Biological and biochemical properties of new anticancer folate antagonists

20
Cancer and Metastasis Reviews 5: 251-270, 1987 © Martinus Nijhoff Publishers, Boston - Printed in the Netherlands. Biological and biochemical properties of new anticancer folate antagonists David W. Fry & Robert C. Jackson Warner-Lambert/Parke-Davis Pharmaceutical Research, Ann Arbor, Michigan 48105, USA Keywords: antifolates, 10-ethyl-10-deazaaminopterin, N10-propargyl-5,8-dideazafolic acid, 5,10-dideaza- tetrahydrofolic acid, piritrexim, trimetrexate Summary We review the biology and biochemical pharmacology of four antifolates that were recently introduced into clinical trial as anticancer agents, and one compound in preclinical development. Toxicology and clinical data are not discussed. 10-Ethyl-10-deazaaminopterin (10-EdAM) is a classical antifolate, structurally related to methotrexate (MTX) but with greater activity against murine tumors. 10-EdAM has more efficient membrane transport, and relatively greater polyglutamylation in murine tumors than in normal mouse tissues, and these differential effects are greater for 10-EdAM than for other 10-deaza antifolates or for MTX. Trimetrexate and piritrexim are nonclassical antifolates, lacking a glutamate substitution. They are lipophilic, cross cell membranes more rapidly than does MTX, and retain activity against tumors resistant to MTX because of impaired drug transport. These nonclassical antifolates are active against several MTX- insensitive murine tumors, and both have demonstrated clinical anticancer activity. 10-EdAM, trimetrexate and piritrexim all inhibit dihydrofolate reductase (DHFR) as their primary site of action. As such, they deplete cellular thymidylate and purine pools, and inhibit DNA replication. N10-Propargyl-5,8-dideazafolic acid (CB3717) differs from the first three compounds in acting primarily on thymidylate synthase. Like DHFR inhibitors, it blocks DNA replication through depletion of dTTP, but it does not exert an antipurine effect. CB3717 retains activity against transport-defective MTX-resistant cells, and also against cells that overproduce DHFR. 5,10-Dideazatetrahydrofolic acid (DDATHF) is a selective inhibitor of glycinamide ribotide transformylase, and its biochemical pharmacology may differ appreciably from that of the other antifolates under study. DDATHF has strong antitumor activity in several murine systems. Introduction Since the introduction of methotrexate (MTX) into clinical cancer chemotherapy in 1949, its utility has been demonstrated in treatment both of leukemias and a range of solid tumors. It has been shown to be therapeutically superior to the earlier agent, aminopterin, against experimental and human tu- mors. The limitations of MTX are its incomplete tumor spectrum, the rather frequent development of resistance, and its acute and chronic toxicity. These limitations are shared, in greater or lesser degree, by all anticancer drugs, but they raise the possibility that better analogues, or at least analo- gues with a different antitumor spectrum, could be discovered. Many newer antifolates have been syn- thesized in attempts to produce agents that are more effective and less toxic than MTX. Recent Address for offprints: D.W. Fry, Warner-Lambert/Parke-Davis Pharmaceutical Research, P.O. Box 1047, Ann Arbor, M148106-1047, USA

Transcript of Biological and biochemical properties of new anticancer folate antagonists

Page 1: Biological and biochemical properties of new anticancer folate antagonists

Cancer and Metastasis Reviews 5: 251-270, 1987 © Martinus Nijhoff Publishers, Boston - Printed in the Netherlands.

Biological and biochemical properties of new anticancer folate antagonists

David W. Fry & Robert C. Jackson Warner-Lambert/Parke-Davis Pharmaceutical Research, Ann Arbor, Michigan 48105, USA

Keywords: antifolates, 10-ethyl-10-deazaaminopterin, N10-propargyl-5,8-dideazafolic acid, 5,10-dideaza- tetrahydrofolic acid, piritrexim, trimetrexate

Summary

We review the biology and biochemical pharmacology of four antifolates that were recently introduced into clinical trial as anticancer agents, and one compound in preclinical development. Toxicology and clinical data are not discussed. 10-Ethyl-10-deazaaminopterin (10-EdAM) is a classical antifolate, structurally related to methotrexate (MTX) but with greater activity against murine tumors. 10-EdAM has more efficient membrane transport, and relatively greater polyglutamylation in murine tumors than in normal mouse tissues, and these differential effects are greater for 10-EdAM than for other 10-deaza antifolates or for MTX. Trimetrexate and piritrexim are nonclassical antifolates, lacking a glutamate substitution. They are lipophilic, cross cell membranes more rapidly than does MTX, and retain activity against tumors resistant to MTX because of impaired drug transport. These nonclassical antifolates are active against several MTX- insensitive murine tumors, and both have demonstrated clinical anticancer activity. 10-EdAM, trimetrexate and piritrexim all inhibit dihydrofolate reductase (DHFR) as their primary site of action. As such, they deplete cellular thymidylate and purine pools, and inhibit DNA replication. N10-Propargyl-5,8-dideazafolic acid (CB3717) differs from the first three compounds in acting primarily on thymidylate synthase. Like DHFR inhibitors, it blocks DNA replication through depletion of dTTP, but it does not exert an antipurine effect. CB3717 retains activity against transport-defective MTX-resistant cells, and also against cells that overproduce DHFR. 5,10-Dideazatetrahydrofolic acid (DDATHF) is a selective inhibitor of glycinamide ribotide transformylase, and its biochemical pharmacology may differ appreciably from that of the other antifolates under study. DDATHF has strong antitumor activity in several murine systems.

Introduction

Since the introduction of methotrexate (MTX) into clinical cancer chemotherapy in 1949, its utility has been demonstrated in treatment both of leukemias and a range of solid tumors. It has been shown to be therapeutically superior to the earlier agent, aminopterin, against experimental and human tu- mors. The limitations of MTX are its incomplete

tumor spectrum, the rather frequent development of resistance, and its acute and chronic toxicity. These limitations are shared, in greater or lesser degree, by all anticancer drugs, but they raise the possibility that better analogues, or at least analo- gues with a different antitumor spectrum, could be discovered. Many newer antifolates have been syn- thesized in attempts to produce agents that are more effective and less toxic than MTX. Recent

Address for offprints: D.W. Fry, Warner-Lambert/Parke-Davis Pharmaceutical Research, P.O. Box 1047, Ann Arbor, M148106-1047, USA

Page 2: Biological and biochemical properties of new anticancer folate antagonists

252

reviews of these synthetic efforts have been pub- lished by Roth and Cheng [1], Montgomery and Piper [2] and Werbel [3]. Some of the more inten- sively studied antifolates produced during the 1950s and 1960s were metoprine (DDMP), methasqin, triazinate, and 5-methyltetrahydrohomofolate. Properties of these agents have been summarized in references [2] and [3], and in articles by Hill and Price [4], and Hutchison and Schmid [5]. Despite the large number of classical (glutamate-contain- ing) and nonclassical (lipophilic, non-glutamate- containing) antifolates produced during this period, and the fact that occasional clinical re- sponses were seen to some of these compounds, none of them appeared to have clear advantages over MTX.

Over the last fifteen years much detailed infor- mation on the mechanism of action of antifolates has made MTX one of the most studied and best understood of all anticancer drugs. The wealth of new data includes biochemical and kinetic studies on MTX and polyglutamylation (reviewed by Sir- otnak and DeGraw, 6), detailed X-ray crys- tallographic and NMR studies of the interaction of DHFR with MTX (reviewed by Freisheim and Matthews, 7), cell cycle effects studied by flow cytometry [8] and measurements of the MTX-in- duced perturbations in cellular pools of ribonucle- otides and deoxyribonucleotides [9]. The new bio- logical and biochemical data suggest ways of screening novel antifolates for greater therapeutic selectivity, activity against MTX-resistant tumor cells, or against sites of action other than DHFR. The high resolution X-ray crystallography studies [7] have already suggested convincing explanations for the tight-binding inhibition of DHFR by 2,4- diamino analogues of folic acid, and for the anti- bacterial selectivity of trimethoprim. Crystallogra- phy studies open up the possibility of computer- assisted drug design of more potent and more selec- tive inhibitors of folate enzymes.

The structures of MTX and of the five novel folate antagonists discussed in this review are shown in Figure 1.10-Ethyl-10-deazaaminopterin is a classical antifolate synthesized by DeGraw et al. [10]. Trimetrexate is one of a series of quinazoline antifolates synthesized in the Parke-Davis labora-

,OH2 A COOH

NH2 CH3CHz N.,-J~-.-N- Cl-I~ ~ QOOH ~,~y, rNy H ~ - C O - N H - C H

CO~H B

"'"AVAV-J bc. c

NH~ CH~ .OC H~

OCH~ 0 HC--C-CH2 J[ ---- -CH~ I ~ ,COOH

HN" Y ~ "~l--(( I~'CO-NH-CH

H2N" "N" Y/"~ ~H~) H E

0 I

0 NH C~OHm

Fig. 1. Structures of methotrexate and of new antifolate anti- cancer drugs. A. methotrexate (MTX); B. 10-ethyl-10-de- azaaminoptcrin (10-EdAM); C. trimetrexate (TMQ or JB-11); D. piritrexim (BW 301U or PTX); E. Nl°-propargyl-5,8-di - deazafolic acid (CB3717); G. 5,10-dideaza-5,6,7,8-tetrahydro- folic acid (DDATHF).

tories [11, 12]. Another lipophilic antifolate is the 2,4-diaminopyridopyrimidine, piritrexim, de- scribed by the Burroughs-Wellcome group [13]. Trimetrexate and piritrexim have some close simi- larities, but also interesting differences, as de- scribed below. CB3717 [14] differs from the pre- vious compounds in having an oxygen substitution at position 4. This is also true of 5,10-di- deazatetrahydrofolic acid [15], which also differs from most other antifolates studied in having a tetrahydro structure.

lO-Ethyl-lO-deazaaminopterin

lO-Ethyl-lO-deaza-aminopterin (IO-EdAM) is one

Page 3: Biological and biochemical properties of new anticancer folate antagonists

of several derivatives synthesized by DeGraw et al. [10] in which the Nl°-amine of aminopterin is re- placed by a methylene group. The initial com- pound produced in this series, 10-deaza-aminop- terin (10-dAM) [16], is clearly superior to meth- otrexate in several murine tumor models [17, 18] and eventually went to clinical trial [19]. The ra- tionale for synthesis of the 10-ethyl derivative is the concept that membrane transport is an important determinant of cytotoxicity [20, 21]. Previous work [20] shows that the Nl°-alkyl derivatives of aminop- terin are transported much less efficiently in nor- mal tissues, such as gut, than in tumor tissue and therefore have the potential for greater selectivity and perhaps an enhanced therapeutic index. The results of an initial study [10] comparing in vivo antitumor properties in L1210 leukemia, inhibition of dihydrofolate reductase, and the transport characteristics in L1210, of methotrexate, aminop- terin 10-dAM and 10-EdAM are consistent with these predictions. Although there were only minor differences in K i for inhibition of dihydrofolate reductase between these analogs, 10-EdAM ex- tended the life span of tumor-bearing mice to a greater extent. An analysis of the kinetic param- eters for transport of these analogs in and out of L1210 cells shows that the influx Vma x and effiux rate constant for 10-EdAM are identical to that ob- tained for the other agents. The influx Km, how- ever, is 4-fold lower than that for methotrexate and is similar or perhaps slightly lower than aminop- terin and 10-dAM. These preliminary observations indicate that alkylation of the N ~° position of aminopterin reduces the affinity for this transport carrier in tumor cells whereas alkylation of the C ~° position in 10-dAM did not produce this effect. Substitution of either position does not alter bind- ing affinity to dihydrofolate reductase.

Subsequent to these initial data more extensive work was completed on the transport as well as the metabolism of 10-EdAM [22]. These studies show that influx is saturable for 10-EdAM with a Km of approximately l uM in L1210 leukemia, Ehrlich carcinoma and Sarcoma 180. In all cell lines 10-EdAM exhibits a lower K m for influx than meth- otrexate and reinforces the conclusion that 10-EdAM and other analogs of this series, as well

253

as aminopterin, are better substrates for this carrier system than methotrexate. Consistant with these data is the observation that the exchangeable steady-state concentrations of these analogs are 3--4 times higher than methotrexate. Secondly, even though there are large differences in the K m of influx for methotrexate in different cell lines, the parameter remains constant for 10-EdAM in L1210, Ehrlich and S180 cells. Since the limiting toxicity in mice is considered to be associated with antifolate effects in crypt cell epithelium of the small intestine [23], kinetic parameters for the transport of 10-EdAM were also determined in isolated murine intestinal epithelial cells. Several relevant observa- tions are notable from this study. First, the K m values for influx for all compounds both N ~° and 10-deaza are much higher than in tumor cells. Sec- ond, chemical modification from N 1° to 10-deaza as well as alkylation at the C l° position contribute to a significantly lower affinity for the carrier in the mucosal cells and, since this lowered affinity did not occur in tumor cells, these characteristics pro- vide a basis for selectivity. The study shows that compounds demonstrating the greatest ratio for influx K m Of gut epithelial cells to tumor cells are the most effective in prolonging survival in tumor bearing mice [24]. Similar relationships between methotrexate, aminopterin, 10-EdAM and 10-dAM with regard to transport kinetic constants and cellular accumulation are found in the estab- lished human tissue culture cell lines Manca (human B-cell Leukemia) and SKBRlll (human mammary carcinoma) and human ovarian tumor cells obtained from patients as a suspension from malignant effusions [25].

An area of intense research in recent years is the metabolism of classical antifolates to their poly- glutamyl derivatives and the relationship between the formation of these metabolites and their selec- tivity and antitumor efficacy. The general concepts emerging from this work have been recently re- viewed [21, 26, 27].

Initial studies examine the formation of poly- glutamyl derivatives of 10-EdAM as well as 10-dAM in L1210 cells, S180 cells and the small intestine four hours after administration to tumor bearing mice [22]. All analogs are substrates for

Page 4: Biological and biochemical properties of new anticancer folate antagonists

254

folypolyglutamate synthetase but net accumulation of the metabolites derived from all analogs is con- sistently lower in the small intestine than in the tumor cells. In the small intestine the extent of polyglutamylation is in the order aminopterin > 10-EdAM >methotrexate >10-dAM (2.8: 0.9: 0.5: 0.3) and consists mainly of the triglutamate with lower levels of di- and heptaglutamates. The same relative order of polyglutamylation is found in L1210 cells. In S180 the order of 10-EdAM and aminopterin is switched. In tumor cells the major metabolites are the tri- and heptaglutamates.

Further studies [28] examine the relationship be- tween transport, polyglutamylation and cytotox- icity of 10-EdAM in 3 cell lines in tissue culture. The kinetics of the formation of each derivative is determined in each cell line and the accumulation of these metabolites reaches 85-95% of the total cellular drug in 24 hours in L1210 and Manca cells, however $180 cells require only 6 hours to attain these levels. Polyglutamates of 10-EdAM consist- ing of 1, 2 and 3 or more additional glutamyl resi- dues appear in sequential order and by 24 hours the predominate forms contain 3 or more polygluta- mates. The extent of polyglutamation for 10-EdAM in all cell types is consistently greater than methotrexate and 10-dAM. The retention of these metabolites is similar to earlier findings for methotrexate in that no efflux of higher polygluta- mates is observed and, although there is some loss of lower derivatives, it is slower than the loss of monoglutamate. The significance of these observa- tions is reflected in the good correlation between the cytotoxicity of the various analogs and their potential for polyglutamylation.

The fluorometric high-performance liquid chro- matographic assay [29, 30] takes advantage of the high native fluorescence exhibited by pteridine- containing compounds which have carbon in the 10 position. This technique was shown to be suitable for analysis of 10-EdAM, its analogs, and poly- glutamyl derivatives in plasma, urine and tissue. The HPLC assay is therefore quite useful for phar- macokinetic studies in patients treated with 10-EdAM.

In summary, the structural modifications and the resulting apparent benefits of 10-EdAM as corn-

pared to methotrexate are as follows: (1) The re- placement of the N ~° with methylene (a) increases its affinity for the folate carrier system in tumor cells (b) decreases the affinity in gut epithelial cells. (2) The substitution of ethyl for methyl at the C 1° position (a) retains affinity for the transport system in tumor cells (alkyl substitution at the N 1° position reduces affinity) (b) further reduces affinity for the influx carrier in gut, and (c) increases poly- glutamylation in tumor cells with little increase of polyglutamylation in gut cells.

10-EdAM is a tight-binding competitive inhibitor of dihydrofolate reductase, with K~ of 2 pM, similar to that of aminopterin and 10-deaza-aminopterin, and slightly tighter than methotrexate [22]. 10-EdAM is only a weak inhibitor of thymidylate synthase. Ueda et al. [31], studying thymidyla~e synthase from human leukemia cells, reports that when 5,10-methylenetetrahydrofolate is used as the folate cofactor, 10-EdAM inhibition is competi- tive with this cofactor, with K~ of 410/zM. Poly- glutamates of 10-EdAM are also competitive, with K i values 50-138-fold tighter than the parent com- pound. However, when 5,10-methylenetetra- hydrofolate pentaglutamate is used as substrate, inhibition by 10-EdAM and its diglutamate is non- competitive, and the 10-EdAM pentaglutamate is a mixed-type inhibitor. Kisliuk et al. [32] studied inhibition of thymidylate synthase from Lac-

tobacillum casei. The IC50 concentration for 10-EdAM is greater than 300/xM. Addition of glu- tamate residues to 10-EdAM increases the inhibi- tory potency, with an ICs0 value for the pentagluta- mate of 7.4/xM. These values are considerably weaker than for methotrexate and its poly- glutamates, and suggest that inhibition of thy- midylate synthase does not contribute to the phar- macological effect of 10-EdAM.

At the time of writing, no published data are available concerning effects of 10-EdAM on cellu- lar nucleotide pools, nor on cell-cycle phase depen- dence. It seems likely that 10-EdAM will generally resemble methotrexate in these respects.

Mandelbaum-Shavit [33] studied the effect of 10-EdAM on MCF-7 human breast cancer cells in vitro. The IC50 concentration is 6 nM, compared to 20 nM for methotrexate. Potencies of the two com-

Page 5: Biological and biochemical properties of new anticancer folate antagonists

pounds as inhibitors of dihydrofolate reductase from MCF-7 cells are similar, with ICs0 of 4.2 nM for methotrexate and 3.9nM for 10-EdAM. It should be noted that this kind of analysis would not distinguish between quite large differences in K i values when the Ki is in the picomolar range. Based upon inhibition of uptake of radiolabelled meth- otrexate, Mandelbaum-Shavit finds a K i for trans- port of 2.4/xM, for 10-EdAM, compared t o K m

values for methotrexate and 5-methyltetra hydro- folate of 7.8 ~M and 3.2/xM, respectively. These results suggest that the greater cytotoxicity to MCF-7 cells of 10-EdAM in comparison with meth- otrexate is attributable to its higher affinity for the transport carrier.

In vivo studies by Sirotnak et al. [24, 34] show that 10-EdAM is superior to methotrexate in four murine ascites tumors (L1210 leukemia, sarcoma 180, Ehrlich ascites carcinoma, and Tapper car- cinosarcoma) and has similar activity to MTX in two other ascites tumors (P288 and C1498 leuke- mias). 10-EdAM gives a proportion of long-term survivors in the L1210 and sarcoma 180 systems. Against subcutaneous solid tumor implants, 10-EdAM is superior to MTX in four of six systems (sarcoma 180, Tapper, E0771 mammary adenocar- cinoma, and T241 fibrosarcoma). 10-EdAM and MTX have similar, modest activity against Lewis lung carcinoma, and both agents are inactive against the B16 melanoma.

Schmid et al. [35] compare the activity of 10-EdAM and MTX against subcutaneous implants of three human tumor xenograft lines in nude mice. In all cases drugs are administered by intra- peritoneal injection daily for five days, beginning two days after implantation. Tumor size is evalu- ated on day 11. Against the MX-1 breast carcinoma, MTX is inactive, and on day 11 tumor mass relative to control (T/C) is 8% for animals treated with 10-EdAM at dose of 2.2 mg/kg, with 2 of 9 animals tumor-free. In the more refractory LX-1 lung car- cinoma, MTX at its optimal dose gives T/C of 51%, and 10-EdAM at dose of 4.5 mg/kg gives T/C of 12%. In the highly refactory CX-1 colon carcinoma xenograft, MTX is totally inactive at all tolerated doses, and 10-EdAM at dose of 3.0 mg/kg gives T/C of 33%. Higher doses of 10-EdAM are toxic. In all

255

three xenograft systems the tumor regresses to some extent, i.e. its size is smaller on the evaluation day than at the start of treatment. Sirotnak [31] reports that 10-EdAM has activity superior to that of MTX in two other human tumor xenografts (SK melanoma 109 and MBA 9812 non-small cell lung cancer). MTX and 10-EdAM have similar activity against two other xenografts. It has been suggested that the marked therapeutic superiority of 10-EdAM relative to MTX against murine tumors and human tumor xenografts may be attributable to its more efficient transport into tumor cells (but not into mouse intestinal epithelium cells) and to the more extensive polyglutamylation of 10-EdAM in tumor cells.

The optimal doses for 10-EdAM are very similar to those for MTX. Kinahan et al. [36] report pre- liminary toxicology findings and pharmacokinetics for 10-EdAM. The terminal plasma half-life in dogs is 9 hours. In rats, 50% of the drug is excreted in the bile, much of it as the 7-hydroxy derivative. Poly- glutamates of 10-EdAM are found in liver, small intestine, and kidney of treated rats soon after drug administration, but decline rapidly thereafter.

Trimetrexate

Trimetrexate (2,4-diamino-5-methyl-6 [(3,4,5-tri- methoxyanilino) methyl] quinazoline), also re- ferred to in some literature as TMQ or JB-11 is one of a series of 2,4-diaminoquinazolines synthesized in an effort to circumvent certain types of resis- tance associated with classical antifolates including transport impairment and poor penetration of the blood-brain barrier [3, 11, 12].

A number of studies have demonstrated, at least in vitro, that trimetrexate is active against meth- otrexate-resistant cells having a defective folate transport system. This has been shown in sublines of several cell types including a human acute lym- phocytic T-cell line, MOLT-3 [37, 38], a human T-lymphoblast, CCRF-CEM [39, 40], a human os- teosarcoma line, SAOS-2 [39], and a human squamous cell carcinoma (SCC15) [41] as well as murine L1210 leukemia [42-44].

Although the above data repeatedly demon-

Page 6: Biological and biochemical properties of new anticancer folate antagonists

256

strate that trimetrexate is active against methotrex- ate transport impaired cells and therefore must enter by a mechanism distinct from the reduced folate carrier system, few studies have appeared that attempt actually to define what that mecha- nism might be. The first account describes some properties of trimetrexate transport in L1210 cells [42]. Uptake was found to be 30-fold faster than methotrexate and proportional to the extracellular concentration of trimetrexate between the concen- trations of 10 -9 and 10 -4 M. Uptake is not inhibited by 20 uM p-chloromecuribenzoate (PCMB), 10 mM sodium azide, the natural folates folic acid or 5-methyltetrahydrofolate or the folate analogs 2,4-diaminoquinazoline, trimethoxyaniline and methotrexate at concentrations of 100/~M each. Uptake of trimetrexate is temperature-sensitive and a Q~0 is calculated to be 2.71. Effiux of tri- metrexate is rapid and not affected by 20/~M PCMB or 10mM sodium azide. After cells are incubated with 4/~M trimetrexate for 30 minutes and then washed out, a nonexchangeable fraction is retained which is 3-5 times higher than the di- hydrofolate reductase level. No metabolites or tri- metrexate are detected under the conditions of these experiments.

A more extensive characterization of trimetrex- ate transport has been conducted in a human lym- phoblastoid cell line, WI-L2 [45, 46, D.W. Fry, unpublished results]. A plot of cell-associated tri- metrexate for these cells versus time at an extra- cellular concentration of 1/~M shows a very high intercept on the y axis implying a component of very rapid association with the cell which may rep- resent some form of surface binding. Uptake is rapid and attains a steady-state concentration of 120 pmol/107 cells within 30 minutes. Undirectional influx studies show that influx rates are 25-fold faster than methotrexate influx. At an extracellular concentration of 1 ~M the influx rate for trimetrex- ate in 107 cells is 3.11pmol/second whereas that for methotrexate in 107 cells is 0.13 pmol/second. The difference in efflux rate constants between tri- metrexate and methotrexate is not nearly as great. Trimetrexate effiux is about 2-3 times faster than methotrexate. Efflux studies show a large nonex- changeable fraction of trimetrexate that is well

above the dihydrofolate reductase binding capac- ity. This nonexchangeable fraction is not constant and is dependent on the level of drug originally associated with the cells. Influx of trimetrexate does not appear to be saturable and a plot of influx rate versus extracellular trimetrexate concentra- tion is linear up to 500 p~M. At higher concentra- tions the response deviates from linearity but even at concentrations as high as 5 mM the system does not fully saturate. Furthermore, influx cannot be inhibited to any significant extent with lmM con- centrations of methotrexate, folic acid, folinic acid, or a wide variety of natural substrates and meta- bolites including purine and pyrimidine bases and nucleosides, sugars and amino acids. Although the above data are consistant with a mechanism for transport of passive diffusion other data are some- what compatible with a carrier mediated process or at least a process more complex than simple diffu- sion. This possibility is supported by the fact that the infux and exchangeable cellular steady-state levels of trimetrexate are reduced by 50% in cells preincubated 30 minutes with 10 mM sodium azide. Glucose, however, does not reverse this effect and indeed reduces the steady-state levels of trimetrex- ate by 50% itself, indicating that these effects on

trimetrexate transport may not be strictly a func- tion of altered energy metabolism. Influx can be completely abolished with 50~M p-(chloromer- curi)benzene-sulfonic acid (PCMBS) or by lower- ing the temperature to 4 °C. The difference be- tween influx rates at 37 ° and 27 ° C is about 2.5 fold. Efflux of trimetrexate is also sensitive to metabolic inhibitors in that 10 mM sodium azide stimulated efflux by 2- to 3-fold and this effect could be par- tially reversed by 10 mM glucose. Finally, the ex- changeable intracellular concentration of tri- metrexate at the steady-state is always 4 to 5 times that of the extracellular concentration.

Previous literature dealing with transport of lipophilic antifolates does little to resolve the poorly understood aspects of trimetrexate trans- port since a number of discrepancies are evident. Kamen et al. [42] (see above) studied trimetrexate transport in L1210 and while most observations are consistant with those reported in the WI-L2 study, he finds no effect with sodium azide or pCMB.

Page 7: Biological and biochemical properties of new anticancer folate antagonists

Duch et al. [47] finds that the rate of uptake of PTX at 4 ° C was the same as at 37 ° degrees which is quite unlike trimetrexate. Hill et al. [48, 49] finds that in L5178Y lymphoblasts, metoprin e transport is highly temperature sensitive, appears to be satura- ble with a K m for influx of 130/zM and is inhibited by folinic acid. Greco and Hakala [50], in contrast, observe in the human carcinoma cells, KB and Hep-2, that uptake of this drug is not sensitive to temperature nor saturable at least at the concentra- tions employed (1-200 p.M) and is not inhibited by folinic acid. These studies complicate any attempt to draw firm conclusions concerning a specific mode of transport for these classes of compounds.

Additional data, however, which indicate a more intricate transport system for trimetrexate than passive diffusion may be derived from the fact that cells are capable of altering the rate of trans- port and thus their sensitivity to trimetrexate. Re- ports [45, 46, D.W. Fry, unpublished results] de- scribing the selection of a trimetrexate-resistant line of human lymphoblastoid cells (WI-L2) by exposing the cells to step-wise increases in drug concentration show that lymphoblastoid cells are 40- to 60-fold resistant to trimetrexate and fully cross-resistant to other lipophilic antifolates in- cluding metoprine and BW 301U. However, they are completely sensitive to methotrexate and only 3-fold cross-resistant to methotrexate-gamma- tertbutylmonoester. Furthermore, these cells are completely sensitive to adriamycin and vincristine implying that this is not a form of pleiotropic or multidrug resistance. Although the resistant cells exhibit a slight elevation in dihydrofolate reductase (2-fold), it seems unlikely that this could account for a 60-fold resistance. Nor are there any differen- ces in the affinities of the enzyme for trimetrexate between the two cell lines. Studies investigating the transport of 14C labeled trimetrexate show a 50% decrease in the influx rate of the resistant relative to the sensitive cell line but equal efflux rates. This correlates with a 50% drop in exchangeable tri- metrexate observed in the resistant line at the steady-state. Transport differences between the two, cell lines are not due to changes in intracellular water and rates of methotrexate transport are iden- tical in each. Furthermore, analysis of the intra-

257

cellular compartment of both cells by high pressure liquid chromatography indicated that no detecta- ble metabolites are formed when cells are incu- bated for up to six hours with 5/zM trimetrexate.

Resistance to trimetrexate possibly due to im- paired transport has also been described in a line of human acute lymphoblastic leukemia, Molt-3 [51]. These cells are 200-fold resistant to trimetrexate and exhibit influx rates 40% of the parent line and steady-state levels of only 23%. Additional abnor- malities are present in these cells, however, includ- ing a 6-fold cross-resistance to methotrexate, a 50% reduction in influx of this agent and a 75% decrease in the intracellular levels at the steady- state. Finally, some cell lines appear to take up trimetrexate intrinsically to lesser or greater ex- tents with corresponding effects on sensitivity. Mini et al. [52] reports that CCRF/CEM cells are 10-fold less sensitive to trimetrexate than HCT-8 cells which corresponded to a 10-fold difference in drug accumulation after four hours. Borman et al. [53] finds that levels of trimetrexate in rat colon adenocarcinoma are about 4-fold the level in L1210 after similar incubation conditions and 10- to 100- fold higher than that observed for methotrexate. The enhanced accumulation of trimetrexate correl- ates with its relative efficacy in experimental colon carcinoma in vivo.

Although multidrug resistance is not associated with the above examples of resistance, recent re- ports imply that trimetrexate and other lipophilic antifolates have this form of resistance. Klohs et al. [54] find that a pleiotropically resistant line of P388 leukemia is cross-resistant to trimetrexate and BW301U and that resistance can be partially re- versed by verapamil and tween-80. Transport stud- ies show that net uptake of trimetrexate is de- pressed by 75% in the resistant line and that intracellular steady-state levels can be restored by the presence of verapamil. These studies indicated that pleiotropic resistance extends to lipophilic antifolates and that the mechanism appears to be alterations in transport similar to what has been observed for adriamycin. Similar results have been observed elsewhere [55]. Multidrug resistance is also found in a line of human acute lymphoblastic leukemia (MOLT-3) made resistant to trimetrex-

Page 8: Biological and biochemical properties of new anticancer folate antagonists

258

ate [56]. These cells are 800-fold resistant to tri- metrexate but are also 280-fold cross-resistant to vincristine, 15-fold to daunorubicin, 14-fold to dox- orubicin and 10-fold to mitoxantrone. Cross resis- tance to these drugs is greatly diminished by 10 -5 M verapamil. Cellular uptake of trimetrexate and vin- cristine in these cells is diminished 82% and 80% respectively. Uptake of both drugs is enhanced in the presence of verapamil.

These observations imply that although tri- metrexate and other lipophilic antifolates are quite capable of circumventing methotrexate resistance due to an impairment of the reduced folate carrier system, the nature of these structures evidently makes them susceptible to at least 2 other forms of resistance related to altered transport. Some stud- ies indicate that since trimetrexate and methotrex- ate elicite different modes of resistance, alternating use of these drugs may delay onset of resistance to antifolates [57].

Many of the biochemical effects of trimetrexate are typical of most antifolates. Trimetrexate has been shown to be a potent inhibitor of dihydrofol- ate reductase from several cell types. Earlier stud- ies do not report true K~ values but express inhibi- tion as the concentration of drug necessary to inhibit by 50% (ICs0). Thus the ICs0 values for dihydrofolate reductase from human acute lym- phocytic leukemia cells and L1210 cells are 1.9 x 10 -9 and 1.2 x 10 -9 M respectively and are compar- able to methotrexate at 9 × 10-~°M for both cells lines [58, 59]. Similar results are obtained for the enzyme from beef liver [60]. A more detailed kine- tic analysis of the inhibition of dihydrofolate reduc- tase are conducted with enzyme purified from WI-L2 cells [44]. A Ki.slop e value of 4.3 × 10 -11M are determined for trimetrexate which is about 10-fold greater than that obtained for methotrexate.

Trimetrexate is also a potent inhibitor of deox- yuridine incorporation into DNA [44, 61-63] and, similar to other antifolates, the cytotoxic effects of trimetrexate can be reversed by leucovorin, thy- midine or a combination of thymidine and hypox- anthine. Broome et al. [62] finds that the cytotoxic effects of 10txM trimetrexate can be reversed by 100/xM leucovorin in P388 and in the M5076 ovarian tumor, the antitumor effects of trimetrex-

ate are reversed by coadministration of leucovorin in vivo. Complete protection from trimetrexate by leucovorin has also been shown in the human lym- phoid cell lines CCRF-CEM and LAZ-007 [65]. Other reports [43] show that protection from both growth inhibition and cytotoxicity of ltxM tri- metrexate in L1210 or WI-L2 cells can be attained when 40~M leucovorin is added either simul- taneously or as late as 5 hours after trimetrexate. McCormack et al. [65] show that 4-6txM thy- midine alone can prevent the growth inhibition produced in L1210 cells by 2 x 10 -8 M trimetrexate. Further studies [44] show that the protection by thymidine alone is dependent on the concentration of trimetrexate and that if concentrations of drug are 10 -7 M or below, thymidine can fully protect. If, however, concentrations of trimetrexate are greater than 10-7M a combination of thymidine and hypoxanthine is necessary.

The requirement that both thymidine and hypo- xanthine be present to protect cells from trimetrex- ate cytotoxicity suggests both an antipyrimidine and antipurine effect. Direct measurement of pu- rine and pyrimidine ribonucleotide pools has been made in L1210 and WI-L2 cells preincubated with 50 nM or 1/xM trimetrexate for 6 hours [44]. Both adenylates and guanylates are reduced to approx- imately the same extent and the depression in these pools is dependent on drug concentration. In con- trast pyrimidine nucleotides, CTP, UTP, and UDP are elevated, most probably due to a reduced rate of RNA synthesis. Qualitatively similar results are obtained in both cell lines although the reduction in purines is much more pronounced in the L1210 cell. In L1210 cells treated 2 hours with 1/xM trimetrex- ate a strong antithymidylate effect is apparent as dTTP levels drop to 29% of the control. The levels of dATP and dGTP also drop significantly to 41% and 15% of control respectively. No significant change is observed in the dCTP pool.

A recent report has examined the cell cycle phase specificity of trimetrexate [66]. Cultures of Chinese hamster ovary (CHO) cells were syn- chronized by mitotic selection and were exposed periodically throughout the cell cycle to 50/xM tri- metrexate for 2 hours. The surviving fractions as determined by a cloning assay show that only those

Page 9: Biological and biochemical properties of new anticancer folate antagonists

cells in S phase are sensitive to trimetrexate. Fur- thermore, CHO cells in plateau phase are rela- tively insensitive to trimetrexate when compared to log phase cells. In other experiments exponen- tionally growing L1210 leukemia cells were con- tinuously exposed to either 3 or 30 nM trimetrexate and their progression through the cell cycle moni- tored by flow cytometry. Cells exposed to 3 nM trimetrexate exhibit a delay in progression through S phase whereas 30 nM produce cell cycle arrest in late G1 or early S phase.

Trimetrexate and methotrexate were evaluated side by side in the human tumor cloning system [67]. A total of 69 (51%) human tumors, obtained from a total of 136 human tumors representing 19 different histological types were prepared and evaluated. A comparison of the relative sensitivity of the two drugs revealed that 15% (8 of 54) of the tumors were sensitive (greater than 50% cell kill) when exposed for 1 hour to 0.1~g/ml of trimetrex- ate but resistant to 0.3 ~g/ml methotrexate whereas 7% (4 of 54) were sensitive to methotrexate but refractory to trimetrexate.

A number of studies have evaluated the anti- tumor properties of trimetrexate especially in com- parison to methotrexate. In general, trimetrexate shows a much broader spectrum of activity especi- ally against solid tumors. Initial studies [59] show trimetrexate to have moderate to good activity against L1210 and P388 leukemias, colon 26 and 38, and B16 melanoma whereas methotrexate is active against just the leukemias. Subsequent reports confirmed the activity in L1210 [44, 68] and also demonstrate mild activity against the M5076 ovarian [44, 62]. A more recent report [69] com- pares trimetrexate and methotrexate against a wide variety of murine tumor models and human tumor xenografts. In this study trimetrexate is active against B16 melanoma, CD8F1 mammary tumor, colon 26 and 38, and L1210 and P388 leukemias. Trimetrexate has no activity against Lewis lung carcinoma. Methotrexate is active only against the CD8F1 mammary tumor cells and the leukemias. Neither drug is active against CX-1 colon, LX-1 lung or MX-1 mammary human tumor xenografts in nude mice.

In a study to determine the potential for thera-

259

peutic synergy between trimetrexate and several clinically proven agents [70, 71] life span assays were conducted in F1 hybrid mice bearing either early or advanced stage P388 leukemia. The mice were treated with each agent singly or in combina- tion. A high degree of therapeutic synergism is observed between trimetrexate and doxorubicin, cytoxan or 6-thioguanine. Combinations of tri- metrexate and 5-fluorouracil, vincristine or cisplatin also produce enhanced tumor cell kill al- though the results are not nearly as dramatic. Over- lapping host toxicities of trimetrexate and meth- otrexate appeared to negate any therapeutic benefit produced by this combination.

The pharmacokinetics of trimetrexate have been evaluated in several animal systems in addition to a number of clinical studies. In mice the elimination of trimetrexate in the plasma is reported to be a single exponential with a tv2 of 50 minutes [65]. In dogs the elimination is biphasic with an alpha t~/2 of 10-17 minutes and a beta tv2 of 2.5-3.6 hours [72, 73]. These parameters appeared to be identical in normal or lymphoma bearing dogs [74]. In Rhesus monkeys the elimination was triphasic having half lives of 3.3-3.9, 39-45, and 199-248 minutes after an IV bolus of 100mg/m 2 [75]. Further analysis shows that elimination is primarily by biotransfor- mation, that there is substantial plasma protein binding and little penetration into the CFS. Several studies in humans report biexponential plasma elimination with alpha tv_, values ranging from 40 to 78 minutes and beta tv2 values from 10 to 44 hours [76-78]. Other studies indicate a triphasic pattern with halflives of approximately 10 minutes, 2 hours and 16 hours [79, 80].

The major metabolic route for trimetrexate is considered to be oxidative demethylation in the liver and occurs via a conjugated glucuronide [81, 82]. Metabolism of trimetrexate readily occurs in the isolated perfused liver [83] and is inhibited by hypoxic conditions [84]. It also can be observed in rat liver microsomes and is inhibited by clinically used imidazole drugs [851.

Page 10: Biological and biochemical properties of new anticancer folate antagonists

260

Piritrexim

Piritrexim (PTX), 2,4.diamino-6-(2,5-dimethox- ybenzyl)-5-methylpyrido-[2,3-d]pyrimidine also referred to in the literature as BW 301U was se- lected out of over 300 heterocyclic structures based on a series of 6-substituted 2,4-diaminopyrido[2,3- d]-pyridines originally synthesized as potential antimicrobial agents [86-88]. In subsequent years the synthetic route to these structures was im- proved and it was found that PTX exhibited signifi- cant antitumor activity against Walker 256 car- cinoma [13].

One criterion for selecting PTX was its greater affinity for DHFR than a previously developed lipophilic antifolate, metoprine (DDMP). PTX in- hibits DHFR by 50% at a concentration of 5 nM which is comparable to MTX at 7 nM [11]. A second advantage PTX has over metoprine is its more favorable plasma pharmacokinetic properties [47]. A third reason for selecting PTX was its lack of inhibition on histamine metabolizing enzymes [90, 91]. Studies with partially purified enzymes show that PTX competitively inhibits histamine meth- yltransferase from bovine cerebral cortex with a K i of 1.8/xM and inhibit diamine oxidase from rat cecum in a noncompetitive manner with a K i of 32/aM. These values, however, represent a much lower affinity for the enzymes than other lipophilic antifolates such as trimetrexate which exhibits K i values for these enzymes of 0.007/aM and 4.1/xM respectively. It was shown that administration of trimetrexate to rats produces a 2-fold elevation of histamine in the brain and a 7-fold elevation in the kidney. In contrast, administration of PTX at the same dose causes no elevation in the brain and a 1.5- to 3-fold elevation in the kidney which is of short duration and returns to normal after 2 hours.

One of the more peculiar aspects of PTX as an antifolate is the nucleoside requirement for rever- sal of cytotoxicity in tissue culture. The growth inhibition of tissue culture cells by most known inhibitors of DHFR such as MTX, trimetrexate and metoprine can be fully reversed by a combina- tion of thymidine and hypoxanthine. In S180 cells, however, full protection from PTX is afforded only by a combination of hypoxanthine, thymidine, ade-

nosine and either cytidine or uridine [47, 89]. Fur- thermore, although leucovorin alone is capable of protecting cells from MTX and other antifolates, it is not sufficient in the presence of PTX. In $180 cells the cytotoxic effects of MTX or metoprine can be completely prevented by the addition of leuco- vorin at a concentration 10-fold higher than that of the drug. However, when cells are grown in a concentration of PTX sufficient to inhibit cell growth by 90%, leucovorin can reverse this inhibi- tion to no more than 60% of the control cells even at concentrations of leucovorin as high as 250 times that of extracellular drug. If, however, thymidine is included along with leucovorin then complete pro- tection can be achieved [47, 89]. The same results are obtained when 5-methyltetrahydrofolate is substituted for leucovorin and similar results are observed in L1210 and Walker 256 cells. In vivo administration of leucovorin alone (200mg/kg, twice a day for 2 days and starting 24 hours after PTX) allows the survival of rats given a lethal dose (100mg/kg, i.p.) of PTX. Thymidine affords no protection from lethal toxicity.

PTX has been tested for its cytotoxicity against a number of ceils both normal and MTX-resistant in tissue culture. PTX inhibits by 50% the growth of Walker 256, sarcoma 180 and L1210 leukemia cells at similar concentrations to MTX; 0.8, 1.2 and 2.2 × 10-SM respectively [47, 89]. Comparable re- suits were obtained in a study with a human lym- phoblastoid cell line, WI-L2, and show a C50 value of 2.5 x 10-8M [92]. Since a major rationale for developing a lipophilic antifolate is the circumven- tion of the more frequently observed forms of resis- tance to classical antifolates, PTX has been tested against several cell lines resistant to MTX. Sedwick et al. [92] tested PTX against mouse 3T6 cells and a mutant line (3T6R400) which overproduces an al- tered form of DHFR having a 100-fold lesser af- finity for MTX than the enzyme from parent 3T6 cells. These cells can be exposed to 1 mM MTX with little reduction in cloning efficiency [93] but they are sensitive to PTX at levels that can be achieved in vivo [92]. In a more recent study [94] PTX was tested against 3 MTX-resistant cell lines. A mutant line of CCRF-CEM ceils (CCRF-CEM/ MTX R) which shows severely impaired transport

Page 11: Biological and biochemical properties of new anticancer folate antagonists

of MTX, and is 100-fold resistant to MTX, is collat- erally sensitive to PTX. Another cell line (L5178Y/ MTXR) overproduces DHFR by 600-fold and is approximately 10,000-fold resistant to MTX. These cells were cross-resistant to PTX, and at least an order of magnitude more sensitive to PTX than MTX. The third cell line, L1210/MTXR, is a double mutant having both impaired MTX transport and a 24-fold overproduction of DHFR as the cause of its 5000-fold resistance to MTX. These cells are markedly sensitive to PTX compared to MTX, al- though not to the degree of the 'sensitive' line and the observed IC50 probably reflects a combination of resistance due to the overproduction of DHFR and collateral sensitivity due to an impaired re- duced folate carrier system. In general these stud- ies show that PTX effectively circumvents the folate carrier system and is quite capable of over- coming MTX-resistance due to its impairment. Secondly, because entry of PTX is apparently not limited by this low capacity carrier system, resis- tance due to an overproduction of DHFR is more easily overcome simply by increasing the extra- cellular concentration.

A number of studies have looked at the effect of PTX on the incorporation of precursors into the macromolecules. Exposure to PTX is associated with a very rapid suppression of deoxyuridine (dUrd) incorporation into DNA. The ability of PTX to inhibit this parameter was assessed in a variety of ceils including human lymphoblastoid cells in culture and normal leukemic myeloid cells from peripheral blood and myeloid cells from bone marrow [95]. In all cells inhibition of deoxyuridine incorporation into DNA is immediate. The degree of inhibition is similar in most leukemic cell culture lines after a 30 or 90 minute incubation, but a somewhat more variable response is found with bone marrow cells. The reversibility of suppressed dUrd incorporation was studied in the WI-L2 cells exposed to PTX as well as MTX and metoprine [92]. After a 1 hour exposure to 1/,M PTX partial recovery occurs when the extracellular drug is re- moved and the cells are resuspended in drug-free media. The initial rate of recovery is faster than in cells exposed to MTX, however the latter phase of recovery is equal to MTX and recovery from both

261

drugs equally augmented by leucovorin. The effects of PTX on the incorporation of 32p into DNA in WI-L2 cells was also assessed in this study [92] as a measure of DNA synthesis. Although this parameter is less sensitive to inhibition by antifdl- ates than dUrd incorporation it is a better indicator of the potential lethal toxicity of these drugs. A drug concentration of 1/xM PTX, MTX or meto- prine produces equivalent levels of inhibition. After a 90 minute exposure the inhibition produced by PTX and metoprine can be completely reversed simply by removing the drug and resuspending the cells in drug-free media. Recovery from the effect of MTX however requires leucovorin or thy- midine.

PTX also inhibits dUrd incorporation much more rapidly than MTX in sarcoma 180 cells [47] but the reversal patterns differ from WI-L2 cells. No difference is observed in reversal of $180 cells exposed to PTX, MTX or metoprine. When resus- pended into drug-free media with no additions the pattern of recovery from all drugs shows an initial period of increased activity followed by a sub- sequent decline of incorporation. The presence of 200/,M leucovorin does not allow cells to recover from the effects of PTX, however, it allows dUrd incorporation to resume at normal rates after inhi- bition by MTX. The ability of PTX or MTX to inhibit dUrd incorporation has been studied in bone marrow cells from patients treated with PTX [96]. The results are unusual in that dUrd incor- poration into DNA is much less susceptible to inhi- bition by MTX in cells from pretreated than from untreated patients. The response to PTX remains the same. The same results are also obtained in HL-60 cells exposed to cytostatic concentrations of PTX in tissue culture.

The transport of PTX across the plasma mem- brane in ceils was initially studied indirectly by a method which utilizes the inhibition of dUrd into DNA as a measure of this parameter [97]. Using this technique reducing the temperature of the cul- tures from 37 ° to 27 ° C does not significantly affect the kinetics of drug entry into WI-L2, indicating a lack of temperature sensitivity for the uptake pro- cess [92, 95]. More direct transport studies for PTX are reported in S180 cells using 14C labeled PTX

Page 12: Biological and biochemical properties of new anticancer folate antagonists

262

[47]. A time course for net uptake indicates that PTX associates rapidly with cells and reaches equilibrium across the membrane after 30 minutes. A very high intercept on the y axis is always present on plots of cellular drug versus incubation time even when very short time points are taken. The high intercept indicates a component of very rapid drug-cell association which may represent some mode of surface binding. Cellular drug levels achieved at the steady-state are directly and lin- early related to the extracellular concentration of drug. Washout experiments show that when cells are in equilibrium with 10/zM PTX there is a large component of exchangeable drug which effiuxes rapidly from the cell. Extended washout reveals a large nonexchangeable fraction which is most likely well above the DHFR binding capacity. Fi- nally, and similar to the indirectly determined mea- surements above, uptake of PTX appears to be temperature insensitive. No differences are ob- served between measurements made at 37 ° and 4 °C. These data are taken as evidence for a non- facilitated, passive, diffusional process as the mechanism of membrane transport for TPX.

A speculative but very appealing theory about the mechanism of antifolate-induced cytotoxicity is that uracil misincorporation into DNA results in accumulation of apyrimidinic sites. Since the nu- cleotide pool imbalance caused by antifolates is characterized by depletion of dTI'P and elevated dUTP, repair synthesis will simply cause further misincorporation, and lethal DNA damage will eventually result. Richards et al. [98] show that piritrexim causes uracil misincorporation, and that addition of deoxyuridine to the growth medium slightly increases the cytotoxicity of piritrexim to HL-60 cells. In MOLT-4 and SB cells, piritrexim causes DNA fragmentation, and in WI-L2 ceils, chromatin maturation is inhibited. It is proposed that these effects are caused by uracil misincor- poration, followed by excision of uracil to leave apyrimidinic sites [98].

Piritrexim has in vivo activity against leukemias P388 and L1210, sarcoma 180, Ehrlich ascites car- cinoma and the Walker 256 carcinosarcoma [99, 100]. Drug distribution studies in rats indicate a plasma half-life of 38 minutes following intra-

venous administration (in dogs, the half-lif~ is 130 minutes). Piritrexim extensively penetrates rat tissues, including the brain. The major pathway of metabolism is O-demethylation. The demethyl- ated metabolite is an effective inhibitor of di- hydrofolate reductase.

Nl°-Propargyl-5,8-dideazafolic acid

This quinazoline antifolate, generally referred to in the literature as CB3717, was synthesized by Jones et al. [14]. The biochemistry and pharmacology of CB3717 have been reviewed recently [101]. The original antitumor studies of this compound were carried out using the tetraploid ICR subline of L1210 leukemia [14]; in this system CB3717 is far more active than MTX, giving 90% cures against intraperitoneal tumor and 40% cures against intra- venous tumors at optimal doses. In the diploid NCI line of L1210, CB3717 is less active. CB3717 also gives 80% inhibition of tumor growth when tested against subcutaneous implants of the ADJ/PC6 plasmacytoma. CB3717 has no activity or only mar- ginal activity against TLX-5 lymphoma, P388 leu- kemia, and colon carcinoma 38 [101]. Jackman et al. caution that the interpretation of data obtained in rodents is complicated by their high circulating thymidine concentration. Thymidine not only an- tagonizes antitumor activity, but may also prevent toxicity to normal tissues. Obviously, using xeno- grafts does not help this situation, since the source of the thymidine is the host animal.

In vitro, CB3717 is only moderately cytotoxic, with IC50 values for continuous exposure failing between 1/zM and 5/zM for most cell lines [14,101- 104]. Cytotoxicity of CB3717 to murine and human cells is largely prevented by 10/xM thymidine [14, 102, 104] even when thymidine addition is delayed for many hours after drug administration [102]. In the case of MTX or trimetrexate, protection from cytotoxicity generally requires a purine, such as hypoxanthine, in addition to thymidine. Leuco- vorin does not protect cells very effectively from CB3717, and hypoxanthine not at all [14, 104].

Supporting evidence that the primary site of CB3717 action is thymidylate synthase comes from

Page 13: Biological and biochemical properties of new anticancer folate antagonists

studies with drug-resistant cells. Cells with over- production of dihydrofolate reductase have little or no cross-resistance to CB3717 [103, 104]. Con- versely, cells with overproduction of thymidylate synthase are highly resistant to CB3717 [105].

A number of kinetic studies have examined the inhibition of thymidylate synthase by CB3717 and its polyglutamate derivatives. In purified enzyme preparations of L1210 mouse leukemia cells, and WI-L2 human lymphoblastoid cells, CB3717 is competitive with 5,10-methylenetetrahydrofotate; kinetic analysis using a zone B procedure for tight- binding inhibition indicates a Ki,slop e of 4 to 5 nM [102,106]. Cheng et al., studying enzyme from KB cells, report noncompetitive inhibition with K i of 20 nM [104]. There are two likely reasons for this discrepancy. First, the analysis of Cheng et al. did not use an appropriate method for tight-binding inhibitors, and competitive tight-binding inhibitors may show apparent noncompetitive kinetics if a zone B technique is not used. (Methotrexate was originally reported to be a noncompetitive inhibi- tor of dihydrofolate reductase for that reason.) Second, Santi (cited in 50) has reported that CB3717 interacts with thymidylate synthase from Lactobacillus casei in a two-stage process. The first stage is competitive with 5,10-methylenetetra- hydrofolate. The second stage, which occurs with a half-time of 50 minutes, produces a ternary com- plex in which dUMP is covalently bound. This ternary complex dissociates with tv2 of 27 hours, If a similar interaction occurs with the human enzyme, long incubation times could result in non- steady-state conditions and apparently anomalous kinetics.

Polyglutamates of CB3717 are more potent in- hibitors of thymidylate synthase than CB3717 itself [104, 1071. Sikora et al. report that CB3717 di- and triglutamates are 30 and 84-fold more potent, re- spectively, as inhibitors of the enzyme from L1210 cells. Cheng et al. report that inhibitory potency increases with addition of successive glutamate res- idues up to the tetraglutamate, but that the pen- taglutamate is weaker. The tri- and tetraglutamates of CB3717 are thus the most potent known inhibi- tors of thymidylate synthase.

CB3717 is a moderately potent inhibitor of di-

263

hydrofolate reductase. It is competitive with dihydrofolate, and has a Ki,slop e estimated at 23-250 nM in various preparations of the enzyme. This inhibitory effect is probably not of great phar- macological importance in vivo; however, in tissue culture cells growing on folic acid-containing me- dium in presence of high concentrations of CB3717 plus thymidine, growth inhibition is seen which may be reversed by hypoxanthine or leucovorin [108]. Since the predominant physiological folate is 5-methyltetrahydrofotate rather than folic acid, this effect would probably not be seen in vivo.

In WI-L2 cells, CB3717 treatment (20/xM, 4 hours) causes marked depletion of dTTP, with minor decreases in dCTP and dTTP. dATP in- creases, as do ATP, GTP, UTP and CTP [102]. These results are consistent with the reversal of CB3717 effects by thymidine. Under normal condi- tions, CB3717 has no antipurine effect. The minor decrease in dGTP may be attributed to the fact that CB3717 causes extensive depletion of DTTP, and dTTP is a required cofactor for the GDP reductase reaction. Unlike MTX, CB3717 does not cause accumulation of dihydrofolates, dUMP accumul- ates in CB3717-treated cells, and significant amounts of dUTP are also seen [102].

The fact that CB3717 polyglutamates are effec- tive thymidylate synthase inhibitors prompted a search for polyglutamate formation in tumor cells and normal tissues. Manteuffel-Cymborowska et al. [109] report formation of di-, tri- and higher glutamates in murine liver and kidney 4 and 24 hours after the drug is administered to mice. Trig- lutamate is predominant. Sikora et al. [107] treated L1210 cells in vitro with radiolabelled CB3717. Twenty four hours after treatment, two-thirds of the extractable radioactivity is unchanged drug, but the remainder is mainly diglutamate. As yet, there is no evidence that CB3717 is selectively poly- glutamylated in malignant cells. Nevertheless, it is clear that polyglutamates are formed in both tumor and normal cells.

At present the transport mechanism for CB3717 is poorly understood. CB3717 appears to penetrate cell membranes very slowly. One study suggested that six hours after the start of incubation, at an extracellular concentration of 50/xM CB3717, the

Page 14: Biological and biochemical properties of new anticancer folate antagonists

264

cellular drug concentration does not exceed 2/xM, despite the fact that tritiated deoxyuridine incor- poration is markedly decreased within one hour [101]. The uptake of CB3717 appears to be non- saturable [110], but it is temperature-dependent [101]. Several cell lines with defective membrane transport of MTX have normal sensitivity to CB3717, suggesting that uptake of the latter cannot be through the carrier responsible for methotrex- ate transport [103].

At present no flow cytometry studies have been reported for CB3717, nor any data concerning cytotoxicity toward nonproliferating cells. On mechanistic grounds it seems likely that CB3717 will be highly S-phase selective in its effect, but experimental confirmation of this would be valu- able.

A preliminary report of CB3717 pharmacokine- tics has been presented [111]. Plasma decay in hu- mans, following a one hour intravenous infusion is biphasic, with half-lives of 70 minutes and 9 hours. Metabolites are not seen in urine or plasma, but the deglutamylated derivative is detected in the feces. The protein binding of CB3717 is more extensive than for methotrexate,

The clinical studies with CB3717 have been re- viewed by Jackman et al. [101]. Renal toxicity and hepatic toxicity have been observed. The hepatic toxicity is not seen with the related compound 5,8- dideaza folic acid, nor is it prevented by thymidine. Thus this effect is not directly attributable to inhibi- tion of thymidylate synthase; possibly it is caused by an unrelated reaction of the propargyl moiety [112]. Clinical responses to CB3717 have been re- ported in several kinds of tumors, including breast cancer patients resistant to the CMF (cyclophos- phamide + methotrexate + 5-fluorouracil) pro- tocol, demonstrating that the approach of design- ing a folate inhibitor aimed at an enzymatic target other than dihydrofolate reductase can overcome clinical MTX resistance [101].

5,10-Dideaza-5,6,7,8-tetrahydrofolic acid

The structure of this compound, which was syn- thesized by Taylor et al. [15] is shown in Fig. 1 (F).

DDATHF is highly cytotoxic to L1210 murine leu- kemia cells in culture, with ICs0 of approximately 20 nM. The cytotoxicity is not prevented by ad- dition of thymidine to the medium but it is pre- vented by hypoxanthine or by aminoimidazole carboxamide [113]. DDATHF retains activity against a methotrexate-resistant subline of L1210. DDATHF gives no significant inhibition of Lac- tobaciUus casei thymidylate synthase when tested at 1 mM concentration, and is a very weak inhibitor of beef liver dihydrofolate reductase, giving 50% inhibition at 0.6 mM. It gives excellent therapeutic activity against a panel of experimental murine tumors, many of which are insensitive to methotrexate [114]. For example, against the sub- cutaneous B16 melanoma, daily doses of DDATHF at 6.25 mg/kg for ten days give complete inhibition of tumor growth. Daily doses below 100 mg/kg have no toxic effect on the mice [115]. DDATHF depletes cellular pools of ATP and GTP, and block formation of N-formylglycinamide ribotide, indicating that its site of action is inhibi- tion of glycinamide ribotide transformylase [114, 115]. It does not inhibit the incorporation into DNA of radiolabelled deoxyuridine or thymidine [113]. DDATHF is an excellent substrate for folylpoly- glutamate synthetase from mouse liver, with K m of 4.7 tzM, and V~a x similar to that of folic acid; high concentrations of DDATHF give excess substrate inhibition [116]. DDATHF appears to be a highly novel kind of antifolate with promising experimen- tal antitumor activity. It will also be a valuable biochemical probe for the study of folate metabo- lism.

Conclusions

- 10-EdAM appears to be the most active classi- cal antifolate yet described against murine tu- mors, based on studies in five ascites systems and five subcutaneous solid tumor transplant lines. In addition, 10-EdAM shows significant activity against three human tumor xenograft lines.

- The increased therapeutic activity of 10-EdAM is achieved by optimization of the membrane

Page 15: Biological and biochemical properties of new anticancer folate antagonists

transport differential between tumor cells and normal tissues, and of the degree of poly- glutamylation of drug in tumor cells relative to normal cells.

- Nonclassical antifolates, such as trimetrexate and piritrexim have activity against murine tu- mors unresponsive to MTX. Tumors with ac- quired resistance to MTX because of DHFR overproduction are totally or partially cross- resistant to lipophilic DHFR inhibitors. How- ever. tumors with resistance to MTX, caused by deletion of membrane transport, retain sen- sitivity to trimetrexate and piritrexim, and may even be collaterally sensitive.

- Resistance to trimetrexate and piritrexim may arise by a mutation that decreases the rate of cellular uptake of these drugs. In addition, tu- mors with pleiotropic drug resistance to an- thracyclines and vinca alkaloids may be cross- resistant to trimetrexate and piritrexim.

- The cytotoxicity of trimetrexate is highly phase- dependent on the cell cycle and occurs only in S-phase.

- Trimetrexate depletes cellular pools of dTTP and of purine nucleotides. Its cytotoxicity can be completely prevented by thymidine plus hy- poxanthine.

- The cytotoxicity of piritrexim in vitro cannot be blocked by thymidine plus hypoxanthine. FuIi protection of S-180 cells requires a mixture of thymidine, hypoxanthine, adenosine and uridine.

- CB3717 has very limited activity against murine tumors, but has shown significant activity against some human tumors. This difference is attributed to the fact that thymidine levels in human plasma are much lower than in mouse plasma.

- CB3717 is active against MTX-resistant tumors with D H F R overproduction and impaired MTX uptake. However, routine and human cells have been described with resistance to CB3717 attributable to overproduction of thy- midylate synthase.

- Like inhibitors of dihydrofolate reductase, CB3717 depletes cellular dTTP pools, but un- like dihydrofolate reductase inhibitors, CB3717

265

appears to have no antipurine effect. - CB3717 is a substrate for folylpolyglutamate

synthetase, and polyglutamates of CB3717 are detectable in L1210 cells treated in vitro, and in normal tissues of CB3717-treated mice.

- D D A T H F is a potent inhibitor of G A R trans- formylase. It does not inhibit dihydrofolate re- ductase or thymidylate synthase. DDATHF is a substrate for folylpolyglutamate synthase.

- DDATHF has broad-spectrum activity against a panel of murine tumors.

K e y u n a n s w e r e d q u e s t i o n s

- Why does MTX have such narrow spectrum activity against murine tumors? If this is be- cause of limited transport, does the four-fold lower influx K m of 10-EdAM partly explain its greater activity in murine tumors7 Does inade- quate transport limit the clinical spectrum of MTX, and if so, will 10-EdAM be more active than MTX against human tumors?

- What are the effects of 10-EdAM on cellular nucleotide pools?

- Since nonclassical antifolates are not trans- ported by the 5-methyltetrahydrofolate mem- brane carrier, and are not polyglutamylated, to what is their antitumor selectivity attributable?

- is the rapid influx of trimetrexate and pi- ritrexim into tumor cells passive or carrier-me- diated? If it is passive, what is the nature of the mutation that confers resistance to these agents by impaired uptake?

- What is the mechanism by which cells with pleiotropic drug resistance become cross-resis- tant to trimetrexate and piritrexim? Does this resistance mechanism, which appears to be present intrinsically in some colon tumors, ex- plain the low response rate of colon tumors to these agents?

- What is the cellular component to which non- exchangeable trimetrexate (in excess of the cel- lular dihydrofolate reductase level) is bound? Does this bound drug provide a cellular reser- voir of pharmacologically active drug?

- Why can t'-,e cytotoxicity of piritrexim not be

Page 16: Biological and biochemical properties of new anticancer folate antagonists

266

reversed by thymidine plus hypoxanthine? Does piritrexim have cellular sites of action other than dihydrofolate reductase?

- What is the mechanism by which CB3717 enters cells?

- What is the mechanism of the renal and hepatic toxicities of CB3717?

- Is the cytotoxic mechanism of DDATHF medi- ated through inhibition of synthesis of D N A or RNA or both?

- What is the cell cycle phase-dependence of the cytotoxicity of DDATHF? Is D D A T H F cytotoxic to non-cycling cells?

References

1. Roth B, Cheng CC: Recent progress in the medicinal chemistry of 2,4-diaminopyrimidines. Prog Med Chem 19: 269-331, 1982

2. Montgomery JA, Piper JR: Design and synthesis of folate analogs as antimetabolites. In: Sirotnak FM, Burchall JJ, Ensminger WD, Montgomery JA (eds) Folate antagonists as therapeutic agents, 1. Academic Press, Orlando, 1984, pp 219-260

3. Werbel LM: Design and synthesis of lipophilic antifols as anticancer agents. In: Sirotnak FM, Burchall JJ, Ensminger WD, Montgomery JA (eds) Folate antagonists as therapeutic agents, 1. Academic Press, Orlando, 1984, pp 261-287

4. Hill BT, Price LA: DDMP (2,4-diamino-5-(3',4'-di- chlorophenyl)-6-methylpyrimidine). Cancer Treat Rev 7: 95-112, 1980

5. Hutchison DJ, Schmid FA: Experimental cancer chemo- therapy with folate antagonists. In: Sirotnak FM, Burchall J J, Ensminger WD, Montgomery JA (eds) Folate antago- nists as therapeutic agents, 2. Academic Press, Orlando, 1984, pp 1-22

6. Sirotnak FM, DeGraw JI: Selective antitumor action of folate analogs. In: Sirotnak FM, Burchall JJ, Ensminger WD, Montgomery JA (eds) Folate antagonists as therapeutic agents, 2. Academic Press, Orlando, 1984, pp 43-95

7. Freisheim JH, Matthews DA: The comparative bio- chemistry of dihydrofolate reductase. In: Sirotnak FM, Burchall J J, Ensminger WD, Montgomery JA (eds) Folate antagonists as therapeutic agents, 1. Academic Press, Orlando, 1984, pp 69--131

8. Taylor IW, Tattersall MHN: Methotrexate cytotoxicity in cultured human leukemic cells studied by flow cytometry, Cancer Res 41: 1549-1558, 1981

9. Jackson RC, Grindey GB: The biochemical basis for methotrexate cytotoxicity. In: Sirotnak FM, Burchall JJ,

Ensminger WD, Montgomery JA (eds) Folate antagonists as therapeutic agents, 1. Academic Press, Orlando, 1984, pp 289-315

10. DeGraw JI, Brown VH, Tagawa H, Kisliuk RL, Gaumont Y, Sirotnak FM: Synthesis and antitumor activity of 10-al- kyl-10-deazaaminopterin. A convenient synthesis of 10-deazaaminopterin. J Med Chem 25: 1227-1230, 1982

11. Elslager EF, Davoll J: Synthesis of fused pyrimidines as folate antagonists. In: Castle RN and Townsend LR (eds) Lectures in heterocyclic chemistry, vol. 2. Hetero Corp, Orem Utah, 1974, pp $97-S133

12. Elslager EF, Johnson JL, Werbel LM: Folate antagonists 20. Synthesis, antitumor, and antimalarial properties of trimetrexate and related 6-((phenylamino)-methyl)-2,4- quinazoline-diamines. J Med Chem 26: 1753-1760, 1983

13. Grivsky EM, Lee S, Sigel CW, Duch DS, Nichol CA: Synthesis and antitumor activity of 2,4-diamino-6-(2,5- dimethoxybenzyl)-5-methylpyrido [2,3-d] pyrimidine. J Med Chem 23: 327-329, 1980

14. Jones TR, Calvert AH, Jackman AL, Brown S J, Jones M, Harrap KR: A potent antitumour quinazoline inhibitor of thymidylate synthetase: synthesis, biological properties and therapeutic results in mice. Eur J Cancer 17: 11-19, 1981

15. Taylor EC, Wong G, Fletcher SR, Harrington PJ, Shih CJ, Beardsley GP: Synthesis of 5,10-dideaza-5,6,7,8- tetrahydrofolic acid (DDATHF) and analogs. In: Cooper BA, Whitehead VM (eds) Pteridines and folic acid deriva- tives. De Gruyter, Berlin, 1986, pp 61-68

16. DeGraw JI: Synthesis and antifolate activity of 10-de- azaaminopterin. J Med Chem 17: 552-552, 1974

17. Sirotnak FM, DeGraw JI, Moccio DM, Dorick DM: Anti- tumor properties of a new folate analog, 10-deaza-aminop- terin, in mice. Cancer Treat Rep 62: 1047-1052, 1978

18. Sirotnak FM, DeGraw JI, Chello PL, Moccio DM, Dorick DM: Biochemical and pharmacologic properties of a new folate analog, 10-deaza-aminopterin, in mice. Cancer Treat Rep 66: 351-358, 1982

19. Currie VE, Warrell RP, Arlin Z, Tan C, Sirotnak FM, Greene G, Young CW: Phase I trial of 10-deaza-aminop- terin in patients with advanced cancer. Cancer Treat Rep 67: 149-154, 1983

20. Sirotnak FM: Correlates of folate analog transport, phar- macokinetics and selective antitumor action. Pharmac Ther 8: 71-103, 1980

21. Goldman ID, Matherly LH: The cellular pharmacology of methotrexate. Pharmac Ther 28: 77-102, 1985

22. Sirotnak FM, DeGraw JI, Moccio DM, Samuels LL, Goutas LJ: New folate analogs of the 10-deaza-aminop- terin series. Basis for structural design and biochemical and pharmacologic properties. Cancer Chemother Phar- macol 12: 18-25, 1984

23. Margolis S, Philips FS, Sternberg SS: The cytotoxicity of methotrexate in mouse small intestine in relation to inhibi- tion of folic acid reductase and DNA synthesis. Cancer Res 31: 2037-2046, 1971

Page 17: Biological and biochemical properties of new anticancer folate antagonists

24. Sirotnak FM, DeGraw JI, Schmid FA, Goutas L J, Moccio DM: New folate analogs of the 10-deaza-aminopterin se- ries. Further evidence for markedly increased antitumor efficacy compared with methotrexate in ascitic and solid routine tumor models. Cancer Chemother Pharmacol 12: 26-30, 1984

25. Moccio DM, Sirotnak FM, Samuels LL, Ahmed T, Yagoda A, DeGraw JI, Piper JR: Similar specificity of membrane transport for folate analogues and their meta- bolites by murine and human tumor cells: a clinically directed laboratory study. Cancer Res 44: 352-357, 1984

26. Chabner BA, Allegra CJ, Curt GA, Clendeninn NJ, Baram J, Koizumi S, Drake JC, Jolivet J: Polyglutamation of methotrexate. Is methotrexate a prodrug? J Clin Invest 76: 907-912, 1985

27. Goldman ID: Proceedings of the second workshop on folyl and antifolyl polyglutamates. Praeger Publishers, Westport, Connecticut, 1985

28. Samuels LL, Moccio DM, Sirotnak FM: Similar differen- tial for total polyglutamylation and cytotoxicity among various folate analogues in human and routine tumor ceils in vitro. Cancer Res 45: 1488--1495, 1985

29. Kinahan J J, Samuels LL, Farag F, Fanucchi MP, Vidal PM, Sirotnak FM, Young CW: Fluorometric high-per- formance liquid chromatographic analysis of 10-de- azaaminopterin, 10-ethyl-10-deazaaminopterin, and known metabolites. Analyt Biochem 150: 203-213, 1985

30. Kinahan J, Samuels L, Farag F, Sirotnak F, Young C: Fluorometric HPLC analysis of 10-deaza-aminopterin (10- dAM), 10-ethyl-10-deaza-aminopterin (10-EdAM), and polyglutamated derivatives in tumor and biological sam- ples. Proc Am Assoc Cancer Res 25: 310, 1984

31. Ueda T, Dutschman GE, Nair MG, DeGraw JI, Sirotnak FM, Cheng YC: Inhibitory action of the 10-deazaaminop- terins and their polyglutamates on human thymidylate synthetase. Proc Amer Assoc Cancer Res 27: 258, 1986

32. Kisliuk RL, Gaumont Y, Kumar P, Nair MG: Inhibition of thymidylate synthase by antifolylpolyglutamates. In: Cooper BA, Whitehead VM (eds) Pteridines and folic acid derivatives. DeGruyter, Berlin, 1986, pp 989-992

33. Mandelbaum-Shavit F: Effect of antifolates 10-methyl- and 10-ethyl-10-deazaaminopterin on a human breast can- cer cell line. In: Cooper BA, Whitehead VM (eds) Pteridines and folic acid derivatives. DeGruyter, Berlin, 1986, in press

34. Sirotnak FM: Biochemical pharmacologic and antitumor properties of 10-ethyl-10-deazaaminopterin. Cancer Treat Symposia, 1986, pp 855-859

35. Schmid FA, Sirotnak FM, Otter GM, DeGraw JI: New folate analogs of the 10-deaza-aminopterin series: markedly increased antitumor activity of the 10-ethyl ana- log compared to the parent compound and methotrexate against some human tumor xenografts in nude mice. Can- cer Treat Rep 69" 551-553, 1985

36. Kinahan J, Fanucchi M, Vidal P, Chou TC, Sternberg S, Farag F, Niedzwiecki D, Samuels L, Sirotnak F, Young C:

267

Preclinical toxicology and pharmacology of 10-ethyl- 10-deaza-aminopterin (10-EdAM). Proc Am Assoc Can- cer Res 26: 356, 1985

37. Ohnoshi T, Ohnuma T, Holland JF: Establishment of methotrexate (MTX)-resistant human acute lymphocytic leukemia cells in culture and effects of various antifols. Proc Am Assoc Cancer Res 21: 298, 1980

38. Ohnuma T, Lo RJ, Scanlon KJ, Kamen BA, Ohnoshi T, Wolman SR, Holland JF: Evolution of methotrexate resis- tance of human acute lymphoblastic leukemia cells in vitro. Cancer Res 45: 1815-1822, 1985

39. Diddens H, Niethammer D, Jackson RC: Patterns of cross-resistance to the antifolate drugs trimetrexate, metoprine, homofolate, and CB3717 in human lymphoma and osteosarcoma cells resistant to methotrexate. Cancer Res 43: 5286-5292, 1983

40. Mini E, Moroson BA, Franco CT, Bertino JR: Cytotoxic effects of folate antagonists against methotrexate-resistant human leukemic lymphoblast CCRF-CEM cell lines. Can- cer Res 45: 325-330, 1985

41. Frei E, Rosowsky A, Wright JE, Cucchi CA, Lippke JA, Ervin TJ, Jolivet J, Haseltine WA: Development of meth- otrexate resistance in a human squamous cell carcinoma of the head and neck in culture. Proc Natl Acad Sci USA 81: 2873-2877, 1984

42, Kamen BA, Eibl B, Cashmore A, Bertino J: Uptake and efficacy of trimetrexate (TMO, 2,4-diamino-5-methyl- 6-[(3,4,5-trimethoxyanilino)methyl] quinazoline), a non- classical antifolate in methotrexate-resistant leukemia cells in vitro. Biochem Pharmacol 33: 1697-1699, 1984

43. Kamen BA, Eibl B, Cashmore AR, Whyte WL, Moroson BA, Bertino JR: Efficacy and transport of a new lipid soluble antifol, 2,4-diamino-5-methyl-6-[(3,4,5-tri- methoxyanilino)methyl] quinazoline (TMQ); JB-11) in methotrexate resistant cells. Proc Am Assoc Cancer Res 22: 26, 1981

44. Jackson RC, Fry DW, Boritzki TJ, Besserer JA, Leopold WR, Sloan B J, Elslager EF: Biochemical pharmacology of the lipophilic antifolate, trimetrexate. Adv Enz Reg 22: 187-206, 1984

45. Besserer JA, Fry DW, Boritzki TJ, Jackson RC: Studies on cellular resistance and membrane transport with the quinazoline antifolate, trimetrexate. Fed Proc 43: 91, 1984

46. Fry DW, Besserer JA, Boritzki TJ, Jackson RC, Elslager EF: Impaired cellular uptake as a mechanism of resistance to lipophilic antifolates. Proc Am Assoc Cancer Res 26: 340, 1985

47. Duch DS, Edelstein MP, Bowers SW, Nichol CA: Bio- chemical and chemotherapeutic studies on 2,4-diamino- 6-(2,5-dimethoxybenzyl)-5-methylpyrido[2,3-d]pyrimid- ine (BW301U), a novel lipid-soluble inhibitor of di- hydrofolate reductase. Cancer Res 42: 3987-3994, 1982

48. Hill BT, Price LA, Goldie JH: Methotrexate resistance and uptake of DDMP by L5178Y cells. Europ J Cancer 11: 545-553, 1975

49. Hill BT, Price LA, Harrison SI, Goldie JH: Studies on the

Page 18: Biological and biochemical properties of new anticancer folate antagonists

268

transport and distribution of diaminopyrimidines in L5178Y lymphoblasts in cell culture. Biochem Pharmacol 24: 535-538, 1975

50. Greco WR, Hakala MT: Cellular pharmacokinetics of lipophilic diaminopyrimidine antifolates. J Pharmacol Ex- per Ther 212: 39-46, 1980

51. Arkin H, Ohnuma T, Takemura Y, Kamen BA, Kano Y, Holland JF: Development and characterization of a tri- metrexate (TMTX, TMQ, JB-ll)-resistant human acute lymphoblastic leukemia cell line. Proc Am Assoc Cancer Res 26: 25, 1985

52. Mini E, Sobrero A, Moroson BA, Bertino JR: Differen- tial efficacy of trimetrexate (2,4-diamino-6-methyl-(3,4,5- trimethoxyanilino)methyl) quinazoline, JB-11, TMQ) and methotrexate (MTX) on human colon carcinoma and human leukemia cells in vitro. Proc Am Assoc Cancer Res 26: 230, 1985

53. Borman LS, McCermack J J: Studies in vitro of the effects of 2,4- diamino- 5- methyl-6- [ (3,4,5-trimethoxyanilino) methyl]-quinazoline (TMQ; trimetrexate) on rat colon carcinoma and L1210 cells. Proc Am Assoc Cancer Res 25: 311, 1984

54. Klohs WD, Steinkampf RW, Besserer JA, Fry DW: Cross resistance of pleiotropically drug resistant P388 leukemia cells to the lipophilic antifolates trimetrexate and BW301U. Cancer Lett, 31: 253-260, 1986

55. Ramu N, Ramu A, Pollard HB, Balis F, Poplack DG: Resistance to trimetrexate (TMQ) is related to pleiotropic drug resistance. Proc Am Assoc Cancer Res 27: 391, 1986

56. Arkin H, Ohnuma T, Holland JF, Greenspan EM: Multi- drug (pleiotropic) resistance in trimetrexate (TMQ)-resis- tant human acute lymphomablastic leukemia (ALL) cell line, MOLT-3. Proc Am Assoc Cancer Res 27: 269, 1986

57. Sobrero AF, Bertino JR: Alternating trimetrexate (TMQ) with methotrexate (MTX) delays the onset of resistance to antifolates in vitro and in vivo. Proc Am Assoc Cancer Res 27: 269, 1986

58. Bertino JR, Sawicki WL: Potent inhibitory activity of trimethoxyquine (TMQ), 'nonclassical' 2+4, diamino- quinazoline, on mammalian DNA synthesis. Proc Am Assoc Cancer Res 18: 168, 1977

59. Bertino JR, Sawicki WL, Moroson BA, Cashmere AR, Elslager EF: 2,4-diamino-5-methyl-6-[3,4,5-trimethox- yanilino) quinazoline (TMQ), a potent non-classical relate antagonist inhibitor-I. Biochem Pharmaco128: 1983-1987, 1979

60. Heusner JJ, McCormack J J: Enzymatic assays for 2,4- diamino-5-methyl-6-[(3,4,5-trimethoxyanilino) methyl] quinazoline, a promising new 'nonclassical' antifolate. J Pharm Sci 70: 827-828, 1981

61. Broome MG, Johnson RK, Wodinsky I: Biochemical and biological characterization of two new antifols, NSC 127755 and TMQ, in comparison with methotrexate. Proc Am Assoc Cancer Res 21: 309, 1980

62. Broome MG, Johnson RK, Evans SF, Wodinsky I: Leuco- vorin reversal studies with TMQ and a triazine antifol in

comparison with methotrexate. Proc Am Assoc Cancer Res 23: 178, 1982

63. Bertino JR, Mini E, Sawicki WL, Moroson BA, Jas- treboff M, McGuire J J: Effects of trimetrexate on human leukemia cells from patients sensitive and resistant to methotrexate (MTX). Proc Am Soc Clin Oncol 4: 44, 1985

64. Browman GP, Spiegl P, Booker L, Rosowsky A: Com- parison of leucovorin protection from variety of antifol- ates in human lymphoid cell lines. Cancer Chemother Pharmacol 15: 111-114, 1985

65. McCormack JJ, Heusner J J, Hacker MP, Mathews LA: Further pharmacological studies of 2,4-diamino-methyl- 6- [ (3,4,5-trimethoxyanilino) methyl] quinazoline (NSC 249008; TMQ). Proc Am Assoc Cancer Res 22: 245, 1981

66. Hook KE, Nelson JM, Roberts BJ, Griswold DP, Leopold WR: Cell cycle effects of trimetrexate (CI-898). Cancer Chemother Pharmacol 16: 116-120, 1986

67. Lathan B, Von HoffD, Elslager E, Rodriguez V: Activity of trimetrexate (TMQ) a non-classical antifolate in a human tumor cloning system. Proc Am Assoc Cancer Res 24: 278, 1983

68. McCormack J J, Heusner JJ, Mathews LA: Preclinical studies with 2,4-diamino-5-methyl-6-[3,4,5-trimethoxy- anilino)methyl]quinazoline (NSC 249008) A 'non-classi- cal' antifolate. Proc Am Assoc Cancer Res 21: 294, 1980

69. O'Dwyer PJ, Shoemaker DD, Plowman J, Cradock J, Grillo-Lopez A, Leyland-Jones B: Trimetrexate: a new antifol entering clinical trials. Invest New Drugs 3: 71-75, 1985

70. Leopold WR, Dykes DD, Griswold Jr DP: Therapeutic synergy of trimetrexate (CI-898) in combination with dox- orubicin, vincristine, cytoxan, 6-thioguanine, cisplatin, or 5-fluorouracil against intraperitoneally implanted P388 leukemia. Cancer Treatment Symposium, in press, 1986

71. Leopold WR, Dykes DD, Griswold Jr DP: Therapeutic synergy of trimetrexate (CI-898) in combination with dox- orubicin, cytoxan, 6-thioguanine, 5-fluorouracil, vin- cristine, or cisplatin against P388 leukemia. Proc Am As- soc Cancer Res 27: 253, 1986

72. Weir EC, Cashmere AR, Dreyer RN, Graham ML, Hsiao N, Moroson BA, Sawicki WL, Bertino JR: Pharmacology and toxicity of a potent +nonclassical' 2,4-diamino quinazoline relate antagonist, trimetrexate, in normal dogs. Cancer Res 42: 1696-1702, 1982

73. Gans JH, Tong WP, Whitfield LR, Mathews LA, McCor- mack J J: Pharmacokinetic studies with trimetrexate in dogs. Proc Am Assoc Cancer Res 27: 405, 1986

74. Schornagel JH, Weir EC, Harris RL, Graham ML, Cash- more AR, Bertino JR: Trimetrexate in normal and lym- phoma-bearing dogs: pharmacology, toxicity and anti- tumor effects. Invest New Drugs 2: 115, 1984

75. Balis FM, Lester CM, Poplack DG: Pharmacokinetics of methotrexate (NSC 352122) in monkeys. Cancer Res 46: 169-174, 1986

76. Fanucchi M, Fleisher M, Vidal P, Williams L, Bauer T, Cassidy C, Chou T-C, Young C: Phase I and phar-

Page 19: Biological and biochemical properties of new anticancer folate antagonists

macologic study of trimetrexate (TMTX). Proc Am Assoc Cancer Res 26: 179, 1985

77. Rosen M, Ohnuma T, Zimet A, Coffey V, Zhang N, Holland JF: Phase I study of trimetrexate (TMTX, TMQ, JB-11) glucuronate in a 5-day infusion schedule. Proc Am Assoc Cancer Res 27: 172, 1986

78. Donehower RC, Graham ML, Thompson GE, Dole GB, Ettinger DS: Phase I and pharmacokinetie study of tri- metrexate (TMTX) in patients with advanced cancer. Proc Am Soc Clin Oncol 4: 32, 1985

79. Legha S, Tenney D, Ho DH, Krakoff I: Phase I clinical & pharmacology study of trimetrexate (TMQ). Proc Am Soc Clin Oncol 4: 48, 1985

80. Ho DHW, Covington WP, Legha S, Newman RA, Kra- koff IH: Clinical pharmacology of trimetrexate (TMX). Proc Am Assoc Cancer Res 27: 173, 1986

81. Heusner JJ, Tong WP, McCormack JJ, Mathews LA: Disposition of the non-classical folate antagonist 2,4- diamino-5-methyl-6- (3,4,5-trimethoxyanilino) methyl quinazoline (TMQ; trimetrexate) in mice. Proc Am Assoc Cancer Res 26: 234, 1985

82. Webster LK, Tong WP, Mathews LA, Stewart JA: Stud- ies of a metabolite of trimetrexate. Proc Am Assoc Cancer Res 27: 256, 1986

83. Webster LK, Tong WP, McCormack JJ: Elimination of trimetrexate (TMTX) by the isolated, perfused rat liver. Proc Am Assoc Cancer Res 26: 120, 1985

84. Webster LK, Tong WP, McCormack JJ: Effect of hypoxia on metabolism of trimetrexate in isolated perfused rat livers. Proc Am Assoc Cancer Res 27: 405, 1986

85. Heusner J J, Franklin MR: Inhibition of metabolism of the 'nonclassical' antifolate, trimetrexate(2,4-diamino- 5-methyl-6- [ (3,4,5-trimethoxyanilino) methyl] quinazoline) by drugs containing an imidazole moiety. Pharmacol 30: 266--272, 1985

86. Hurlbert BS, Ledig KW, Valenti BF, Hitchings GH: Stud- ies on condensed pyrimidine systems. XXIII. Synthesis of 2,4-diaminopyrido[2,3-d]pyrimidines from beta-keto esters. J Med Chem 11: 703-707, 1968

87. Hurlbert BS, Valenti BF: Studies on condensed pyrimidine systems. XXIV. The condensation of 2,4,6- triaminopyrimidine with malondialdehyde derivatives. J Med Chem 11: 708-710, 1968

88. Hurlbert BS, Ferone R, Herrmann TA, Hitchings GH: Studies on condensed pyrimidine systems. XXV. 2,4-di- aminopyrido-[2,3-d]pyrimidines. Biological data. J Med Chem 1l: 711-717, 1968

89. Duch DS, Sigel CW, Bowers SW, Edelstein MP, Cavallito JC, Foss RG, Nichol CA: Lipid-soluble inhibitors of di- hydrofolate reductase: selection and evaluation of the 2,4- diaminopyridopyrimidine BW 301U and related com- pounds as anticancer agents. In: Nelson JD, Grassi C (eds) Current chemotherapy and infectious disease. The Amer- ican Society for Microbiology, Washington DC, 1980, pp 1597-1599

90. Duch DS, Edelstein MP, Nichol CA: Inhibition of his-

269

tamine metabolizing enzymes and elevation of histamine (HA) levels in tissues by anticancer folate antagonists. Pharmacologist 21: 266, 1979

91. Duch DS, Edelstein MP, Nichol CA: Inhibition of his- tamine-metabolizing enzymes and elevation of histamine levels in tissues by lipid-soluble anticancer folate antago- nists. Mol Pharmacol 18: 100-104, 1980

92. Sedwiek WD, Hamrell M, Brown OE, Laszlo J: Meta- bolic inhibition by a new antifolate, 2,4-diamino-6-(2,5- dimethoxybenzyl)-5-methyl-pyrido[2,3-d]pyrimidine (BW301U), an effective inhibitor of human lymphoid and dihydrofolate reductase-overproducing mouse cell lines. Mol Pharmacol 22: 766-770, 1982

93. Hamrell M, Laszlo J, Brown OE, Sedwick WD: Toxicity of methotrexate and metoprine in a dihydrofolate reduc- tase gene-amplified mouse cell line. Mol Pharmacol 20: 637-643, 1981

94. Taylor IW, Slowiaczek P, Friedlander ML, Tattersall MHN: Selective toxicity of a new lipophilic antifolate, BW301U, for methotrexate-resistant cells with reduced drug uptake. Cancer Res 45: 978--982, 1985

95. Sedwick WD, Birdwell R, Laszlo J: Lipid-soluble pyridopyrimidine, BW301U, with metabolic inhibitory ca- pacity equal to methotrexate in human lymphoblastoid, normal myeloid, human leukemia, and dihydrofolate re- ductase-overproducing mouse cells. In: Nelson JD, Grassi C (eds) Current chemotherapy and infectious disease. The American Society for Microbiology, Washington DC, 1980, pp 1593-1595

96. Iland HJ, Laszlo J. Sedwick WD: Paradoxical effect of BW 301U, a lipophilic antifolate, on methotrexate-inhib- itable deoxyuridine incorporation by human hematopoie- tic cells, Cancer Res 45: 3962-3968, 1985

97. Sedwick WD, Fyfe M J, Brown OE, Frazer TA, Kutler M, Laszlo J: Deoxyuridine incorporation as a useful measure of methotrexate and metoprine uptake and metabolic effectiveness. Mol Pharmacol 16: 607-613, 1979

98. Richards RG, Sowers LC, Laszlo J, Sedwick WD: The occurrence and consequences of deoxyuridine in DNA. Adv in Enzyme Regul 22: 157-185, 1984

99. Sigel CW, Macklin AW, Woolley JL, Blum MR, Clen- denin NJ, Everitt BS, Foss RG, Duch DS, Nichol CA: Preclinical biochemical pharmacology and toxicology of piritrexim, a lipophilic inhibitor of dihydrofolate reduc- tase. Cancer Treatment Symposia, in press, 1986

100. Laszlo J, Iland H J, Sedwick WD: Overcoming methotrex- ate resistance by a lipophilic antifolate (BW301U): from theory to models to practice. Advances in Enzyme Regul 24: 357-375, 1986

101. Jackman AL, Jones TR, Calvert AH: Thymidylate syn- thetase inhibitors: experimental and clinical aspects. In: Muggia FM (ed) Experimental and clinical progress in cancer chemotherapy. Martinus Nijhoff, Boston, 1985

102. Jackson RC, Jackman AL, Calvert AH: Biochemical effects of a quinazoline inhibitor of thymidylate syn- thetase, CB3717, on human lymphoblastoid cells. Bio-

Page 20: Biological and biochemical properties of new anticancer folate antagonists

270

ehem Pharmacoi 32: 3783-3790, 1983 103. Diddens H, Niethammer D, Jackson RC: Human cells

resistant to methotrexate: sensitivity to the nonclassical antifolates trimetrexate, metoprine, homofolic acid and CB3717. In: Blair JA (ed) Chemistry and Biology of Pteridines. DeGruyter, Berlin, 1983, pp 953-957

104. Cheng YC, Dutschman GE, Starnes MC, Fisher MH, Nanavathi NT, Nair MG: Activity of the new antifolate N10-propargyl-5,8-dideazafolate and its polyglutamates against human dihydrofolate reductase, human thymidyl- ate synthetase, and KB cells containing different levels of dihydrofolate reductase. Cancer Res 45: 598-600, 1985

105. Jackman AL, Alison DL, Calvert AH, Harrap KR: In- creased thymidylate synthase in L1210 cells possessing ac- quired resistance to N10-propargyl-5,8-dideazafolic acid (CB3717): development, characterization, and cross-resis- tance studies. Cancer Res 46: 2810-2815, 1986

106. Jackman AL, Calvert AH, Hart LI, Harrap KR: Inhibi- tion of thymidylate synthetase by the new quinazoline antifolate CB3717: enzyme purification and kinetics. In: DeBruyn CHM, Simmonds HA, Muller MM (eds) Purine metabolism in man, IV part B. Biochemical, immunologi- cal and cancer research. Plenum Publishing Corp, New York, 1984, pp 375-378

107. Sikora E, Pawelczak K, Rzeszotarska B, Harrap KR, Calvert AH, Jones TR, Jackman AL, Newell DR: N10- propargyl-5,8-dideazafolic acid: intracellular binding and polyglutamylation. In: Cooper BA, Whitehead VM (eds) Pteridines and folic acid derivatives. DeGruyter, Berlin, 1986, pp 675-679

108. Jackman AL, Moran RG, Calvert AH: The reversal of the cytotoxicity of folate-based thymidylate synthase inhibi- tors in cultured L1210 cells. In: Cooper BA, Whitehead VM (eds) Pteridines and folic acid derivatives. DeGruy- ter, Berlin, 1986, pp 644-649

109. Manteuffel-Cymborowska, Kaminska B, Grzelakowska- Sztabert B: Polyglutamylation of the thymidylate syn- thetase inhibitor, N10-propargyl-5,8-dideazafolic acid (CB3717) in organs of Ehrlich ascites carcinoma-bearing mice. In: Cooper BA, Whitehead VM (eds) Pteridines and folic acid derivatives. DeGruyter, Berlin, 1986, pp 993--996

110. Jackman AL, Grzelakowska-Sztabert B, Newell DR, Cal- vert AH, Harrap KR: Transport studies with H CB3717. Brit J Cancer 52: 431, 1985

111. Alison DL, Newell DR, Calvert AH: Pharmacokinetic studies in humans with CB3717. Brit J Cancer 48:126,1983

112. Newell DR, Alison DL, Jackman AL, Sessa C. Calvert AH, Harrap KR: Clinical and preclinical pharmacokinetic studies with the thymidylate synthetase inhibitor N10-pro- pargyl-5~8-dideazafolic acid (CB3717). Proc Amer Assoc Cancer Res 26: 350, 1985

113. Moran RG, Taylor EC, Beardsley GP: 5,10-dideaza- 5,6,7,8-tetrahydrofolic acid, a potent antifolate inhibitory to de n o v o purine biosynthesis. Proc Amer Assoc Cancer Res 26: 231, 1985

114. Beardsley GP, Taylor EC, Grindey GB, Moran RG: De- aza derivatives of tetrahydrofolic acid. A new class of folate antimetabolite. In: Cooper BA, Whitehead VM (eds) Pteridines and folic acid derivatives. De Gruyter, Berlin, 1986, pp 953-957

115. Beardsley GP, Taylor EC, Shih C, Poore GA, Grindey GB, Moran RG: A new class of antifolates. 5,10-di- deazatetrahydrofolic acid (DDATHF), an inhibitor of GAR transformylase with broad in vivo activity. Proc Amer Assoc Cancer Res 27: 259, 1986

116. Moran RG: Structural requirements for the activity of folate analogs as substrates for folylpolyglutamate syn- thetase. Cancer Treat Symp, in press, 1986