the DNA. Because DNA is the important target, side effects might be trap@ complex in which the drug molecule stacks against a flanking reduced by minimizing the non-DNA reactions. DNA targeting of base pair at the cleavage site (Fig. 9; Ref. 121). nitrogen mustards has been studied in compounds containing an Similarly, we found for the interaction of camptothecin with acridine moiety capable of DNA intercalation (131, 132). Some of topoisomerase I that the strongest sequence preference was for the these compounds have unusual DNA sequence preferences, and in bases immediately flanking the cleavage site (123, 124). The stacking model can therefore apply also to the camptothecin 4 Y. Fan, Y. Pommier, K. W. Kohn, and J. N. Weinstein. Molecular modeling studies topoisomerase I trapped complexes. A molecular modeling study of the DNA-topoisomerase ternary complex with camptothecin, submitted for publication. 5539

Downloaded from cancerres.aacrjournals.org on September 27, 2021. © 1996 American Association for Cancer Research. HISTORYAND PROSPECTSOF DNA-TARGETEDCANCERTREATMENT

some cases, the monofunctional alkylating derivatives are active, suggesting mechanistic differences worthy of clinical investigation. The relationship between DNA site preferences and cytotoxicity for topoisomerase-targeted drugs is also unknown. In some cell lines, the potency of these drugs correlates with DNA and/or RNA syntheses (133—135),butit is not know how this relates to the genomic sites of the lesions. .5' Extensive structure-activity data on have given a camptothecin coherent picture that may help in the development of improved versions of these drugs (125, 136, 137). Ability to block topoisomer @ ‘I ReplIcatIon ase I was closely correlated with antitumor activity. The features of complex the molecule that are essential for activity include the a-hydroxy structure of ring E and the steric configuration of the asym metric carbon on this ring (position 20; Fig. 10). Positions 9 and 10 on ring A can be substituted without loss of activity, whereas substitution of position 12 on the same ring destroys activity. This suggests that the region of the molecule near position 12 is juxtaposed against the DNA or topoisomerase in the complex, whereas the other side of the Fig.11.Howacollisionbetweenamovingreplicationforkanda camptothecin-trapped molecule, in the vicinity of positions 9 and 10 in ring A and position topoisomerase I complex may result in an irretrievable DNA lesion. 7 in ring B, is accessible to solvent. In searching for improved camptothecin derivatives, substitutions can therefore be made at po covalently with the enzyme or the DNA to make the complexes nearly sitions 7, 9, 10, and sometimes I 1. Several such derivatives are in irreversible (127). These derivatives do form nearly irreversible corn clinical use or under investigation (138). plexes, and they have enhanced cytotoxic potency, which, unlike that What basis in mechanism might there be for improved activity of of other camptothecins, is not reversed by DNA synthesis inhibitors.5 such derivatives? One possibility emerged from the current partial Clinical trials will be needed to determine whether enhanced cytotoxic understanding of how camptothecin produces potentially lethal le potency by this mechanism can translate into improved therapeutic sions in cells. When camptothecin forms a complex with DNA and results. topoisomerase I, one strand of the DNA helix is cleaved, and the enzyme is covalently bound via a tyrosine residue to a 3'-phosphate Toward Better Therapy through Cell Biology terminus. The drug may stabilize this cleaved complex, in part, by forming a labile covalent bond between an opened lactone ring (ring Although we know a great deal about the chemical lesions pro E) and a nucleophilic group on the enzyme (98, 139, 140). Covalency duced by anticancer drugs and about their repair, we still do not know notwithstanding, the complex can dissociate within minutes to re-form much about the origin of the antitumor activity: we do not know how undamaged DNA. A potentially lethal lesion can result if a cleaved cancer cells may be killed selectively or why the drugs work in some complex is encountered by a moving replication fork (141—143), cases and not in others. The emerging detailed knowledge of how cell which could have the consequence of producing a DNA double-strand proliferation, differentiation, and apoptosis are regulated may open break (Ref. 144; Fig. 11). Such lesions are considered to be only new possibilities for therapy. potentially lethal, because they could be subject to repair. How then The physiological responses of cells to a variety of DNA-damaging might structure modification enhance the production of such lesions? drugs or topoisomerase blockers are remarkably similar in that treated A key parameter is the lifetime of the drug-DNA-topoisomerase cells tend to arrest at certain points in the cell cycle and, in some complex: the longer its lifetime, the greater the probability of a fatal cases, to undergo apoptotic death. The cell cycle arrest generally is not interaction with a genomic event, such as a moving replication fork or a manifestation of drug toxicity. Rather, it is a protective mechanism transcription process (135, 142, 145). In support of this picture, ofcell cycle checkpoints in which the initiation of one cell cycle event camptothecin derivatives were found to differ in the rate of revers is inhibited while another is in progress (147—152). ibility of their complexes, and cytotoxic potency was found to in In this context, a cell cycle checkpoint response to DNA damage is crease with increasing lifetime of the complexes (146). With this indicated when there is evidence that (a) treated cells tend to arrest or conceptual underpinning, camptothecin derivatives were prepared become delayed at a specific point in the cell cycle; (b) the arrest can with an alkylating substituent in the 7 position, which could then react be prevented by adding an inhibitor; and (c) prevention of arrest increases cytotoxicity. Thus, a (12 checkpoint is indicated by the findings that (a) tumor 9 7 cells that have sustained DNA damage often arrest in G2 (or very late 10 in S); (b) this arrest can be prevented by caffeine, pentoxifylline or certain protein kinase inhibitors which are not by themselves cyto 11 toxic; and (c) when the G2 arrest is prevented, cell survival is reduced (152—158). A current view of the G2 checkpoint controlling the G2 to M transition in drug-treated or irradiated cells suggests a system of mutually interacting kinases and phosphatases (Fig. 12; Refs. 150, Et 152, 159—164).An interesting feature of this system is that it appears to contain positive feedback loops, such that the components may HO have switch-like behavior and may not constitute a linear causal 0

Fig. 10. Structure of camptothecin. 5 M. Valenti, K. W. Kohn, and Y. Pommier, unpublished observations. 5540

Downloaded from cancerres.aacrjournals.org on September 27, 2021. © 1996 American Association for Cancer Research. HISTORYAND PROSPECTSOF DNA-TARGETEDCANCERTREATMENT

These opposing actions of can account for the conflicting results in different cell systems that have been reported on the effect of p53 on drug sensitivities (157, 173—176).Cells that are inherently apoptosis competent would be expected to exhibit reduced sensitivity to DNA damage when p53 is inactivated by mutation. Lymphoma cells, for example, generally undergo apoptosis very readily. Thus, most p53-mutant Burkitt's lymphoma cell lines were found to have reduced sensitivities to ionizing radiation, the DNA cross-linking agents nitrogen mustard and cisplatin, and the (179). Some carcinoma cells, on the other hand, have weak or nonexistent capabilities to undergo apoptosis, even when they have normal p53 function. In such cases, cells can have increased sensitivity to DNA cross-linking agents when p53 function is disrupted (157). Without p53 function, treated cells then do not arrest in G1, but proceed Mltotic through S and arrest in G2. Survival then becomes increasingly events dependent on an intact G2 checkpoint. When the G2 checkpoint is

Fig. 12. Integrated switch model of the G2 to M phase transition (152). Mitotic events inhibited (e.g., by means of pentoxifylline or by means of a stauro are stimulated by the hypophosphorylated cyclin B-cdc2 complex, which is the active sporine derivative, UCN-Ol), survival is reduced further, and the form of the kinase. During S or 02 phaseScells accumulate this kinase in an inactive form selective toxicity against p53 mutant cells is increased (157, 158). in which the cdc2 component is hyperphosphorylated. The conversion of inactive to active form is accomplished by a phosphatase, cdc25C, the activity of which is enhanced by Another confounding factor in p53 disruption is chromosome in hyperphosphorylation, this being carried out by cydlin B-cdc2. Thus we stability, which arises in cells that cycle unchecked despite persistent have a positive feedback loop that can have switch-like behavior. A small initial stimulus can activate the switch. The switch is kept in check by ongoing DNA replication or repair, DNA damage (183). Chromosome instability increases the heteroge so that when these processes are completed, the switch is thrown. How the switch is neity of cell populations and favors the selection of rapidly growing or controlled is, however, unknown. One controlling component is the weal kinase, which drug-resistant malignant cells. Although p53 represents a well-defined tends to keep cyclin B-cdc2 hyperphosphorylated and inactive. molecular and functional distinction between normal tissues and cer tamtumors,theutilizationofsuchdifferencestoselectivelyeradicate sequence of events (152). The inputs which control the switch and the these tumors by means of chemotherapy is a complex problem re components that sense the presence of DNA damage however remain quiring a detailed understanding of integrated regulatory functions. to be elucidated. p53 affects the transcription of several genes that are believed to In early studies, arrest in G2 was the most prominent effect, because mediate its diverse cell biological effects (184, 185). Included among the tumor cells studied tended to be defective in tumor suppressor p53 these are p2lwafl/cipl and gadd45, which inhibit the G1-S transition function. More recent studies in cells having intact p53 function show and probably also itself (186—188), and bax and bcl-2 that that G@ arrest (arrest prior to the G1 to S transition) also is an affect the threshold for apoptosis ( 189—191).The logic of this system important feature of drug responses (165, 166). may be as follows. Increased levels of active p53 protein lower the Much of the complex regulatory network controlling the G1 to S threshold for apoptosis in all cells that are inherently competent to transition has now been revealed, including its component cyclins, protein kinases and phosphatases, kinase inhibitor proteins, transcrip carry out the process. By increasing the transcription of p2lwafl/cipl, tion factors and their inhibitors, and promoter elements of genes for which blocks cyclin-dependent kinase cdk2, p53 at the same time some of these components (Fig. 13; Refs. 164, 167—169).Itwill be causes E2F activities to fall. Because E2F can cooperate with p53 to challenging to integrate the vast amount of information accumulating enhance apoptosis (192), the reduced E2F activity could prevent cells about various parts of the network. Therapeutic implications currently from undergoing apoptosis despite the lowered threshold. Thus, center on p53, which is defective in about one-half of human cancers Polyak et a!. (178) have recently shown that disruption of p2lwafl/ @ (170, 171), and on the pathway involving cyclin D, tumor suppressor cipl can convert the p53-induced response from arrest to apopto protein p16, and pRb,6 one or another of which is altered in most sis, indicating that p2lwafl/cipl can mediate cell cycle arrest and clinical tumors (Refs. 152, 157, 172—176;Fig. 13). protect against apoptosis. The functional value of this logic could be p53 and other regulatory components that control the G1 checkpoint that cells that inappropriately elevate E2F levels through expression of are also involved in the control of cell death by apoptosis (177, 178). oncogenes or viral genes would be eliminated. [As a countermeasure, Because p53 is so often defective in clinical cancer (171), the effects some oncogenic viruses synthesize products that disrupt p53 function of p53 on drug sensitivities have been of intense recent interest. The (193).] functions of p53, however, are now seen to be multiple and complex (177), and some of its different actions affect drug sensitivity in Studies on a Panel of Human Tumor Cell Lines opposing directions (Refs. 157 and 179; Table 1). When p53 function is disrupted, DNA damage does not cause cells to arrest in G1 (180), Detailed knowledge at the level of molecular and cellular biology and DNA repair capability is reduced (157, 181). The enhancement of may not immediately lend itself to clinical application. One formula DNA repair by p53 may be mediated in part through p53-induced tion of the problem is that is driven by altered expression transcription of GADD45 (182). The effects of p53 disruption in of particular gene products and that such alterations could make cells disturbing cell cycle arrest and DNA repair tend to reduce cell selectively sensitive to particular cytotoxic therapies. The objective is survival. On the other hand, the loss of the role of p53 in apoptosis to develop cytotoxic therapies that would be tailored to tumors of would tend to increase cell survival. certain types defined by their molecular characteristics. Some of the important molecular changes in common cancers may be preserved in

6The abbreviations used are: pRb, retinoblastoma protein; NCI, National Cancer appropriately chosen cell lines. The molecular alterations in human Institute. tumor cell lines could then serve to link preclinical and clinical 5541

Downloaded from cancerres.aacrjournals.org on September 27, 2021. © 1996 American Association for Cancer Research. HISTORYAND PROSPECTSOF DNA-TARGETEDCANCERTREATMENT

Fig. 13. General logic of the control of S phase and apoptosis in cells treated with DNA-damaging drugs. Tumor cells often have defects in one or another component of this regulatory system, and future therapies may be designed to take advantage of these defects (157, 173-175). This regulatory system makes a life-or-death decision for the cell, dictating either cell cycle arrest or apoptosis (178, 188, 204, 205). DNA damage or other forms of genotoxic stress cause p53 to accumulate, mainly as a result of decreased degradation of the protein. p53 causes p2lwafl/cipl to increase due to transcriptional activation. p21 inhibits the function of 01-cyclin-dependent kinases by binding to the cycin-cdk complexes. The cyclin-cdk complexes hyperphosphorylate, and thereby block, the function of pRb. pRb, when not hyperphosphorylated, binds to and blocks E2F transcription factors. E2F stimulates many genes that are required for and that may initiate S phase. p53 also cooperates with E2F to stimulate apoptosis (192), in part through its transcriptional control of the Bax-Bcl2 couple (189—191).Quotation marks indicate that related protein species may function similarly in the same or in parallel pathways. The best defined parallel pathways consist essentially of (a) cyclin D-cdk4, pRb, and E2F1, and (b) cyclin E-cdk2, p107, and E2F4. Tumor suppressor protein p16 also is frequently defective in tumor cells and may play a target role in future therapies (172); it competes with cyclin D for binding to cdk4 and thereby specifically blocks this pathway.

Table 1 Factors affecting cell survival when p53 function is disrupted Decreased cell survival Increased cell survival Impaired cell cycle checkpoint functions Reduced apoptosis Impaired DNA repair Chromosome instability leading to selection of malignant Chromosome instability leading to lethal genotypes or drug-resistant cell populations development of therapies for specific tumors based on molecular tics could guide clinical trials based on molecular diagnosis of the diagnosis. tumors. Thus drugs that act selectively against cell lines having certain The NC! cell screen has generated a large amount of data on cell molecular characteristics could be tested in patients stratified with growth inhibition by many thousands of compounds tested against a respect to a molecular diagnosis of their tumors. panel of 60 human tumor cell lines representing a variety of tissues of One phase of the cell line characterization studies has focused on origin ( 194). Individual cell lines differ greatly in their drug re p53.7 Of 58 cell lines analyzed, 39 were found to be p53 mutant. sponses. However, the pattern of responses of the set of 60 cell lines Most, but not all, of the mutant lines showed elevated basal levels of is strongly correlated with drug action mechanisms (195—197).The p53, and their responses to ionizing radiation were deficient with response patterns, considered as 60-dimensional vectors, have been respect to G@arrest and with respect to inducibility of p2lwafl/cipl, analyzed by means of a variety of statistical tools, including cluster gadd45, and mdm2 transcription. Strong G1 arrest in response to analysis, neural networks, and principle component regression. These ionizing radiation occurred only in cell lines that exhibited strong analyses have been done for large data sets of >40,000 compounds induction of p2lwafl/cipl mRNA. The p53 mutant lines exhibited and for a subset of compounds for which probable mechanisms of statistically less growth inhibition than p53-normal lines to DNA action could be inferred from experimental evidence or chemical cross-linking agents, topoisomerase inhibitors, and structure. Cluster analysis over large sets of tested compounds (198, (l99).@ Apoptosis may therefore be an important factor in drug re 199) corroborates the observation that compounds related by structure sponses in the NC! screening assay. Because there was considerable or by action mechanism usually group together. The coherence of overlap in the sensitivity distributions, factors in addition to p53 these groupings is remarkable in view of the huge size of the data set contribute to the growth inhibition, some of which may be identified and demonstrates the informational power inherent in the behavior as this work progresses. Antimitotic agents, such as taxol and vincris patterns of a large panel of cell lines. tine, on the other hand, did not show any such dependence on p53, To test the ability of the cell screen data to predict mechanism of suggesting that these agents may have an advantage over other stand action, a set of 123 standard drugs were assigned probable mecha ard drugs in the treatment of p53-mutant tumors (l99).@ nisms (196). The assignments were made without prior knowledge of the cell screen results. A neural network, trained on a subset of drugs, Postscript was then able to correctly assign the remaining drugs in nearly all cases. The strong relationship between mechanism of action and cell The painting by Edouard Manet featured on the cover of this issue screen response pattern was verified by other statistical methods of Cancer Research (and Fig. 14) could be viewed in relation to the (198—200). half-century quest for anticancer drugs. The painting, “GareSt. The informational power of patterns encompassing sets of cell lines Lazare,―was brought to my attention in a recent Sunday afternoon is now being extended to include molecular and cellular biology lecture at the National Gallery of Art presented by the art historian characteristics. The 60 cell lines of the NCI screen have been or are Juliet Wilson-Bareau, to whom I am also indebted for an illuminating being characterized with respect to a variety of oncogenes, tumor response to my proposed use of the picture. Her lecture focused on suppressor genes, and cell cycle responses. A variety of correlations “themysterybehind the steam―inthe painting. The steam is billowing are being examined to determine how individual cell lines are related to each other and how molecular characteristics relate to drug re 7 P. M. O'Connor, J. Jackman, I. Bae, T. 0. Myers, S. Fan, D. A. Scudiero, A. Monks, E. A. Sausville, J. N. Weinstein, S. Friend, A. J. Fornace, Jr., and K. W. Kohn. sponses. Assuming that the cell lines retain some of the molecular Characterization of the p53 tumor suppressor pathway in cell lines of the NC! anticancer characteristics that are common in clinical tumors, these characteris drugscreenandrelationshipswithchemosensitivity,manuscriptinpreparation. 5542

Downloaded from cancerres.aacrjournals.org on September 27, 2021. © 1996 American Association for Cancer Research. HISTORY AND PROSPECTS OF DNA-TARGETEDCANCER TREATMENT

of Art, for her help in obtaining slides and permission to reproduce the Manet parnting that appears on the cover of this issue.

References

I. Kohn, K. W. DNA filter elution: a window on DNA damage in mammalian cells. BioEssays, 18: 505—514,1996. 2. Krumbhaar, E. B. Bone marrow changes in mustard gas poisoning. 3. Am. Med. Assoc., 73: 715, 1919. 3. Marshall, E. K., Jr., Lynch, V., and Smith, H. W. On dichioroethylsulphide, the systemic effects and mechanism of action. J. Pharmcol. Exp. Ther., 12: 291, 1918. 4. Smith, H. W., Clowes, 0. H. A., and Marshall, E. K., Jr. On dichloroethylsulphide, the mechanism of absorption by the skin. J. Pharmacol. Exp. Ther., 13: 1, 1919. 5. Lillie, R. S., Clowes, G. H. A., and Chambers, R. On the penetration of dichloro ethylsulphide (mustard gas) into marine organisms, and the mechanism of its destructive action on protoplasm. J. Pharmacol. Exp. Ther., 14: 75—120,1919. 6. Krumbhaar, E. B. Role of the blood and the hone marrow in certain forms of gas poisoning. J. Am. Med. Assoc., 72: 39—41, 1919. 7. Knimbhaar, E. B., and Krumbhaar, H. D. The blood and bone marrow in Yellow Cross gas (mustard gas) poisoning. J. Med. Res., 40: 497—506,1919. 8. Pappenheimer, A. M., and Vance, M. The effects of intravenous injections of dichioroethylsulfide in rabbits, with special reference to its leukotoxic action. J. Exp. Med., 31: 71—94,1920. 9. Hektoen, L., and Corper, H. J. The effects of mustard gas (dichlorethyl suiphid) on antibody formation. J. Infect. Dis., 28: 279—285, 1921. 10. Magne, H., and Remy, P. Les actions toxiques elementaires de quelques corps I•i@;14.1tiniiri@ h\ II@U@it4i \1@nct. (‘i @!.Ii:', 1 ( I, @i@i1 Ii I pc generateurs d'acide chlorhydrique par hydrolyse. Bull. Soc. Chim. Biol., 19: 1092— @@@ l@@@)kt11pt@@rtl@@i1@\@lLI\l@L'l@i1@@ftl1L@ )L' L\L@!l@l. \\Itll 1104, 1937. @ @i!,i;-I@:@/ . h@. I@J@uctJ \1iiict ( @ti t r@i@c 11c@crnc\ I 11 F1ICfli@\ t I 1. Adair, F. E., and Bagg, H. J. Experimental and clinical studies on the treatment of @ n@tlicr. 1t!Nilic \\ l1t\c11lc\cL l)U Rtid 1 @IrtI\@c@.\ttiritI ( IIICF\ I \it. cancer by dichlorethylsuiphide (mustard gas). Ann. Surg., 93: 190—199,1931. \\ t@hiiictn. D@( . H@ @. ii n @lH\t 12. Kabelik, J. Experiences chimiotherapiques sur Ic carcinome d'Ehrlich de Ia souris. Compt. Rend. Soc. Biol., 110: 394—397,1932. 13. Visser, J., and de Vos, J. J. Over den invloed van mosterdgas op kwaadaardige up from locomotives that are passing below into and out of the Paris gezwellen. Geneesk. Tijdschr. Nederlandsch-Indie., 75: 1363—1380,1935. 14. Infield, G. B. Disaster at Bad. New York: Macmillan, 1971. railway station, which is just to the right beyond the bridge. In the 15. Gilman, A., and Philips, F. S. The biological actions and therapeutic applications of upper left-hand corner, one can make out Manet's studio window. The the @-chloroethylaminesand sulfides. Science (Washington DC), 103: 409—415, woman seated on the ledge, with the studio across the tracks of the 1946. 16. Goodman, L. S., Wintrobe, M. M., Dameshek, W., Goodman, M. J., and Oilman, A. broad railway cutting just behind her, served as model in many of Nitrogen mustard therapy. Use of methyl-bis(beta-chloroethylamine hydrochloride Manet's paintings. The child, in a white dress with her back toward us, and tris(beta-chlomethyl)amine hydrochloride for Hodgkin's disease, lymphosar coma, leukemia and certain allied and miscellaneous disorders. J. Am. Med. Assoc., is looking through the steam in the direction of the studio, thus uniting 132:126—132,1946. the two figures whose relationship to each other so puzzled the critics 17. Wilkinson, J. F., and Fletcher, F. Effects of j3-chlorethylamine hydrochlorides in of the time. The billows of steam can be likened to smoke and dust in leukemia, Hodgkin's disease, and polycythemia vera. Lancet, 1946: 540—545,1947. 18. Rhodes, C. P. Nitrogen mustards in the treatment of neoplastic disease. J. Am. Med. other Manet paintings, most notably the smoke from the guns in his Assoc., 131: 656—658, 1946. painting, “TheExecutionof the Emperor Maximilian.―The child, 19. Auerbach, C., Robson, J. M., and Carr, J. G. The chemical production of mutations. with her left arm extended, can be referenced to a symbol of hope in Science (Washington DC), 105: 243—247,1947. 20. Elmore, D. T., Gulland, J. M., Jordan, D. 0., and Taylor, H. F. W. The reaction of other art works, and thus could be looking through the steam in the nucleic acids with mustard gas. Biochem. J., 42: 308—315,1948. hope of seeing a better world beyond the smokes of war. “GareSt. 21. Beanie, J. W., and Howell, L. H. Biological actions and therapeutic applications of Lazare―was painted in 1873, shortly after the Franco-Prussian War, nitrogen mustard. Q. J. Med., New Series., 23: 231—254,1954. 22. Fruton, J. S., Stein, W. H., and Bergmann, M. Chemical reactions of the nitrogen during which Paris was invaded. The era of cancer chemotherapy mustard gases. V. The reactions of the nitrogen mustard gases with protein constit began out of the gases of war; now, as the mists of ignorance clear at uents. J. Org. Chem., II: 559—570,1946. 23. Goldacre, R. J., Loveless, A., and Ross, W. C. J. The mode of production of last, we may look for better therapy through basic science. chromosome abnormalities by nitrogen mustard: possible role of crosslinking. Nature (Lond.), 163: 667—669, 1949. Acknowledgments 24. Ward, K., Jr. The chlorinated ethylamines: a new type of vesicant. J. Am. Chem. Soc., 57: 914—916,1935. It is a greatpleasureto acknowledgemy colleagueswho hadmajorpartsin 25. Loveless, A., and Revell, S. New evidence on the mode of action of “mitotic the studies from my laboratory. Regina A. Ewig worked with me during the poisons.―Nature(Lond.), 164: 938—944,1949. 26. Haddow, A., Kon, 0. A. R., and Ross, W. C. J. Effects upon tumors of various discovery and development of the DNA alkaline elution technique. Major haloalkylarylamines. Nature (Land.), 162: 824—825,1948. contributions to the variations and applications of the technique were made by 27. Burchenal, J. H., and Riley, J. B. Nitrogen mustards: the relationship between Leonard C. Erickson, Albert J. Fomace, Jr., Matthews 0. Bradley, Warren E. chemical structure and chemotherapeutic activity. Cancer Res., 9: 553—554,1949. Ross, andLeonardA. Zwelling.Manyof the earlytopoisomerasestudieswere 28. Marmur, J., and Doty, P. Heterogeneity in DNA. Nature (Land.). 183: 1427—1429, 1959. conducted by Warren Ross, Leonard Zwelling, Donna Kerrigan, Jan Filipski, 29. Doty, P., Marmur, J., Eigner, J., and Schildkraut, C. Strand separation and specific and John K. Minford and by Yves Pommier, who is leading the current studies recombination in DNA: physical chemical studies. Proc. NaIl. Acad. Sd. USA, 46: related to topoisomerase inhibitors in our laboratory. Most of the work on 461—476,1960. chloroethylnitrosoureas was done by Leonard Erickson, Neil Gibson, and John 30. Kohn, K. W., Spears, C. L., and Doty, P. Inter-strand crosslinking of DNA by nitrogen mustard. J. Mol. Biol., 19: 266—288,1966. A. Hartley.The studies of DNA sequence-dependentalkylationincludedthe 31. Kohn, K. W., Steigbigel, N. H., and Spears, C. L. Crosslinking and repair of DNA efforts of William B. Mattes, John Hartley, and Ann Orr. The cell cycle work in sensitive and resistant strains of E. coli treated with nitrogen mustard. Proc. Nail. was and is continumg to be spearheaded by Patrick M. O'Connor. The Acad. Sci. USA, 53: 1154—1161,1965. 32. Brookes, P., and Lawley, P. D. Reaction of mustard gas with nucleic acids in vitro information-intensive approach to human tumor cell line panels and their and in vivo. Biochem. J., 77: 478—484,1960. molecular pharmacology is being developed under the leadership of John N. 33. Brookes, P., and Lawley, P. D. The alkylation of guanosine and guanylic acid. J. Weinstein. Chem. Soc., 1961: 3923—3928,1961. The book by G. B. Infield aboutthe Disasterat Bail was broughtto my 34. Iyer, V. N., and Szybalski, W. A molecular mechanism of mitomycin action: linking attentionbyDr.ShaiIzraeli,who hadreviewedthebookfortheIsraelMedical of complementary DNA strands. Proc. Nail. Acad. Sci. USA, 50: 355—362,1963. 35. Roberts, J. J., and Pascoe, J. M. Crosslinking of complementary strands of DNA in Corps Journal (22: 17—19,1987). mammalian cells by antitumor platinum compounds. Nature (Land.), 235: 282—284, I thankGailFeigenbaum,Curatorof AcademicPrograms,NationalGallery 1971. 5543

Downloaded from cancerres.aacrjournals.org on September 27, 2021. © 1996 American Association for Cancer Research. HISTORYAND PROSPECTSOF DNA-TARGETEDCANCERTREATMENT

36. Kohn, K. W. Interstrand cross-linking of DNA by l,3-bis(2-chloroethyl)-l-nitro cross-linking by haloethylnitrosoureas in L12l0 cells. Cancer Res., 38: 3197—3203, sourea and other l-(2-haloethyl)-l-. Cancer Res., 37: 1450—1454,1977. 1978. 37. Kohn, K. W., and Ewig, R. A. Alkaline elution analysis, a new approach to the study 65. Erickson, L C., Laurent, 0., Sharkey, N. A., and Kohn, K. W. DNA crosslinking of DNA single-strand interruptions in cells. Cancer Res., 33: 1849—1853,1973. and monoadduct repair in -treated human tumor cells. Nature (Lond.), 38. Kohn, K. W. DNA as a target in cancer chemotherapy: measurement of macromo 288: 727—729,1980. lecular DNA damage produced in mammalian cells by anticancer agents and 66. Erickson, L. C., Bradley, M. 0., and Kohn, K. W. Measurements of DNA damage carcinogens. Methods Cancer Res., 16: 291—345,1979. in Chinese hamster cells treated with equitoxic and equimutagenic doses of nitro 39. Kohn, K. W., Ewig, R. A., Erickson, L. C., and Zwelling, L. A. Measurement of soul-eat.Cancer Res., 38: 3379—3384,1978. strand breaks and crosslinks in DNA by alkaline elution. In: E. C. Friedberg and P. 67. Erickson, L. C., Bradley, M. 0., Ducore, J. M., Ewig, R. A., and Kohn, K. W. DNA C.Hanawalt(eds.),DNARepair:ALaboratoryManualofResearchTechniques,pp. crosslinkingandcytotoxicityinnormalandtransformedhumancellstreatedwith 379—401.New York: Marcel Dekker, 1981. antitumor nitrosoureas. Proc. Nail. Acad. Sci. USA., 77: 467—471,1980. 40. Kuhn, K. W. The significance of DNA damage assays in toxicity and carcinoge 68. Kohn,K. W.Prospectsforimprovedchloroethylnitrosoureasandrelatedhaloethy nicity assessment. Ann. NY Acad. Sci., 407: 106—118,1983. lating agents. In: P. L. Kornblith and M. D. Walker (edt.), Advances in Neuro 41. Kohn, K. W. DNA filter elution methods in anticancer drug development. Cancer Oncology, pp. 491—513.Mount Kisco, NY: Futura Publishing Co., Inc., 1988. Treat Res., 36: 3—38,1987. 69. Gibson, N. W., Erickson, L. C., and Kohn, K. W. DNA damage and differential 42. Kohn, K. W. Principles and practice of DNA filter elution. Pharmacol. Ther., 49: cytotoxicity produced in human cells by 2-chioroethyl(methylsulfonyl)- 55—77,1991. methanesulfonate, a new DNA chloroethylating agent. Cancer Res., 45: 1674—1679, 43. Bradley, M. 0., and Kohn, K. ‘A'.X-rayinduced DNA double strand break produc 1985. tion and repair in mammalian cells as measured by neutral filter elution. Nucleic 70. Sariban, E., Erickson, L. C., and Kohn, K. W. Effects of carbamoylation on cell Acids Res., 7: 793—804,1979. survival and DNA repair in normal human embryo cells treated with various 44. Fornace, A. J., Jr. Measurement of M. luteus endonuclease sensitive lesions by l-(2-chloroethyl)-l-nitrosoureas. Cancer Rca., 44: 1352—1357,1984. alkaline elution. Mutat. Res., 94: 263—276,1982. 71. Wang, J. C. Interaction between DNA and an protein omega. J. 45. Kohn, K. W., and Ewig, R. A. DNA-protein crosslinking by trans-platinum(ll)di Mol. Biol., 55: 523—533,1971. amminedichloride in mammaliancells, a new method ofanalysis. Biochim. Biophys. 72. Blair, D. G., and Helinski, D. R. Relaxation complexes of plasmid DNA and protein. Acts, 562: 32—40,1979. I. Strand-specific association of protein and DNA in the relaxed complexes of 46. Ross, W. E., Glaubiger, D. L., and Kohn, K. W. Protein-associated DNA breaks in plasmids ColEl and Co1E2.J. Biol. Chem., 250: 8785—8789,1975. cells treated with adriamycin or ellipticine. Biochim. Biophys. Acts, 519: 23—30, 73. Champoux, J. J., and Dulbecco, R. An activity from mammalian cells that untwists 1978. superhelical DNA: a possible swivel for DNA replication. Proc. Nail. Acad. Sci. 47. Ross, W. E., Glaubiger, D., and Kohn, K. W. Qualitative and quantitative aspects USA, 69: 143—146,1972. of intercalator-induced DNA strand breaks. Biochim. Biophys. Acta, 562: 41— 74. Champoux, J. J. Strand breakage by the DNA untwisting enzyme results in 50, 1979. covalent attachment of enzyme to DNA. Proc. Nail. Acad. Sci. USA, 74: 48. Kohn, K. W., Erickson, L. C., Ewig, R. A., and Friedman, C. A. Fractionation of 3800—3804,1977. DNA from mammalian cells by alkaline elution. Biochemistry, 15: 4629—4637, 75. Zwelling, L. A., Michaels, S., Erickson, L. C., Ungerleiter, R. S., Nichols, M., and 1976. Kohn,K.W.Protein-associatedDNAstrandbreaksin Ll2lO cellstreatedwiththe 49. Fornace, A. J. J., Kohn, K. W., and Kann, H. E. J. DNA single-strand breaks during DNA intercalating agents amsacrine and adriamycin. Biochemistry, 20:6553-6563, repair of UV damage in human fibroblasts and abnormalities of repair in xeroderma 1981. pigmentosum. Proc. Nail. Acad. Sci. USA, 73: 39—43,1976. 76. Zwelling. L A., Michaels, S., Kerrigan, D., Pommier, Y., and Kohn, K. W. 50. Fomace, A. J., Jr., and Kohn, K. W. DNA-protein cross-linking by ultraviolet Protein-associated DNA strand breaks produced in mouse leukemia Ll210 cells by radiation in normal human and xeroderma pigmentosum fibroblasts. Biochim. Bin ellipticine and 2-methyl-9-hydroxyellipticinium. Biochem. Pharmacol., 31: 3261— phys. Acts, 435: 95—103.1976. 3267, 1982. 51. Ewig, R. A., and Kohn, K. W. DNA damage and repair in mouse leukemia L1210 77. Zwelling, L. A., Kemgan, D., and Michaels, S. Cytotoxicity and DNA strand breaks cells treated with nitrogen mustard, 1,3-bit (2-chloroethyl)-l-nitrosourea. and other by 5-iminodaunorubicin in mouse leukemia cells. Cancer Res., 42: 2687—2691, nitrosoureas. Cancer Res., 37: 2114—2122,1977. 1982. 52. Ross, W. E., Ewig, R. A., and Kohn, K. W. Differences between and 78. Pommier, Y., Covey, J., Kerrigan, D., Mattes, W., Markovits, J., and Kohn, K. W. nitrogen mustard in the formation and removal of DNA cross-links. Cancer Res., 38: Role of DNA intercalation in the inhibition of purified mouse DNA topoisomerase 1502—1506,1978. II by 9-aminoacridines. Biochem. Pharmacol., 36: 3477—3486, 1987. 53. Zwelling, L. A., Filipski, J., and Kohn, K. W. Effect of thiourea on survival and 79. Filipski, J., and Kohn, K. W. Ellipticine-induced protein-associated DNA breaks in DNA crosslink formation in cells treated with platinum(ll) complexes. L-phenylala isolated 11210 nuclei. Biochim. Biophys. Acts. 698: 280—286,1982. nine mustard, and Bis(2-chloroethyl)methylamine. Cancer Res., 39: 4989—4995, 80. Pommier, Y., Schwartz, R. E., Kobn, K. W., and Zwelling, L. A. Formation and 1979. rejoining of DNA double-strand breaks induced in isolated cell nuclei by antineo 54. Erickson, L. C., Ramonas, L. M., Zaharko, D. S., and Kohn, K. W. Cytotoxicity and plastic intercalating agents. Biochemistry, 23: 3194—3201, 1984. DNA crosslinking activity of 4-sulfidocyclophosphamides in mouse leukemia cells. 81. Glisson, B. S., Smallwood, S. E., and Ross, W. E. Characterization of VP-16- CancerRes.,40:4216—4220,1980. induced DNA damage in isolated nuclei from Ll2lO cells. Biochim. Biophys. Acts, 55. Ducore, J. M., Erickson, L. C., Zwelling, L. A., Laurent, 0., and Kohn, K. W. 783: 74—79,1984. Comparative studies of DNA crosslinking and cytotoxicity in Burkiti's lymphoma 82. Covey, J. M., Kohn, K. W., Kerrigan, D., Tilchen, E. J., and Pommier, Y. Topoi cell lines treated with cid-diamminedichloroplatinum(ll) and i-phenylalanine mus somerase H-mediated DNA damage produced by 4'-(9-acridinylamino)-methanesul taid. Cancer Res., 42: 897—902,1982. fon-m-anisidide and related acridines in Ll2lO cells and isolated nuclei: relation to 56. Sariban, E., Kohn, K. W., Zlotogorski, C., Laurent, 0., D'Incalci, M., Day, R., cytotoxicity. Cancer Res., 48: 860—865, 1988. Smith, B. H., Kornblith, P. L., and Erickson, L. C. DNA cross-linking responses of 83. Filipski, J., Yin, J., and Kohn, K. W. Reconstitution of intercalator-induced DNA human malignant glioma cell strains to chloroethylnitrosoureas, cisplatin, and dia scission by an active component from nuclear extracts. Biochim. Biophys. Acts, ziquone. Cancer Res., 47: 3988—3994, 1987. 741:116—122,1983. 57. Zwelling, L. A., Kohn, K. W., Ross, W. E., Ewig, R. A., and Anderson, T. Kinetics 84. Liu, L. F., Liu, C. C., and Alberta, B. M. Type II DNA topoisomerases. Enzymes that of formation and disappearance of a DNA cross-linking effect in mouse leukemia can unknot a topologically knotted DNA molecule via a reversible double-strand L1210 cells treated with cis- and trans-diamminedichloroplatinum(II). Cancer Ret., break. Cell, 19: 697—707,1980. 38: 1762—1768,1978. 85. Liu, L. F., Rowe, T. C., Yang, L., Tewey, K. M., and Chen, G. L. Cleavage of DNA 58. Zwelling, L. A., Anderson, T., and Kohn, K. W. DNA-protein and DNA interstrand by mammalian DNA topoisomerase II. J. Biol. Chem., 258: 15365—15370,1983. crosslinking by cis- and trans-platinum(H) diamminedichloride in Ll210 mouse 86. Tewey, K. M., Rowe, T. C., Yang, L., Halligan, B. D., and Liu, L. F. Adriamycin leukemia cells and relation to cytotoxicity. Cancer Res., 39: 365—369,1979. induced DNA damage is mediated mammalian DNA topoisomerase II. Science 59. Zwelling, L. A., Bradley, M. 0., Sharkey, N. A., Anderson, T., and Kohn, K. W. (Washington DC), 226: 466—468,1984. Mutagenicity, cytotoxicity and DNA crosslinking in V79 Chinese hamster cells 87. Tewey, K. M., Chen, G. L., Nelson, E. M., and Liu, L. F. Intercalative antitumor treated with cis- and trans-Pt(H)diammindichloride. Mutat. Res., 67: 271—280,1979. drugs interfere with the breakage-reunion reaction of mammalian DNA topoisomer 60. Erickson, L. C., Zwelling, L. A., Ducore, J. M., Sharkey, N. A., and Kohn, K. W. ate II. J. Biol. Chem., 259: 9182—9187, 1984. Differential cytotoxicity and DNA crosslinking in normal and transformed human 88. Nelson, E. M., Tewey, K. M., and Liu, L. F. Mechanism of antitumor drug action: fibroblasts treated with cis-diamminedichioroplatinum(II). Cancer Res., 41: 2791— poisoning of mammalian DNA topoisomerase II on DNA by amsacrine. Proc. Nail. 2794,1981. Aced. Sci. USA, 81: 1361—1365,1984. 61. Laurent, 0., Erickson, L. C., Sharkey, N. A., and Kohn, K. W. DNA crosslinking 89. Minford, J., Pommier, Y., Filipski, J., Kohn, K. W., Kerrigan, D., Mattern, M., and cytotoxicity induced by cis-diammineplatinum(Il) in human normal and tumor Michaels, S., Schwartz, R., and Zwelling, L. A. Isolation of intercalator-dependent cell lines. Cancer Res., 41: 3347—3351,1981. protein-linked DNA strand cleavage activity from cell nuclei and identification as 62. Micetich, K., Zwelling, L. A., and Kohn, K. W. Quenching of DNA:platinum(II) topoisomerase II. Biochemistry, 25: 9—16,1986. monoadducts as a possible mechanism of resistance to cis-diamminedichloroplati 90. Ross, W. E., Rowe, T. C., Glisson, B. S., Yalowich, J., and Liu, L. F. Role of num(II) in L12l0 cells. Cancer Res., 43: 3609—3613, 1983. topoisomerase II in mediating epipodophyllotoxin-induced DNA cleavage. Cancer 63. Cohen, L. F., Ewig, R. A., Kohn, K. W., and Glaubiger, D. Interstrand DNA Res., 44: 5857—5860,1984. crosslinking by 4,5',8-trimethylpsoralen plus monochromatic ultraviolet light. Stud 91. Rowe, T. C., Kupfer, 0., and Ross, W. E. Inhibition of epipodophyllotoxin cyto ies by alkaline elution in mouse Ll2lO leukemia cells. Biochim. Biophys. Acta, 610: toxicity by interference with topoisomerase-mediated DNA cleavage. Biochem. 56—63,1980. Pharmacol., 34: 2483—2487, 1985. 64. Ewig, R. A., and Kohn, K. W. DNA-protein cross-linking and DNA interstrand 92. Long, B. H., Musial, S. T., and Brattain, M. G. Single- and double-strand DNA 5544

Downloaded from cancerres.aacrjournals.org on September 27, 2021. © 1996 American Association for Cancer Research. HISTORYAND PROSPECTSOF DNA-TARGETEDCANCERTREATMENT

breakage and repair in human lung adenocarcinoma cells exposed to etoposide and cleavages in the transcriptionally active tyrosine aminotransferase gene. Cell, 50: teniposide. Cancer Res., 45: 3106—3112, 1985. 1109—1117,1987. 93. Long, B. H., Musial, S. T., and Brattain, M. 0. DNA breakage in human lung 120. Culotta, V., and Sollner-Webb, B. Sites of topoisomerase I action on X. laevis carcinoma cells and nuclei that are naturally sensitive or resistant to etoposide and ribosomal chromatin: transcriptionally active rDNA has an —200bp repeating teniposide. Cancer Ret., 46: 3809—3816, 1986. structure. Cell, 52: 585—597,1988. 94. Kerrigan, D., Pommier, Y., and Kohn, K. W. Protein-linked DNA strand breaks 121. Pommier, Y., Capranico, 0., Orr, A., and Kohn, K. W. Local base sequence produced by etoposide and teniposide in Mouse L12l0 and Human VA-l3 and preferences for DNA cleavage by mammalian topoisomerase II in the presence of HT-29 cell lines: relationship to cytotoxicity. NC! Monogr., 4: 117—121, amsacrine or temposide. Nucleic Acids Res., 19: 5973—5980,1991. 1987. 122. Capranico, 0., Zunino, F., Kohn, K. W., and Pommier, Y. Sequence-selective 95. Long, B. H., Musial, S. T., and Brattain, M. 0. Comparison ofcytotoxicity and DNA topoisomerase H inhibition by anthracydline derivatives in SV4O DNA: relationship breakage activity of congeners of podophyllotoxin including VP16-.213 and VM with DNA binding affinity and cytotoxicity. Biochemistry, 29: 562-569, 1990. 26: a quantitative structure-activity relationship. Biochemistry, 23: 1183—1188, 123. Jaxel, C., Capranico, 0., Kerrigan, D., Kohn, K. W., and Pommier, Y. Effect of local 1984. DNA sequence on topoisomerase I cleavage in the presence or absence of campto 96. Liu, S-Y., Hwang, B-D., Haruna, M., Imakura, Y., Lee, K-H., and Cheng, Y-C. thecin. J. Biol. Chem., 266: 20418—20423, 1991. Podophyllotoxin analogs: effects on DNA topoisomerase II, tubulin polymerization. 124. Tanizawa, A., Kohn, K. W., and Pommier, Y. Induction of cleavage in topoisomer human KB cells, and their VP-l6-resistant variants. Mol. Pharmacol., 36: 78—82, ase I cDNA by topoisomerase I enzymes from calf thymus and wheat germ in the 1989. presence and absence of camptothecin. Nucleic Acids Ret., 21: 5157—5166,1993. 97. Chang, J-Y., Han, F-S., Liu, S-Y., Wang, Z-Q., Lee, K-H., and Cheng, Y-C. Effect 125. Jaxel, C., Kohn, K. W., Wani, M. C., Wall, M. E., and Pommier, Y. Structure of 4b-arylamino derivatives of 4'-O-demethylepipodophyllotoxin on human DNA activity study of the actions of camptothecin derivatives on mammalian topoisomer topoisomerase II, tubulin polymerization, KB cells, and their resistant variants. ase I: evidence for a specific receptor site and a relation to antitumor activity. Cancer CancerRes., 51: 1755—1759,1991. Ret.,49: 1465—1469,1989. 98. Pommier, Y., Tanizawa, A., and Kohn, K. W. Mechanisms of topoisomerase I 126. Leteurtre, F., Fesen, M., Kohthagen, G., Kohn, K. W., and Pommier, Y. Specific inhibition by anticancer drugs. Adv. Pharmacol., 29B: 73—91,1994. interaction of camptothecin, a topoisomerase I inhibitor, with guanine residues of 99. Gupta, M., Fujimori, A., and Pommier, Y. Eukazyotic DNA topoisomerases I. DNA detected by photoactivation at 365 nm. Biochemistry, 32: 8955—8962, 1993. Biochini.Biophys.Acts,1262:1—14,1995. 127. Pommier, Y., Kohthagen, 0., Kohn, K. W., Leteuitre, F., Wan!, M. C., and Wall, M. E. 100.Wall, M. E., and Warn,M. C. Camptothecinandtaxol: discoveryto clinic Interaction of an alkylating camptothecin derivative with a DNA base at topoisomerase thirteenth Bruce F. Cain Memorial Award Lecture. Cancer Ret., 55: 753-760, 1995. I-DNA cleavage sites. Proc. Ned. Acad. Sd. USA, 92: 8861-8865, 1995. 101. Horwitz, M. S., and Horwitz, S. B. Intracellular degradation of HeLa and adenovirus 128. Mattes, W. B., Hartley, J. A., and Kohn, K. W. DNA sequence selectivity of type 2 DNA induced by camptothecin. Biochem. Biophys. Ret. Commun., 45: guanine-N7 alkylation by nitrogen mustards. Nucleic Acids Ret., 14: 2971—2987, 723—727,1971. 1986. 102. Hsiang, Y. H., Hertzberg, R., Hecht, S., and Liu, L. F. Camptothecin induces 129. Kohn, K. W., Hartley, J. A., and Mattes, W. B. Mechanisms of DNA sequence protein-linked DNA breaks via mammalian DNA topoisomerase I. J. Biol. Chem., selective alkylation of guanine-/'P positions by nitrogen mustards. Nucleic Acids 260: 14873—14878,1985. Ret.,15:10531—10549,1987. 103. Jaxel, C.. Kohn, K. W., and Pommier, Y. Topoisomerase I interaction with SV4O 130. Mattes, W. B., Hartley, J. A., and Kohn, K. W. GC-rich regions in genomes as DNA in the presence and absence of camptothecin. Nucleic Acids Ret., 16: 11157— targets for DNA alkylation. Carcinogenesis (Lond.), 9: 2065—2072, 1988. 11170,1988. 131. Kohn, K. W., Orr, A., O'Connor, P. M., Guziec, L. J., and Guziek, F. S., Jr. 104. Covey, J. M., Jaxel, C., Kohn, K. W., and Ponunier, Y. Protein-linked DNA strand Synthesis and DNA-sequence selectivity of a series of mono and difunctional breaks induced in mammalian cells by camptothecin, an inhibitor of topoisomerase 9-aminoacridine nitrogen mustards. J. Med. Chem., 37: 67-72, 1994. I. Cancer Ret., 49: 5016—5022, 1989. 132. Creech, H. J., Preston, R. K., Peck, R. M., O'Connell, A. P., and Ames, B. N. 105. Andoh, T., Ishii, K., Suzuki, Y., Ikegami, Y., Kusunoki, Y., Takemoto, Y., and Antitumor and mutagenic properties of a variety of heterocyclic nitrogen and sulfur Okada, K. Characterization of a mammalian mutant with a camptothecin-resistant mustards. J. Med. Chem., 15: 739—746,1972. DNA topoisomerae I. Proc. Nail. Aced. Sci. USA, 84: 5565—5569,1987. 133. Markovits, J., Pommier, Y., Kemgan, D., Covey, J., Tilchen, E. J., and Kohn, K. W. 106. Kjeldsen, E., Bonven, B. J., Andoh, T., Ishii, K., Okada, K., Bolund, L., and Topoisomerase 11-mediated DNA breaks and cytotoxicity in relation to cell prolif Westergaard, 0. Characterization of a camptothecin-resistant human DNA topoi eration and the cell cycle in NIH3T3 fibroblasts and Ll210 leukemia cells. Cancer somerase I. J. Biol. Chem., 263: 3912—3916,1988. Ret., 47: 2050—2055, 1987. 107. Gupta, R. S., Gupta, R., Eng, B., Lock, R. B., Ross, W. E., Hertzberg, R. P., Caranfa, 134. Kaufmann, S. H. Antagonism between camptothecin and topoisomerase H-directed M. J., and Johnson, R. K. Camptothecin-resistant mutants of Chinese hamster ovary chemotherapeutic agents in a human leukemia cell line. Cancer Ret., 51: 1129— cells containing a resistant form of topoisomerase I. Cancer Ret., 48: 6404—6410, 1136, 1991. 1988. 135. O'Connor, P. M., Nieves-Neira, W., Kerrigan, D., Bertrand, R., Goldman, J., Kohn, 108. Sullivan, D. M., Latham, M. D., Rowe, T. C., and Ross, W. E. Purification and K. W., and Pommier, Y. S-phase population does not correlate with the cytotoxicity characterization of an altered topoisomerase II from a drug-resistant Chinese ham of camptothecin and 10,1 1-methylenedioxycamptothecin in human colon carcinoma ster ovary cell line. Biochemistry, 28: 5680—5687, 1989. HT-29 cells. Cancer Commun., 3: 233—240,1991. 109. Tamura, H., Kohchi, C., Yamada, R., Ikeda, T., Koiwai, 0., Patterson, E., Keene, 136. Hsiang, Y., Liu, L. F., Wall, M. E., Wani, M. C., Nicholas, A. W., Manikumar, G., J. D., Okada, K., Kjeldsen, E., Nishikawa, K., and Andoh, T. Molecular cloning of Kirschenbaum, S., Silber, R., and Potmesil, M. DNA topoisomerase I-mediated a cDNA of a camptothecin-resistant human DNA topoisomerase I and identification DNA cleavage and cytotoxicity of camptothecin analogues. Cancer Ret.. 49: 4385— of mutation sites. Nucleic Acids Ret., 19: 69—75,1991. 4389, 1989. 110. Tanizawa, A., and Pommier, Y. Topoisomerase I alteration in a camptothecin 137. O'Connor, P. M., Kerrigan, D., Bertrand, R., Kohn, K. W., and Pommier, Y. resistant cell line derived from Chinese hamster DC3F cells in culture. Cancer Ret., 10,11-Methylenedioxycamptothecin. a topoisomerase I inhibitor of increased po 52: 1848—1854,1992. tency: DNA damage and correlation to cytotoxicity in human colon carcinoma 111. Tanizawa, A., Bertrand, R., Kohlhagen, 0., Tabuchi, A., Jenkins, J., and Pommier, (HT-29) cells. Cancer Commun., 2: 395—400,1990. Y. Cloning of Chinese hamster DNA topoisomerase I cDNA and identification of a 138. Potmesil, M. Camptothecins: from bench research to hospital wards. Cancer Ret., single point mutation responsible for camptothecin resistance. J. Biol. Chem., 268: 54: 1431—1439,1994. 25463—25468,1993. 139. Hertzberg, R. P., Caranfa, M. J., and Hecht, S. M. On the mechanism of topoi 112. Fujimori, A., Harker, 0., Kohthagen, 0., Hold, Y., and Pommier, Y. Mutation at the somerase I inhibition by camptothecin: evidence for binding to an enzyme-DNA catalytic site of topoisomerase I in CEM/C2, a human leukemia cell line resistant to complex. Biochemistry, 28: 4629—4638, 1989. camptothecin. Cancer Res., 55: 1339—1346,1995. 140. Crow, R. T., and Crothers, D. M. Structural Modifications of camptothecin and 113. Pommier, Y., Capranico, 0., Orr, A., and Kohn, K. W. Distribution of topoisomerase effects on topoisomerase I inhibition. J. Med. Chem., 35: 4160—4164, 1992. II cleavage sites in SV4O DNA and effects of drugs. J. Mol. Biol., 222: 909—924, 141. Hsiang, Y., Lihou, M. G., and Liu, L. F. Arrest of replication forks by drug 1991. stabilized topoisomerase I-DNA cleavable complexes as a mechanism of cell killing 114. Wasserman, K., Markovits, J., Jaxel, C., Capranico, 0., Kohn, K. W., and Pommier, by camptothecin. Cancer Ret., 49: 5077—5082,1989. Y. Effects of morphollnyldoxorubicin, doxorubicin, and actinomycin D on mam 142. Holm, C., Covey, J. M., Kerrigan, D., and Pommier, Y. Differential requirement of malian DNA topoisomerases I and H. Mol. Pharmacol., 38: 38-45, 1990. DNA replication for the cytotoxicity of DNA topoisomerase I and II inhibitors in 115. Capranico, G., Jaxel, C., Roberge, M., Kolm, K. W., and Pommier, Y. Nucleosome Chinese hamster DC3F cells. Cancer Ret., 49: 6365—6368, 1989. positioning as a critical determinant for the DNA cleavage sites of mammalian DNA 143. Zhang, H., D'Arpa, P., and Liu, L F. A model for tumor cell killing by topoisomer topoisomerase II in reconstituted SV4Ochromatin. Nucleic Acids Res., 18: 4553— ase poisons. Cancer Cells, 2: 23—27,1990. 4559, 1990. 144. Avemann, K., Knippers, R., Koller, T., and Sogo, J. M. Camptothecin, a specific 116. Pommier, Y., Cockerill, P. N., Kohn, K. W., and Garrard,W. T. Identification within inhibitor of type I DNA topoisomerase, induces DNA breakage at replication forks. the SV4Ogenome of a chromosomal loop attachment site that contains topoisomer Mol. Cell. Biol., 8: 3026—3034,1988. ase H cleavage sites. J. Virol., 64: 419—423, 1990. 145. Pommier, Y., Leteurtre, F., Fesen, M. R., Fujimori, A., Bertrand, R., Solary, E., 117. Riou, J-F., Lefevre, D., and Riou, G. Stimulation of the topoisomerase H induced Kohthagen, 0., and Kohn, K. W. Cellular determinants of sensitivity and resistance DNA cleavage sites in the c-myc gene by antitumor drugs is associated with gene to DNA topoisomerase inhibitors. Cancer Invest., 12: 530—542,1994. expression. Biochemistry, 28: 9104—9110, 1989. 146. Tanizawa, A., Kohn, K. W., Kohihagen, 0., Leteurtre, F., and Pommier, Y. Differ 118. Pommier, Y., Orr, A., Kohn, K. W., and Riou, J. F. Differential effects of amsacrine ential stabilization of eukaryotic DNA topoisomerase I cleavable complexes by and epipodophyllotoxins on topoisomerase H cleavage in the human c-myc proto camptothecin derivatives. Biochemistry, 34: 7200—7206, 1995. oncogene. Cancer Res., 52: 3125—3130,1992. 147. Hartwell, L. H., and Weinert, T. A. Checkpoints: controls that ensure the order of 119. Stewart, A. F., and Schutz, 0. Camptothecin-induced in vivo topoisomerase I cell cycle events. Science (Washington DC), 246: 629—643, 1989. 5545

Downloaded from cancerres.aacrjournals.org on September 27, 2021. © 1996 American Association for Cancer Research. HISTORY AND PROSPECTS OF DNA-TARGETEDCANCER TREATMENT

148. Hartwell, L. Defects in a cell cycle checkpoint may be responsible for the genomic 181. Smith, M. L., Chen, I. T., Zhan, Q., O'Connor, P. M., and Fornace, A. J., Jr. instability of cancer cells. Cell, 71: 543—546,1992. Involvement of the p53 tumor suppressor in repair of u.v.-type DNA damage. 149. Murray, A. W. Creative blocks: cell-cycle checkpoints and feedback controls. Oncogene, 10: 1053—1059,1995. Nature (Lond.), 359: 599—604, 1992. 182. Smith, M. L., Udo Kontny, H., Than, Q., Sreenath, A., O'Connor, P. M., and 150. O'Connor, P. M., Ferns, D. K., Pagano, M., Draetta, G., Pines, J., Hunter, 1., Longo, Fornace, A. J., Jr. Antisense GADD45 expression results in decreased DNA repair D. L., and Kohn, K. W. G2delay induced by nitrogen mustard in human cells affects and sensitizes cells to u.v.-irradiation or cisplatin. Oncogene, 13: in press, 1996. cyclin A/cdk2 and cyclin Bl/cdc2-kinase complexes differently. J. Biol. Chem., 183. Smith, M. L., and Fornace, A. J., Jr. Genomic instability and the role of p53 268:8298—8308,1993. mutations in cancer cells. Curr. Opin. Oncol., 7: 69—75,1995. 151. O'Connor, P. M., and Kohn, K. W. A fundamental role for cell cycle regulation in 184. Zambetti, 0. P., and Levine, A. J. A comparison of the biological activities of the chemosensitivity of cancer cells? Semin. Cancer Biol., 3: 409—416, 1992. wild-type and mutant p53. FASEB J., 7: 855—865, 1993. 152. Kohn, K. W., Jackman, J., and O'Connor, P. M. Cell cycle control and cancer 185. Ko, L. J., and Privet, C. pS3: puzzle and paradigm. Genes Dcv., 10: 1054—1072, chemotherapy. J. Cell. Biochem.. 54: 440—452, 1994. 1996. 153. Lou, C. C., and Pardee,A. B. Mechanismby which caffeine potentiates lethality of 186. Kastan, M. B., Than, Q., El-Deiry, W. S., Carrier, F., Jacks, T., Walsh, W. V., nitrogen mustard. Proc. Natl. Acad. Sci. USA, 79: 2942—2946, 1982. Plunkett, B. S., Vogelstein, B., and Fornace, A. J. J. A mammalian cell cycle 154. Fingert, H. J.. Chang, J. D., and Pardee, A. B. Cytotoxic, cell cycle, and chromo checkpoint pathway utilizing p53 and GADD45 is defective in ataxia-telangiectasia. somal effects of methylxanthines in human tumor cells treated with alkylating Cell, 71: 587—597,1992. agents. Cancer Res., 46: 2463—2467, 1986. 187. El-Dairy, W. S., Tokino, T., Velculescu, V. E., Levy, D. B., Parsons, R., Trent, J. M., 155. Fingert, H. J., Pu, A. T., Chen, Z. Y., Gouge, P. B., Alley, M. C., and Pardee, A. B. Lin, D., Mercer, W. E., Kinzler, K. W., and Vogelstein, B. WAF1, a potential in vivo and in vitro enhanced antitumor effects by pentoxifylline in human cancer mediator of p53 tumor suppression. Cell, 75: 817—825,1993. cells treated with . Cancer Ret.. 48: 4375—4381, 1988. 188. El-Deiry, W. S., Harper, J. W., O'Connor, P. M., Velculescu, V. E., Canman, C. E., 156. Teicher, B. A., Holden, S. A., Herman, T. S., Epelbaum, U., Pardee, A. B., and Jackman, J., Pietenpol, J. A., Burrell, M., Hill, D. E., Wang, Y., Wilman, K. 0., Dezube, B. Efficacy of pentoxifylline as a modulator of alkylating agent activity in Mercer, W. E., Kastan, M. B., Kohn, K. W., Elledge, S. J., Kinzler, K. W., and vitro and in vivo. Anticancer Ret., II: 1555—1560,1991. Vogeistein, B. WAF1/CIP1 is induced in p53-mediated G@arrest and apoptosis. 157. Fan, S., Smith, M. L.. Rivet, D. J., Duba, D., Zhan, Q., Kohn, K. W., Fornace, A. J., Cancer Ret., 54: 1169—1174,1994. Jr., and O'Connor, P. M. Disruption of p53 function sensitizes MCF-7 189. Miyashita, T., Krajewski, S., Krajewska, M., Wang, H. 0., Lin, H. K., Liebermann, cells to cisplatin and pentoxifylline. Cancer Ret., 55: 1649—1654, 1995. D. A., Hoffman, B., and Reed, J. C. Tumor suppressor p53 is a regulator of bcl-2 and 158. Wang, Q., Fan, S., Eastman, A., Worland, P. J., Sausville, E. A., and O'Connor, bax gene expression in vitro and vivo. Oncogene, 9: 1799—1805,1994. P. M. UCN-Ol: a potent abrogator of G2 checkpoint function in cancer cells with 190. Miyashita, T., and Reed, J. C. Tumor suppressor p53 is a direct transcriptional ditrupted p53. J. NatI. Cancer Inst., 88: 956—963,1996. activator of the human box gene. Cell, 80: 293—299,1995. 159. Lock, R. B. Inhibition of p34cdc2 kinase activation, p34cdc2 tyrosine dephospho 191. Than, Q., Fan, S., Bae, I., Guillouf, C., Liebermann, D. A., O'Connor, P. M., and rylation, and mitotic progression in Chinese hamster ovary cells exposed to etopo Fornace, A. J., Jr. Induction of bax by genotoxic stress in human cells correlates with side. Cancer Ret., 52: 1817—1822,1992. normal status and apoptosis [published erratum appears in Oncogene, 10: 1259, 160. Tsao, Y. P., DArpa, P., and Liu, L. F. The involvement of active DNA synthesis in 1995]. Oncogene, 9: 3743—3751,1994. camptothecin-induced G2 arrest: altered regulation of p34cdc2/cyclin B. Cancer 192. Wu, X., and Levine, A. J. p53 and E2F-l cooperate to mediate apoptosis. Proc. Nail. Ret., 52: 1823—1829,1992. Acad. Sci. USA, 91: 3602—3606,1994. 161. Mutchel, R. J., Zhang, H. B., Iliakis, 0., and McKenna, W. B. Cyclin B expression 193. Kessis, T., Slebos, R. J., Nelson, W. 0., Kastan, M. B., Plunkett, B. S., Han, S. M., in HeLa cells during the 02 block induced by ionizing radiation. Cancer Ret., 51: Lorincz, A. T., Hedrick, L., and Cho, K. R. Human papillomavirus 16E6 expression 5113—5117,1991. disrupt the p53 mediated cellular response to DNA damage. Proc. Nail. Aced. Sci. 162. Muschel, R. J., Zhang, H. B., and McKenna, W. 0. Differential effect of ionizing USA, 90: 3988—3992,1993. radiation on the expression of cyclin A and cyclin B in HeLa cells. Cancer Ret., 53: 194. Monks, A., Scudiero, D., Skehan, P., Shoemaker, R., Paull, K., Vistica, D., Hose, C., 1128—1135,1993. Langley, J., Cronise, P., Vigro-Wolff, A., et aL Feasibility of a high-flux anticancer 163. O'Connor, P. M., Ferns, D. K., Hoffmann, I., Jackman, J., Draeua, G., and Kohn, drug screen using a diverse panel of cultured human tumor cell lines. J. Nail. Cancer K. W. Role of the cdc25C phosphatase in 02 arrest induced by nitrogen mustard. Inst., 83: 757—766,1991. Proc. NatI. Acad. Sci. USA, 91: 9480—9484,1994. 195. Paull, K. D., Shoemaker, R. H., Hodes, L., Monks, A., Scudiero, D. A., Rubinstein, 164. O'Connor, P. M., and Fan, S. DNA damage checkpoints: implications for cancer L., Plowman, J., and Boyd, M. R. Display and analysis of patterns of differential therapy. Prog. Cell Cycle Res., 2: 1—9,1996. activity of drugs against human tumor cell lines: Development of mean graph and 165. Kastan, M. B. Participation of p53 protein in the cellular response to DNA damage. COMPARE algorithm. J. Nail. Cancer Inst., 81: 1088—1092,1989. Cancer Ret., 51: 6304—6311,1991. 196. Weinstein, J. N., Kohn, K. W., Graver, M. R., Viswanadhan, V. N., Rubinstein, 166. Kuerbitz, S. J., Plunkett, B. S., Walsh, W. V., and Kastan, M. B. Wild-type p53 is L. V., Monks, A. P., Scudiero, D. A., Welch, L., Koutsoukos, A. D., Chiausa, A. J., a cell cycle checkpoint determinant following irradiation. Proc. Natl. Acad. Sci. and Paull, K. D. Neural computing in cancer drug development: predicting mech USA, 89:7491—7495,1992. anism of action. Science (Washington DC), 258: 447—451, 1992. 167. Sherr, C. J. Mammalian G, cyclins. Cell, 73: 1059—1065, 1993. 197. van Osdol, W. W., Myers, T. 0., Paull, K. D., Kohn, K. W., and Weinstein, J. N. Use 168. Sherr, C. J. G, phase progression: cycling on cue. Cell, 79: 551—555,1994. of the Kohonen self-organizing map to study the mechanisms of action chemother 169. Morgan, D. 0. Principles of CDK regulation. Nature (Lond.), 374: 131—134,1995. apeutic agents. J. Nail. Cancer Inst., 86: 1853—1859,1994. 170. Levine, A., Momand, J., and Finlay, C. A. The p53 tumor suppressor gene. Nature 198. Weinstein,J. N., Myers,T., Buolamwini,J., Raghavan,K., Viswanadhan,V. N., Licht, (Lond.), 351: 453—456,1991. J., Rubinstein,L V., Koutsoukos,A. K., Kohn, K. W., Zaharevitz,D., Graver, M. R., 171. Hollstein, M., Sidransky, D., Vogelstein, B., and Harris, C. C. p53 mutations in Monks, M. R., Scudiero, K. A., Chabner, B. A., Anderson, N. L, and Paul!, K. D. human cancers. Science (Washington DC), 253: 49—53,1991. Predictive statistics and artificial intelligencein the U. S. National Cancer Institute's 172. Stone, S., Dayananth, P., and Kamb, A. Reversible, p16-mediated cell cycle arrest drug discovery program for cancer and AIDS. Stem Cells, 12: 13—22,1994. as protection from chemotherapy. Cancer Ret., 56: 3199—3202, 1996. 199. Weinstein, J. N., Myers, T. 0., O'Connor, P. M., Friend, S. H., Fornace, A. J., Jr., 173. Lowe, S. W., Ruley, H. E., Jacks, T., and Housman, D. E. p53-dependent apoptosis Kohn, K. W., Fojo, R., Bates, S. E., Rubinstein, L. V., Anderson, N. L., Buolomwini, modulates the cytotoxicity of anticancer agents. Cell, 74: 957—967,1993. J. K., van Osdol, W. W., Monks, A. P., Scudiero, D. A., Sausville, E. A., Zaharevitz, 174. Lowe, S. W. Cancer therapy and p53. Cun. Opin. Oncol., 7: 547—553,1995. D. W., Bunow, B., Johnson, 0. S., Wines, R. E., and Paull, K. D. An information 175. Hawkins, D. S., Demers, G. W., and Galloway, D. A. Inactivationof p53 enhances intensive approach to the molecular pharmacology of cancer. Science (Washington sensitivity to multiple chemotherapeutic age. Cancer Ret., 56: 892—898, 1996. DC), in press, 1997. 176. Canman, C. E., Gilmer, T. M., Couus, S. B., and Kastan, M. B. Growth factor 200. Koutsoukos, A. D., Rubinstein, L. V., Faraggi, D., Kalyandrug, S., Weinstein, J. N., modulation of p53-mediated growth arrest versus apoptosis. Genes Dcv., 9: 600— Paull, K. D., Kohn, K. W., and Simon, R. M. Discrimination techniques applied to 611,1995. the NCI in vitro antitumor drug screen: predicting biochemical mechanism of action. 177. Smith, M. L., and Fomace, A. J., Jr. The two faces of tumor suppressor p53. Am. J. Stat. Med., 13: 719—730, 1994. Pathol., 148: 1019—1022,1996. 201 . Lefebure, V. The Riddle of the Rlsine. Chemical Strategy in Peace and War, p. 282. 178. Polyak, K., Waldman, T., He, T-C., Kinzler, K. W., and Vogelstein, B. Genetic New York: Chemical Foundation, Inc., 1923. determinants of p53-induced apoptosis and growth arrest. Genes Dcv., 10: 1945— 202. Cain, B. F., Atwell, 0. J., and Seelye, R. N. Potential antitumour agents. 11. 1952, 1996. 9-anilinoacridines. J. Med. Chem., 14: 311—315,1971. 179. Fan, S., el-Deiry, W. S., Bae, I., Freeman, J., Jondle, D., Bhatia, K., Fornace, A. I., 203. Cain, B. F., and Atwell, 0. J. The experimental antitumour properties of three Jr., Magrath, I., Kohn, K. W., and O'Connor, P. M. p53 gene mutations are congeners of the acridylmethanesulphonanilide (AMSA) series. Eur. J. Cancer., 10: associated with decreased sensitivity of human lymphoma cells to DNA damaging 539—549,1974. agents. Cancer Ret., 54: 5824—5830,1994. 204. Smith, M. L, Than, Q., Baa, I., and Fomace, A. J., Jr. Role of retinoblastomagene 180. O'Connor, P. M., Jackman, J., Jondle, D., Bhatia, K., Magrath, I., and Kohn, K. W. productinp53-mediatedDNAdamageresponse.Exp.CellRet.,215:386-389,1994. Role of the p53 tumor suppressor gene in cell cycle arrest and radiosensitivity of 205. Smith, M. L., and Fomace, A. J., Jr. Mammalian DNA damage-inducible genes Burkitt's lymphoma cell lines. Cancer Ret., 53: 4776—4780, 1993. associated with growth arrest and apoptosis. Mutat. Ret., 340: 109—124,1996.

5546

Downloaded from cancerres.aacrjournals.org on September 27, 2021. © 1996 American Association for Cancer Research. Beyond DNA Cross-Linking: History and Prospects of DNA-targeted Cancer Treatment−−Fifteenth Bruce F. Cain Memorial Award Lecture

Kurt W. Kohn

Cancer Res 1996;56:5533-5546.

Updated version Access the most recent version of this article at: http://cancerres.aacrjournals.org/content/56/24/5533

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

Reprints and To order reprints of this article or to subscribe to the journal, contact the AACR Publications Subscriptions Department at [email protected].

Permissions To request permission to re-use all or part of this article, use this link http://cancerres.aacrjournals.org/content/56/24/5533. Click on "Request Permissions" which will take you to the Copyright Clearance Center's (CCC) Rightslink site.

Downloaded from cancerres.aacrjournals.org on September 27, 2021. © 1996 American Association for Cancer Research.