Cytidine Analogues



Cytidine Analogues


Bruce A. Chabner

Jacob Glass



Nucleoside analogs have earned an important place in the treatment of acute leukemia. Through modification of the base or sugar component, chemists have been able to create molecules that mimic the physiological counterparts but potently inhibit aspects of DNA synthesis or function. Among these analogues are the arabinose nucleosides, a unique class of antimetabolites isolated from the sponge Cryptothethya crypta1 but now produced synthetically.2 They differ from the physiologic deoxyribonucleosides in the presence of a 2′-OH group in the cis configuration relative to the N-glycosyl bond between cytosine and the arabinose sugar (Fig. 10-1). Several arabinose nucleosides, including cytosine arabinoside (ara-C, cytarabine), 2-fluoro-ara-adenosine monophosphate, and nelarabine, a guanine analogue, have been important agents for treating hematological cancers.


Cytosine Arabinoside

Ara-C is one of the most effective agents in the treatment of acute myelogenous leukemia (AML)3 and is incorporated into virtually all standard induction regimens for this disease, generally in combination with an anthracycline (daunorubicin hydrochloride or idarubicin hydrochloride). Ara-C is also a component of consolidation and maintenance regimens in AML after remission is attained.4 High-dose ara-C confers particular benefit in AML patients with certain cytogenetic abnormalities related to the core binding factor that regulates hematopoiesis (t8:21, inv 16, del 16, t16:16).5 Ara-C is also active against other hematologic malignancies, including Burkitt’s lymphoma,6 acute lymphocytic leukemia,7 and chronic myelogenous leukemia8 but has little value as a single agent against solid tumors. This limited spectrum of activity has been attributed to the lack of metabolic activation of this agent in solid tumors and its selective action against rapidly dividing cells. The essential features of ara-C pharmacology are described in Table 10-1.


Mechanism of Action

Ara-C acts as an analogue of deoxycytidine (CdR) and has multiple effects on DNA synthesis. Ara-C undergoes phosphorylation to form arabinosylcytosine triphosphate (ara-CTP), which competitively inhibits DNA polymerase α in opposition to the normal substrate deoxycytidine 5′-triphosphate (dCTP).9 This competitive inhibition has been demonstrated with crude DNA polymerase from calf thymus9 and with purified enzyme from human leukemic cells10 and from murine tumors.11 Ara-CTP has an affinity for human leukemia cell DNA polymerase α in the range of 1 × 10−6 mol/L and competes with dCTP. In intact cells, its effects are antagonized by the addition of CdR, the precursor of dCTP.12 Ara-CTP inhibits DNA polymerase β with lesser potency.13 The effects of ara-C on DNA polymerase activity extend not only to semiconservative DNA replication but also to DNA repair. Repair of ultraviolet light damage to DNA, a function of polymerase α, is blocked more potently than the repair of photon-induced or γ radiation-induced strand breaks,14 which are repaired by a different polymerase. In addition to its inhibition of DNA synthesis, however, it becomes incorporated into DNA, a feature that correlates closely with cytotoxicity15,16 (Fig. 10-2). In fact, a preponderance of evidence suggests that this is the major cytotoxic lesion in ara-C-treated cells. Drugs that prevent ara-C incorporation into DNA, such as aphidicolin, block its cytotoxicity.17 In cell culture experiments, a linear relationship exists between picomoles of ara-C incorporated into DNA and the log of cell survival for a wide range of drug concentrations and durations of exposure. Thus, drug toxicity is a direct function of incorporation into DNA, and the latter varies directly with drug concentration and duration of exposure.18 Once incorporated into DNA, ara-C is excised slowly,19 and the incorporated ara-C inhibits template function and chain elongation.17,20,21 In experiments with purified enzyme and calf thymus DNA, the consecutive incorporation of two ara-C or two arabinosyl-5-azacytidine (ara-5-aza-C) residues effectively stops chain elongation.10 At high concentrations of ara-C, one finds a greater than expected proportion of ara-C residues at the 3′-terminus, a finding that implies potent chain termination.19 These observations support the hypothesis that ara-C incorporation into DNA is a prerequisite for drug action and is responsible for cytotoxicity.

Ara-C also causes an unusual reiteration of DNA segments.22 Human lymphocytes exposed to ara-C in culture synthesize small reduplicated segments of DNA, which results in multiple copies of limited portions of DNA. These reduplicated segments increase the possibility of recombination, crossover, and gene amplification; gaps and breaks are observed in karyotype preparations after ara-C treatment. The same mechanism, reiteration of DNA synthesis after its inhibition by an antimetabolite, may explain the high frequency of gene reduplication induced by methotrexate, 5-fluorouracil, and hydroxyurea (see relevant chapters). In summary, although ara-C has multiple effects on DNA synthesis, the most important antitumor mechanism seems to be its incorporation into DNA.

Other biochemical actions of ara-C have been described, including a relatively weak inhibition of ribonucleotide reductase23 and formation of arabinosylcytosine diphosphate (ara-CDP)-choline. The latter functions as an analog of cytidine 5′-diphosphocholine (CDP-choline) and inhibits synthesis of membrane glycoproteins and glycolipids.24 Ara-C also has the interesting property of promoting differentiation of leukemic cells in tissue culture, an effect that is accompanied by decreased c-myc oncogene expression.25 These changes in morphology and oncogene expression occur at concentrations above the threshold for cytotoxicity and may simply represent terminal injury of cells. After exposure to ara-C, both normal and malignant cells undergo apoptosis in experimental models.26







FIGURE 10-1 Structure of cytidine analogs.








TABLE 10.1 Key features of ara-C pharmacology





























Factor


Result


Mechanism of action


Inhibits DNA polymerase α, is incorporated into DNA, and terminates DNA chain elongation


Metabolism


Activated to triphosphate in tumor cells. Degraded to inactive ara-U by deamination
Converted to ara-CDP choline derivative


Pharmacokinetics


Plasma: t½α 7-20 min, t1/2β 2 h; CSF: t1/2 2 h


Elimination


Deamination in liver, plasma, and peripheral tissues—100%


Drug interactions


Methotrexate increases ara-CTP formation
THU, 3-deazauridine inhibit deamination
Fludarabine phosphate increases ara-CTP formation
Ara-C blocks DNA repair, enhances activity of alkylating agents


Toxicity


Myelosuppression
Gastrointestinal epithelial ulceration
Intrahepatic cholestasis, pancreatitis
Cerebellar and cerebral dysfunction (high dose)
Conjunctivitis (high dose)
Hidradenitis
Noncardiogenic pulmonary edema


Precautions


High incidence of cerebral-cerebellar toxicity with high-dose ara-C in the elderly, especially in those with compromised renal function


ara-CDP, arabinosylcytosine diphosphate; ara-CTP, arabinosylcytosine triphosphate; ara-U, uracil arabinoside; CSF, cerebrospinal fluid; t1/2, half-life; THU, tetrahydrouridine; ara-C, cytosine arabinoside.








FIGURE 10-2 Relationship between AML blast clonogenic survival and incorporation of tritium-labeled ara-C into DNA at ara-C concentrations of 10−7 mol/L (▲), 10−6 mol/L (•), 10−5 mol/L (▪), and 10−4 mol/L (○) during periods of 1, 3, 6, 12, and 24 hours. (From Kufe DW, Spriggs DR. Biochemical and cellular pharmacology of cytosine arabinoside. Semin Oncol 1985;12:34.)

The mechanism by which ara-C induces apoptosis is uncertain. It may be triggered by p53 in response to DNA breaks (see above) and by futile attempts to excise incorporated ara-C nucleotides. Ara-C stimulates the formation of ceramide, a potent inducer of apoptosis.27 Ara-C induces production of diacylglycerol, which activates protein kinase C (PKC), a response that opposes apoptosis in hematopoietic cells. The lethal actions of ara-C may depend, at least partially, on its relative effects on the PKC and ceramide pathways.

Cells exposed to ara-C display marked alterations in transcription factors. Ara-C induces AP-1 (a dimer of jun-fos or jun-jun proteins) and NF-kB, and induction has been temporally associated with apoptosis.28,29 PKC inhibitors promote ara-C-induced apoptosis despite their antagonizing c-jun up-regulation, calling into question the involvement of c-jun expression in apoptosis secondary to ara-C.30 Ara-C induces pRb phosphatase activity, another possible mechanism responsible for p53-independent G1 arrest and apoptosis.31 The resulting hypophosphorylated pRb binds to and inactivates the E2F transcription factor, which inhibits the transcription of genes responsible for cell-cycle progression.32


Cellular Pharmacology and Metabolism

Ara-C penetrates cells by a carrier-mediated process shared by physiologic nucleosides.33 Several different classes of transporters for nucleosides have been identified in mammalian cells; the most extensively characterized in human tumors is hENT1, the equilibrative transporter, identified by its binding to nitrobenzylthioinosine (NBMPR). The number of transport sites on the cell membrane is greater in AML cells than in acute lymphocytic leukemia cells and can be enumerated by incubation of cells with NBMPR.34 The hENT1 transporter is highly upregulated in biphenotypic leukemia associated with the 11q23 MLL gene (4:11) translocation.35 Uptake occurs rapidly. A steady-state level of intracellular drug is achieved within 90 seconds at 37°C.36,37 Studies of Wiley et al.33 and others34,37 suggest that the NBMPR transporter plays a limiting role in the action of this agent in that the formation of the ultimate toxic metabolite ara-CTP is strongly correlated with the number of transporter sites on leukemic cells (Fig. 10-3). A single point
mutation in the hENT1 carrier confers resistance in leukemic cell lines in vitro.36 At drug concentrations above 10 μmol/L, the transport process becomes saturated, and further entry takes place by passive diffusion.37 hENT1 is strongly inhibited by various receptor tyrosine kinase inhibitors, an interaction that could limit ara-C use with targeted drugs.38






FIGURE 10-3 Correlation between accumulation of ara-CTP and nucleoside transport capacity measured by the maximal number of NBMPR binding sites on leukemic cells (r = 0.87; P < 0.0001). Ara-CTP accumulation was measured after incubation of cells with 1 μmol/L of tritium-labeled ara-C for 60 minutes. ●, acute myelogenous leukemia; ○, non-T-cell acute lymphoblastic leukemia; ▲, T-cell leukemia/lymphoma, lymphoblastic leukemia; ▪, acute undifferentiated leukemia; γ, chronic lymphocytic leukemia. (From Wiley JS, Taupin J, Jamieson GP, et al. Cytosine arabinoside transport and metabolism in acute leukemias and T-cell lymphoblastic lymphoma. J Clin Invest 1985;75:632.)






FIGURE 10-4 Metabolism of ara-C by tumor cells. The conversion of ara- UMP to a triphosphate has not been demonstrated in mammalian cells. Ara-CMP, arabinosylcytosine monophosphate; ara-CDP, arabinosylcytosine diphosphate; ara- CTP, arabinosylcytosine triphosphate; ara-U, uracil arabinoside; dCMP, deoxycytidine monophosphate; NDP, nucleoside diphosphate.

As shown in Figure 10-4, ara-C must be converted to its active form, ara-CTP, through the sequential action of three enzymes: (a) CdR kinase, (b) deoxycytidine monophosphate (dCMP) kinase, and (c) nucleoside diphosphate (NDP) kinase. Ara-C is subject to degradation by cytidine deaminase, forming the inactive product uracil arabinoside (ara-U); arabinosylcytosine monophosphate (ara-CMP) is likewise degraded by a second enzyme, dCMP deaminase, to the inactive arabinosyluracil monophosphate (ara-UMP). Each of these enzymes, with the exception of NDP kinase, has been examined in detail because of its possible relevance to ara-C resistance.








TABLE 10.2 Kinetic parameters of enzymes that metabolize ara-C



















































Enzyme


Substrate


Km (mol/L)


Activity in AML cells (nmol/h/mg protein at 37°C)


CdR kinase


Ara-C


2.6 × 10−5


15.4 ± 16



CdR


7.8 × 10−6


dCMP kinase


Ara-CMP


6.8 × 10−4


1,990 ± 1,500



dCMP


1.9 × 10−3


dCDP kinase


Ara-CDP


?


Not known



Other NDPs


?


CR deaminase


Ara-C


8.8 × 10−5


372 ± 614



CdR


1.1 × 10−5


dCMP deaminase


Ara-CMP dCMP


Ara-CMP has higher Km than dCMP; exact Km not determined


1,250 (five patients)


AML, acute myelogenous leukemia; Ara-C, cytosine arabinoside; ara-CDP, arabinosylcytosine diphosphate; ara-CMP, arabinosylcytosine monophosphate; CdR, deoxycytidine; CR, cytidine; dCDP, deoxycytidine diphosphate; dCMP, deoxycytosine monophosphate; NDPs, nucleoside diphosphates.


The first activating enzyme, CdR kinase, is found in lowest concentration (Table 10-2) and is believed to be rate limiting in the process of ara-CTP formation. The enzyme is a 30.5-kDa protein that phosphorylates ara-C and many pyrimidine and purine nucleosides and their analogs. The gene coding for CdR kinase and a mutated version found in ara-C resistant cells have been cloned.39,40 The rate-limiting role of CdR kinase in ara-C activation is illustrated by transfection of malignant cell lines with retroviral vectors containing CdR kinase cDNA, a maneuver that substantially increases the susceptibility of cells to ara-C, gemcitabine, and purine analogues.41 Ara-C cytotoxicity increases when intracerebral gliomas in rats are transduced with CdR kinase.42

CdR kinase activity is highest during the S phase of the cell cycle.43 The Km, or affinity constant, for ara-C is 20 μmol/L, compared with the higher affinity or 7.8 μmol/L for the physiologic substrate CdR.44 This enzyme is strongly inhibited by dCTP but weakly inhibited by ara-CTP. This lack of “feedback” inhibition allows accumulation of the ara-C nucleotide to higher concentrations. PKC-α, the activity of which is increased after ara-C exposure, has been implicated in phosphorylation and activation of CdR kinase. This observation raises the possibility that ara-C at high doses may potentiate its own metabolism by induction of the PKC activator, diacylglycerol.45

The second activating enzyme, dCMP kinase,46 is found in several hundred-fold higher concentration than CdR kinase. Its affinity for ara-CMP is low (Km = 680 μmol/L) but greater than the affinity for the competitive physiologic substrate dCMP. Because of its relatively poor affinity for ara-CMP, this enzyme could become rate limiting at low ara-CMP concentrations. The third activating enzyme, the diphosphate kinase, appears not to be rate limiting because it is present in very high concentration and the intracellular pool of ara-CDP is only a fraction of the ara-CTP pool.47


Opposing the activation pathway are two deaminases found in high concentration in some tumor cells as well as in normal tissues. Cytidine deaminase is widely distributed in mammalian tissues, including intestinal mucosa, liver, and granulocytes.48 It is found in granulocyte precursors and in leukemic myeloblasts in lower concentrations than in mature granulocytes, but even in these immature cells, the deaminase level exceeds the activity of CdR kinase, the initial activating enzyme.44,48 The second degradative enzyme, dCMP deaminase (Fig. 10-4), regulates the flow of physiologic nucleotides from the dCMP pool into the deoxyuridine monophosphate (dUMP) pool from which point dUMP is ultimately converted to deoxythymidine 5′-phosphate (dTMP) by thymidylate synthase.49 The enzyme dCMP deaminase is strongly activated by intracellular dCTP (Km = 0.2 μmol/L) and strongly inhibited by deoxythymidine triphosphate in concentrations of 0.2 μmol/L or greater. Ara-CTP weakly activates this enzyme (Km = 40 μmol/L50 and, thus, would not promote degradation of its own precursor nucleotide, ara-CMP. The affinity of dCMP deaminase for ara-CMP is somewhat higher than the affinity of dCMP kinase for the same substrate, but the activity of these competitive enzymes depends greatly on their degree of activation or inhibition by regulatory triphosphates (dCTP), and dCMP deaminase concentration in leukemic myeloblasts is slightly less than that of dCMP kinase (Table 10-2).

The balance between activating and degrading enzymes, thus, is crucial in determining the quantity of drug converted to the active intermediate, ara-CTP. This enzymatic balance varies greatly among cell types.44 CdR kinase activity is higher and cytidine deaminase activity lower in lymphoid leukemia than in acute myeloblastic leukemia. Enzyme activities vary also with cell maturity; deaminase increases dramatically with maturation of granulocyte precursors, whereas kinase activity decreases correspondingly.48 Thus, admixture of normal granulocyte precursors with leukemic cells in human bone marrow samples complicates the interpretation of enzyme measurements unless normal and leukemic cells are separated. In general, cytidine deaminase (D) activity greatly exceeds kinase (K). The kinase:deaminase ratio averages 0.03 in human AML, whereas the enzyme activities are approximately equal in acute lymphoblastic leukemia and Burkitt’s lymphoma. Thus, the biochemical setting seems to favor drug activation by lymphoblastic leukemia cells if these initial enzymes play a rate-limiting role.

In fact, this may not be the case. Chou et al.47 found that human AML cells formed 12.8 ng of ara-CTP per 106 cells after 45 minutes of incubation with 1 × 10−5 mol/L ara-C. Acute lymphoblastic leukemia cells formed less ara-CTP, 6.3 ng/106 cells, and as expected, the more mature chronic myelocytic and chronic lymphocytic leukemia cells formed lesser amounts of ara-CTP (4.7 to 5.2 ng/106 cells). The likelihood is that other factors, such as transport across the cell membrane and regulatory effects of intracellular nucleoside triphosphate concentrations, may limit ara-CTP formation.

In addition to its activation to ara-CTP, ara-C is converted intracellularly to ara-CDP-choline,51 an analog of the physiologic CDP-choline lipid precursor. However, ara-C does not inhibit incorporation of choline into phospholipids of normal or transformed hamster embryo fibroblasts.24 Ara-CMP does inhibit the transfer of galactose, N-acetylglucosamine, and sialic acid to cell surface glycoproteins. Further, high concentrations (approaching 1 mM) of ara-CTP inhibit the synthesis of cytidine monophosphate (CMP)-acetylneuraminic acid, an essential substrate in sialylation of glycoproteins.52 Thus, ara-C treatment could alter membrane structure, antigenicity, and function.


Biochemical Determinants of Cytosine Arabinoside Resistance

The foregoing consideration of ara-C metabolism and transport makes it clear that a number of factors could affect ara-C response. Not surprisingly, many of these factors have been implicated in various preclinical models of ara-C resistance. The most frequent abnormality found in resistant leukemic cells recovered from mice treated with ara-C has been decreased activity of CdR kinase.53 In cultured cells exposed to a mutagen and then to low concentrations of ara-C, some single-step mutants developed high-level resistance to ara-C through loss of activity of CdR kinase, whereas other resistant clones exhibited markedly expanded dCTP pools, presumably through increased cytidine-5′-triphosphate (CTP) synthetase activity or through deficiency of dCMP deaminase.54,55 As mentioned previously, specific mutations and deletions in the gene coding for CdR kinase derived from resistant cells have been described.39,40

The role of cytidine deaminase in experimental models of resistance is less clear. Retrovirus-mediated transfer of the cytidine deaminase cDNA into 3T3 murine fibroblast cells significantly increases drug resistance to ara-C and other analogs such as 5-aza- 2′-CdR and gemcitabine. This phenotype of increased cytidine deaminase activity and drug resistance is reversed by the cytidine deaminase inhibitor tetrahydrouridine (THU).56 Other genes, including proto-oncogenes, may affect ara-C response. Transfection of rodent fibroblasts and human mammary HBL 100 cells with c-H-ras conferred resistance to ara-C, an event attributed to decreased activity of CdR kinase.57 On the other hand, N-ras or K-ras mutations strongly correlated with increased ara-C sensitivity in the screening of human tumor cell lines from the National Cancer Institute’s in vitro drug screen.58 Ras mutations are found in 20% of AML cases, and these patients appear to derive greatest benefit from high-dose ara-C regimens.59 Although various molecular lesions have been implicated as causing ara-C resistance in animals, their relevance to resistance in human leukemia is less certain. Clinical studies have described specific biochemical changes in drug-resistant cells from patients with leukemia, including deletion of CdR kinase,60 increased cytidine deaminase,61 a decreased number of nucleoside transport sites,62 and increased dCTP pools.63 Other clinical investigators have not been able to correlate resistance with either CdR kinase or cytidine deaminase or their ratio.64,65 All studies have shown extreme variability in enzyme levels among patients with leukemia. Thus, no agreement exists as to the specific changes responsible for resistance in human leukemia.

Although specific biochemical lesions associated with resistance in humans are unclear, the current understanding of ara-C action suggests that the ultimate formation of ara-CTP and the duration of its persistence in leukemic cells determine response.47,66 Chou et al.47 found greater ara-CTP formation in leukemic cells of responders when these cells were incubated in vitro with ara-C, but even this correlation was not confirmed in other studies.67, 68, 69

Preisler et al.70 found that the duration of remission induced by ara-C-containing regimens was strongly correlated with the ability of cells to retain ara-CTP in vitro after removal of ara-C from
the medium. Attempts to monitor ara-CTP formation in leukemic cells taken from patients during therapy have not disclosed useful correlations of ara-CTP levels or intracellular persistence with response.69,71 Ara-CTP has an intracellular half-life of about 3 to 4 hours. Again, considerable variability has been observed in the rates of formation of ara-CTP, and this rate does not correlate well with plasma ara-C pharmacokinetics in individual patients (Fig. 10-5).

The cellular response to ara-C-mediated DNA damage also governs whether the genotoxic insult results in cell death. Ara-C incorporation into DNA stalls the replication fork for cells in active DNA synthesis, activating ATR and Chk 1, checkpoint kinases that block cell cycle progression and allow for removal of ara-C from DNA. Absence of either of these checkpoints sensitizes cells to apoptosis. Levels of expression of apoptotic proteins influence response. Overexpression of the antiapoptotic proteins Bcl-2 and Bcl-XL in leukemic blasts causes in vitro resistance to ara-C-mediated apoptosis.72 The intracellular metabolism of ara-C and its initial effects on DNA are not modified by Bcl-2 expression, which suggests that Bcl-2 primarily regulates the more distal steps in the ara-C-induced cell death pathway. Although the precise mechanism by which these proteins prevent ara-C-induced cytotoxicity remains to be elucidated, Bcl-2 and Bcl-XL have been shown to antagonize ara-C-mediated cell death by caspase activation.72 The fact that antisense oligonucleotides directed against Bcl-2 increase the susceptibility of leukemic blasts to ara-C-induced apoptosis in vitro,73 and that patients whose blasts express high levels of Bcl-2 respond poorly to ara-C-containing regimens,74 further suggests the pivotal role of Bcl-2 in ara-C resistance.






FIGURE 10-5 Pharmacokinetics of ara-CTP in leukemia cells and of ara-C in plasma. Blood samples were drawn at the indicated times during and after infusion of ara-C, 3 g/m2, to patients with acute leukemia in relapse. Symbols for each analysis are the same for individual patients. (From Plunkett W, Liliemark JO, Estey E, et al. Saturation of ara-CTP accumulation during high-dose ara-C therapy: pharmacologic rationale for intermediate-dose ara-C. Semin Oncol 1987;14[2 (Suppl 1)]:159.)

Phosphorylation of apoptotic or DNA damage response factors may also determine the outcome of ara-C exposure. Phosphorylation of Bcl-2 is required for its antiapoptotic function, and a functional role for PKC-α in Bcl-2 phosphorylation and suppression of apoptosis has been postulated,75 although this observation has not been confirmed by others.76 Attempts at pharmacological inhibition of PKC-α activity have increased ara-CTP formation but had variable effects on cytotoxicity in culture, possibly due to their simultaneous activation of BcL-2.76 Altered phosphorylation of transcription factors also influences the cellular response to ara-C toxic insult. Ara-C-induced activation of PKC and mitogen-activated protein kinase (MAPK) increases c-jun expression and phosphorylation,27,77 and hyperphosphorylation of the AP-1 transcription factor has been associated with ara-C resistance in human myeloid leukemic cell lines in vitro.78

Clinical studies of determinants of ara-C response are complicated by the fact that ara-C is almost always given in combination with an anthracycline or an anthraquinone. Thus, a complete response or long remission duration does not necessarily imply sensitivity to ara-C. A lack of response does imply resistance to both agents in the combination, except for the not-infrequent cases in which failure can be attributed to infection or inability to administer full dosages of drug. With these limitations, the duration of complete response is probably the most appropriate and most important single yardstick of drug sensitivity because it reflects the fractional cell kill during induction therapy, but no single factor has emerged as a determinant of remission duration.


Cell Kinetics and Cytosine Arabinoside Cytotoxicity

In addition to biochemical factors that determine response, cell kinetic properties exert an important influence on the results of ara-C treatment. As an inhibitor of DNA synthesis, ara-C has its greatest cytotoxic effects during the S phase of the cell cycle perhaps because of the requirement for its incorporation into DNA and the greater activity of anabolic enzymes during S phase. The duration of exposure of cells to ara-C is directly correlated with cell kill because the longer exposure period allows ara-C to be incorporated into the DNA of a greater percentage of cells as they pass through S phase. The cytotoxic action of ara-C is not only cell-cycle phasedependent but is influenced by the rate of DNA synthesis. That is, cell kill in tissue culture is greatest if cells are exposed during periods of maximal rates of DNA synthesis, as in the recovery period after exposure to a cytotoxic agent. In experimental situations, it has been possible to schedule sequential doses of ara-C to coincide with the peak in recovery of DNA synthesis and thus to improve the therapeutic results.79

In humans, the influence of tumor cell kinetics on response is unclear. Although, earlier studies showed that the complete remission rate seems to be higher in patients who have a high percentage of cells in S phase,80 remissions are longer in patients with leukemias that have long cell-cycle time.81



Clinical Pharmacology—Assay Methods

The preferred method for assay of ara-C and its primary metabolite ara-U is high-pressure liquid chromatography, which has the requisite specificity and adequate (0.1 μmol/L) sensitivity.82 An alternative method using gas chromatography-mass spectrometry combines high specificity with greater sensitivity (4 nmol/L) but requires derivatization of samples and thus prolonged performance time.83 Because of the presence of cytidine deaminase in plasma, the deaminase inhibitor THU must be added to plasma samples immediately after blood samples are obtained.


Pharmacokinetics

The important factors that determine ara-C pharmacokinetics are its high aqueous solubility and its susceptibility to deamination in liver, plasma, granulocytes, and gastrointestinal tract. Ara-C is amenable to use by multiple schedules and routes of administration and has shown clinical activity in dosages ranging from 3 mg/m2 twice weekly to 3 g/m2 every 12 hours for 6 days. Remarkably, over this wide dosage range, its pharmacokinetics remains quite constant and predictable.


Distribution

As a nucleoside, ara-C is transported across cell membranes by a nucleoside transporter and distributes rapidly into total-body water.84 It then crosses into the central nervous system (CNS) with surprising facility for a water-soluble compound and reaches steady-state levels at 20% to 40% of those found simultaneously in plasma during constant intravenous infusion. At conventional doses of ara-C (100 mg/m2 by 24-hour infusion), spinal fluid levels reach 0.2 μmol/L, which is probably above the cytotoxic threshold for leukemic cells. High doses of ara-C yield proportionately higher ara-C levels in the spinal fluid.85


Plasma Pharmacokinetics

The pharmacokinetics of ara-C are characterized by rapid disappearance from plasma owing to deamination, with some variability seen among individual patients.83 Peak plasma concentrations reach 10 μmol/L after bolus doses of 100 mg/m2 and are proportionately higher (up to 150 μmol/L) for doses up to 3 g/m2 given over a 1- or 2-hour infusion86 (Fig. 10-6). Thereafter, the plasma concentration of ara-C declines, with a half-life of 7 to 20 minutes. A second phase of drug disappearance has been detected after high-dose ara-C infusion, with a terminal half-life of 30 to 150 minutes, but the drug concentration during this second phase has cytotoxic potential only in patients treated with high-dose ara-C.87,88 Seventy to eighty percent of a given dose is excreted in the urine as ara-U,87 which, within minutes of drug injection, becomes the predominant compound found in plasma. Ara-U has a longer half-life in plasma (3.2 to 5.8 hours) than does ara-C and may enhance the activation of ara-C through feedback inhibition of ara-C deamination in leukemic cells.87 The steady-state level of ara-C in plasma achieved by constant intravenous infusion remains proportional to dose for dose rates up to 2 g/m2/d. At this dosage, steady-state plasma levels approximate 5 μmol/L. Above this rate of infusion, the deamination reaction is saturated and ara-C plasma levels rise unpredictably, which leads to severe toxicity in some patients.88 To accelerate the achievement of a steady-state concentration, one may give a bolus dose of three times the hourly infusion rate before infusion.89 Equivalent drug exposure (area under the curve [AUC]) is achieved by subcutaneous or intravenous infusion of ara-C, although one study has reported higher ara-CTP concentrations in leukemia cells after subcutaneous administration.90






FIGURE 10-6 Ara-C pharmacokinetics in plasma after doses of 3 g/m2 given over 2 hours, 100 mg/m2/h by continuous infusion for 24 hours, 4 mg/m2/h (a conventional antileukemic dose) by continuous intravenous infusion, and 10 mg/m2 subcutaneously or intravenously as a bolus.

Owing to the presence of high concentrations of cytidine deaminase in the gastrointestinal mucosa and liver, orally administered ara-C provides much lower plasma levels than does direct intravenous administration. Threefold to tenfold higher doses must be given in animals to achieve a biologic effect equivalent to that produced by intravenous drug. The oral route, therefore, is not routinely used in humans.

Ara-C may also be administered by intraperitoneal infusion for treatment of ovarian cancer.91 After instillation of 100 μmol/L of drug, ara-C levels fall in the peritoneal cavity with a half-life of approximately 2 hours. Simultaneous plasma levels are 100- to 1,000-fold lower, presumably because of deamination of ara-C in liver before it reaches the systemic circulation. In 21-day continuous infusion, patients tolerated up to 100 μmol/L intraperitoneal concentrations but developed peritonitis at higher concentrations.92


Cerebrospinal Fluid Pharmacokinetics

After intravenous administration of 100 mg/m2 of ara-C, parent drug levels reach 0.1 to 0.3 μmol/L in the cerebrospinal fluid (CSF). Thereafter, levels decline with a half-life of 2 hours. Proportionately higher CSF levels are reached by intravenous high-dose ara-C regimens; for example, a 3 g/m2 infusion intravenously over 1 hour yields peak CSF concentrations of 4 μmol/L,86 whereas the same dose over 24 hours yields peak CSF ara-C concentrations of 1 μmol/L.88

Ara-C is effective when administered intrathecally for the treatment of metastatic neoplasms. A number of dosing schedules for giving intrathecal ara-C have been recommended, but twice weekly or weekly schedules of administration are the most often used. The dose of ara-C ranges from 30 to 50 mg/m2. The dose is generally
adjusted in pediatric patients according to age (15 mg for children below 1 year of age, 20 mg for children between 1 and 2 years, 30 mg for children between 2 and 3 years, and 40 mg for children older than 3 years). The clinical pharmacology of ara-C in the CSF following intrathecal administration differs considerably from that seen in the plasma following a parenteral dose. Systematically administered ara-C is rapidly eliminated by biotransformation to the inactive metabolite ara-U. In contrast, little conversion of ara-C to ara-U takes place in the CSF following an intrathecal injection. The ratio of ara-U to ara-C is only 0.08, a finding that is consistent with the very low levels of cytidine deaminase present in the brain and CSF. Following an intraventricular administration of 30 mg of ara-C, peak levels exceed 1 to 2 mM, and levels decline slowly, with the terminal half-life of approximately 3.4 hours.93 Concentrations above the threshold for cytotoxicity (0.1 μg/mL, or 0.4 μmol/L) are maintained in the CSF for 24 to 48 hours. The CSF clearance is 0.42 mL/min, which is similar to the CSF bulk flow rate. This finding suggests that drug elimination occurs primarily by this route. Plasma levels following intrathecal administration of 30 mg/m2 of ara-C are less than 1 μmol/L, which illustrates again the advantage of intracavitary therapy with a drug that is rapidly cleared in the systemic circulation.

Depocytarabine (DTC 101) is a depot formulation in which ara-C is encapsulated in microscopic Gelfoam particles (DepoFoam) for sustained release into the CSF so that the need for repeated lumbar punctures is avoided. The encapsulation of ara-C in DepoFoam results in a 55-fold increase in CSF half-life after intraventricular administration in rats, from 2.7 to 148 hours. Cytotoxic concentrations of free ara-C (>0.4 μmol/L) in CSF are maintained for more than 1 month following a single intrathecal dose administration of 2 mg of DTC 101 in rhesus monkeys. A phase I trial of DTC 101 given intraventricularly has been performed in patients with leptomeningeal metastasis. Free ara-C CSF concentration decreased biexponentially. After a dose of 50 mg of DTC, ara-C concentrations were maintained above the cytotoxic threshold for 12 ± 3 days. The maximum tolerated dosage was 75 mg administered every 3 weeks, and the dose-limiting toxicity was headache and arachnoiditis.94 A randomized study involving patients with lymphomatous meningitis demonstrated a possible prolongation of time to neurologic progression in patients treated with 50 mg of DTC 101 every 2 weeks compared with patients treated with standard intrathecal ara-C.95 DTC appears to give equivalent results to standard intrathecal methotrexate, given every 4 days, for treatment of carcinomatous meningitis.96


Alternate Schedules of Administration

Although ara-C is used most commonly in regimens of 100 to 200 mg/m2/d for 7 days, other high-dose and low-dose schedules have been used in treating leukemia. The more effective of these newer regimens have been high-dose schemes, usually 2 to 3 g/m2 every 12 hours for six doses.97 High-dose ara-C is used primarily in the consolidation phase for acute myelocytic leukemia.4 The rationale for the higher-dose regimen initially rested on the assumption that ara-C phosphorylation is the rate-limiting intracellular step in the drug’s activation and could be promoted by raising intracellular concentrations to the Km of CdR kinase for ara-C, or approximately 20 μmol/L. Above this level, further increases in ara-C do not lead to increased ara-CTP because the phosphorylation pathways enzymes become saturated.98

Others have examined the clinical activity of low-dose ara-C, particularly in older patients with myelodysplastic syndromes.99 These regimens have used dosages in the range of 3 to 20 mg/m2/d for up to 3 weeks, with the expectation that low doses would produce less toxicity and promote leukemic cell differentiation (or apoptosis). The persistence of chromosomal markers for the leukemic cell line in remission granulocytes has been documented, findings that support induction of differentiation.100 In general, although the low-dose regimens produce less toxicity, myelosuppression often supervenes and less than 20% of patients achieve meaningful improvement in blood counts.


Toxicity

The primary determinants of ara-C toxicity are drug concentration and duration of exposure. Because ara-C is cell cycle phase specific, the duration of cell exposure to the drug is critical in determining the fraction of cells killed.101 In humans, single-bolus doses of ara-C as large as 4.2 g/m2 are well tolerated because of the rapid inactivation of the parent compound and the brief period of exposure, whereas constant infusion of drug for 48 hours using total doses of 1 g/m2 produces severe myelosuppression.102

Myelosuppression and gastrointestinal epithelial injury are the primary toxic side effects of ara-C. With the conventional 5- to 7-day courses of treatment, the period of maximal toxicity begins during the first week of treatment and lasts 14 to 21 days. The primary targets of ara-C are platelet production and granulopoiesis, although anemia also occurs. Little acute effect is seen on the lymphocyte count, although a depression of cell-mediated immunity is found in patients receiving ara-C.103 Megaloblastic changes consistent with suppression of DNA synthesis are observed in both the white and red cell precursors.104

Only gold members can continue reading. Log In or Register to continue

Stay updated, free articles. Join our Telegram channel

May 27, 2016 | Posted by in ONCOLOGY | Comments Off on Cytidine Analogues

Full access? Get Clinical Tree

Get Clinical Tree app for offline access