Skip to main content

New insights into the synergism of nucleoside analogs with radiotherapy

Abstract

Nucleoside analogs have been frequently used in combination with radiotherapy in the clinical setting, as it has long been understood that inhibition of DNA repair pathways is an important means by which many nucleoside analogs synergize. Recent advances in our understanding of the structure and function of deoxycytidine kinase (dCK), a critical enzyme required for the anti-tumor activity for many nucleoside analogs, have clarified the mechanistic role this kinase plays in chemo- and radio-sensitization. A heretofore unrecognized role of dCK in the DNA damage response and cell cycle machinery has helped explain the synergistic effect of these agents with radiotherapy. Since most currently employed nucleoside analogs are primarily activated by dCK, these findings lend fresh impetus to efforts focused on profiling and modulating dCK expression and activity in tumors. In this review we will briefly review the pharmacology and biochemistry of the major nucleoside analogs in clinical use that are activated by dCK. This will be followed by discussions of recent advances in our understanding of dCK activation via post-translational modifications in response to radiation and current strategies aimed at enhancing this activity in cancer cells.

Overview of purine and pyrimidine nucleoside analogs that synergize with ionizing radiation

Nucleoside analogs comprise a class of rationally designed agents that emerged in the 1950s from insight gained by advances made in understanding DNA structure and DNA synthesis. In many ways, the underlying logic behind the creation of these compounds presaged the development of more recent targeted therapies by modeling cancer drugs after endogenous nucleotides in an effort to corrupt key cellular processes. By acting in this manner, it has become possible to kill rapidly dividing cancer cells by exploiting differences in the rate and amount of DNA synthesis between normal cells and cancer cells. The nucleoside analogs can be divided into sub-classes based on their structural similarity to purine bases (adenine and guanine) or pyrimidine bases (cytosine, uracil, or thymine). In general, these agents exert their cytotoxic actions through common means such as disruption of DNA function, inhibition of DNA replication, or a combination thereof. Additionally, these drugs as a class share the need to be transported into the cell through nucleoside transporters and are metabolically activated following internalization into the cell. A number of these drugs have been shown to work synergistically with radiation, a feature that is exploited clinically to enhance tumor regression, and the subject of this review. We will briefly review the pharmacology for each of these compounds followed by putative mechanisms by which radiosensitization or chemosensitization may be achieved. We will conclude with a discussion of recent efforts to identify patient suitability for combination chemotherapy and ways to enhance these synergistic effects in susceptible individuals in an optimal fashion.

Purine based analogs

Fludarabine

Originally synthesized by John Montgomery and Kathleen Hewson in 1969, the fluorinated arabinosyl nucleoside analog fludarabine is a prodrug that incorporates a number of structural features that extend its half-life by protecting it from degradative enzymes [1]. Some of these structural features are shared by other nucleoside analogs. For example, the presence of a fluorine atom on the 2- position of the adenine ring makes fludarabine resistant to ADA mediated metabolism. The presence of a hydroxyl group (β instead of α) at the 2-position of the sugar ring is also a common structural modification that helps reduce glycosidic bond cleavage by the bacterial purine nucleoside phosphorylase (PNP), although fludarabine is still susceptible to phosphorolysis by human PNP [2]. Degradation by bacterial PNP can limit oral bioavailability whereas human PNP is involved in normal cellular metabolism of purine nucleosides to bases [3]. Unlike other nucleoside analogs, fludarabine is administered in the monophosphate form to increase solubility and bioavailability [4, 5]. However, fludarabine-monophosphate is rapidly dephosphorylated by plasma localized 5′-nucleotidases in a rapid and complete fashion prior to cellular uptake by the hENT1, hENT2, hCNT2, or hCNT3 nucleoside transporters [5]. Once internalized into the target cell, fludarabine is phosphorylated by deoxycytidine kinase (dCK) to a monophosphate form facilitating its retention inside the cell. Phosphorylation of fludarabine to the diphosphate and triphosphate forms appear to be catalyzed by adenylate kinase and nucleoside diphosphate kinase, respectively [6]. The affinity of dCK for fludarabine appears to be relatively weak as evidenced by low Km values (100-600 μmol/L) and this may be explained by the aforementioned structural modifications [69]. Nevertheless, the interaction of dCK with fludarabine is specific and occurs rapidly when dCK is abundant [6]. As a consequence T-lymphoblasts, which contain high levels of dCK, are particularly sensitive to fludarabine due to the increased production of fludarabine-monophosphate [10]. At the present time, however, the major clinical utility of fludarabine is for the treatment of refractory chronic lymphocytic leukemia (CLL) [11]. Fludarabine has been tested for efficacy against a wide variety of solid tumors in the absence of radiation yielding unimpressive results. This may be potentially explained by the exceedingly low levels of dCK activity seen in non-lymphoid tissues [10]. However, more recent efforts using fludarabine in combination with radiation have shown promise in treating non-small cell lung cancer (NSCLC) and head and neck squamous cell carcinomas (HNSCC), at least in terms of tolerability and safety [12]. The major molecular actions of fludarabine tri-phosphate, which lead to cytotoxicity and radiosensitivity, may be explained in part by inhibition of DNA polymerases and inhibition of ribonucleotide reductase with consequent depletion of deoxynucleotide pools [6]. Incorporation of fludarabine into DNA can lead to chain termination and induction of apoptosis in a cell cycle specific manner [13]. Alternative mechanisms relating to inhibition of DNA repair machinery have been proposed to explain cell death initiation in quiescent tumor cells in response to fludarabine [14].

Cladribine

Cladribine, a chlorinated deoxyadenosine nucleoside analog, has been the agent of choice in the treatment of hairy cell leukemia since the early 1990s [15]. Cladribine has also demonstrated utility in the treatment of chronic myelogenous leukemias and non-hodgkins lymphomas but, similar to what has been reported for fludarabine, it has not produced impressive outcomes with solid tumors [1618]. The synthesis of cladribine in the 1960s grew out of efforts to produce agents with enhanced cytotoxicity and decreased susceptibility to catabolism using insight gained from studies of agents like ara-A [19]. Indeed, the chloride atom placed at the 2-position of the adenine ring is a key modification that interferes with catabolism through inhibition of deamination by ADA [20]. However, oral administration of cladribine is hindered by poor oral bioavailability due to degradation by the actions of bacterial PNP [21]. Thus, cladribine is administered intravenously (IV). Following IV infusion, cladribine is rapidly internalized by cells via the hENT1, hENT2, hCNT2, and hCNT3 transporters and phosphorylated to the monophosphate form by dCK and deoxyguanosine kinase (dGK) [2225], respectively. The relative role of dGK in mediating activation of cladribine is unclear in light of published data showing dGK is localized to the mitochondria whereas endogenous dCK is localized in the cytoplasm (exogenously overexpressed dCK results in nuclear localization) [26, 27]. Thus, phosphorylation by dCK is thought to be a critical event that is responsible for both enriching cladribine inside the cell and preparing it for its cytotoxic actions [28]. The triphosphate form of cladribine is achieved after successive phosphorylation of the mono- and di-phosphate forms by the nucleoside monophosphate kinase and the nucleoside diphosphate kinase. Once the triphosphate is generated it serves as an effective substrate for DNA replication enzymes like DNA polymerases [29, 30]. While incorporation of cladribine into DNA does not block chain extension per se, it is an inefficient substrate for extension and facilitates miss-incorporation of nucleotides [29]. Nevertheless, incorporation of cladribine into DNA leads to inhibition of DNA synthesis and, importantly, inhibition of DNA repair which in turn leads to formation of single strand breaks in DNA, poly(ADP-ribose) polymerase (PARP) activation and apoptosis via p53 dependent and independent pathways [3133]. Interestingly, cladribine is toxic to both resting and actively proliferating cells (a feature shared with fludarabine) possibly in a p53 dependent manner [34]. In addition to its effects on DNA synthesis and repair, the tri-phosphate form of cladribine has also been shown to inhibit ribonucleotide reductase (RR) [30]. Inhibition of RR by cladribine results in depletion of deoxynucleoside triphosphate (dNTP) pools leading to further inhibition of DNA synthesis and inappropriate activation of endonucleases that promote formation of stand breaks [35]. Other mechanisms regarding the cytotoxic actions of cladribine have also been reported. Fabianowska-Majewska et al. reported that cladribine can inhibit deoxyadenosine deamination and phosphorylation suggesting a role in regulating deoxyadenosine metabolism [36]. Several groups have demonstrated that cladribine may directly damage the mitochondria, disrupt mitochondrial function and promote the release of AIF [37, 38]. Despite the similarities in structure and mechanism to other nucleoside analogs, we are not aware of any successful clinical trials using cladribine as a radiosensitizer [39, 40].

Clofarabine

Clofarabine (2-Chloro-9-(2-deoxy-2-fluoro-β-D-arabinofuranosyl)-adenine) is a deoxyadenosine analog used in the treatment of acute lymphoblastic leukemia (ALL) and acute myelogenous leukemia (AML) [41]. A number of favorable structural features of other nucleoside analogs were incorporated into clofarabine to help improve its pharmacokinetic profile and reduce toxicity, without altering pharmacodynamics [41, 42]. For example, the addition of a fluoride atom to the 2’-position on the sugar group increases resistance to acidic conditions and bacterial PNP [43]. Also, reminiscent of modifications to cladribine and fludarabine, a halogen (chloride) at the 2-position of the adenine ring helps protect clofarabine from adenosine deaminase [43, 44]. In addition, the presence of this chloride atom appears to enhance the catalytic efficiency of dCK for clofarabine [45].

After administration, clofarabine is transported into the cell by the concerted action of three types of nucleoside transporters, hENT1, hENT2, and hCNT2 [22, 23]. Passive transport across the plasma membrane may also occur depending on the concentration of drug administered [22]. Upon entry into the cell, clofarabine undergoes rapid phosphorylation by dCK [41]. This is followed by rapid phosphorylation to the diphosphate form by purine nucleotide monophosphate kinase and to the triphosphate form by purine nucleotide diphosphate kinase [46]. The phosphorylation of clofarabine by dCK appears to occur with greater efficiency than fludarabine and cladribine [41, 47, 48]. Phosphorylation by dCK facilitates retention of clofarabine inside the cell thereby enriching intracellular drug concentrations. Indeed, the ABCG2 drug efflux pump can eject unphosphorylated clofarabine from the cell but not the monophosphate form [49]. As a result, the tissue distribution and relative expression levels of dCK in normal and cancer cells can influence both therapeutic efficacy and toxicity of clofarabine and other nucleoside analogs phosphorylated by dCK. Equally important is the fraction of active dCK present in normal cells as compared to cancerous cells. As discussed later in this review, while dCK is active in its native unphosphorylated state, its kinase activity is greatly enhanced when it is phosphorylated [50, 51]. Once phosphorylated to the tri-phosphate form, clofarabine acts as a fraudulent nucleoside and is incorporated into DNA but serves as a poor substrate for subsequent addition of nucleosides onto the growing chain [52]. This, in turn, results in chain termination and strand breaks. The triphosphate form of clofarabine has also been shown to interfere with DNA polymerase-α, but not β or γ [52, 53]. Finally, clofarabine-triphosphate is a potent inhibitor of ribonucleotide reductase (RR) that appears to result in an increase in clofarabine-triphosphate incorporation into DNA by depleting cellular concentrations of the normal, endogenous nucleotides [52, 53]. Several studies suggest that clofarabine ultimately promotes induction of apoptosis through a combination of direct and indirect effects on the mitochondria [37, 54]. As with the other radiosensitizing nucleoside analogs, in vivo activity against solid tumors have failed to reveal any objective responses in the absence of co-treatment with radiation [55]. However, the effectiveness of clofarabine for solid tumors has shown promise in vitro when used in tandem with radiation [56].

Nelarabine

Nelarabine is a prodrug of the guanosine analog, 9-β-D-arabinofuranosyl guanine (ara-G), that was granted accelerated approval by the FDA in 2005 for the treatment of T-cell ALL (T-ALL) and T-cell lymphoblastic lymphoma (T-LBL) [57]. Although ara-G was originally synthesized in the early 1960s its maturation into a viable clinical treatment modality was hindered because of its poor solubility [58]. However, in the 1970s work on human PNP deficiency rekindled interest in ara-G. Several key observations emerged from these studies: 1) human PNP deficiency can lead to depletion of T cells, 2) T cell cytotoxicity is associated with elevations in intracellular levels of dGTP (because dGTP is normally degraded by human PNP), and 3) B lymphocytes are largely unaffected, possibly as a result of differences in metabolism or cell cycle dependent accumulation of dGTP [5963]. Based on these observations, it became apparent that T cells would be subject to killing by a guanine based analog, such as ara-G, which is not subject to degradation by human PNP [64]. Subsequent studies revealed that the cytotoxic actions of ara-G were principally directed towards T cells [64]. This was followed by the successful synthesis of nelarabine via addition of a methyl group to the N6 position of the guanine ring [58]. Upon administration, nelarabine is converted to ara-G by plasma localized ADA [58]. Ara-G readily enters cells via the hENT whereupon it is rapidly phosphorylated by either dCK or dGK, in a rate limiting manner, to the monophosphate form [65, 66]. Phosphorylation to the di-phosphate and tri-phosphate forms may also be catalyzed by dGK [63, 66]. The intracellular concentrations of ara-GTP appear to be highly dependent and related to kinase activity of dCK or dGK. An increase in dCK activity or dGK activity facilitates higher intracellular concentrations of ara-GTP which shifts the preference of ara-G from dGK to dCK [66]. Additionally, the presence of a β hydroxyl group at the 2’-position of the sugar moiety leads to high intracellular levels of ara-G by reducing its susceptibility to human PNP [64]. Recent studies have suggested that the ABCB1 transporter may play a role in development of resistance by pumping ara-G out of the cell although the prevalence of this mechanism is uncertain [67]. Nevertheless, once sufficient intracellular levels of ara-GTP are reached, incorporation of ara-GTP into DNA blocks chain extension leading to strand breaks and, ultimately, apoptosis [64, 68]. Although there is currently no evidence that nelarabine or ara-G act as radiosensitizers, their reliance on dCK may make them subject to activation by the IR/ATM/dCK pathway thereby facilitating synergism as discussed below.

Pyrimidine based analogs

Cytarabine

The deoxycytidine analog cytarabine (also known as ara-C or 1-β-arabinofuranosylcytosine) has been in clinical use for leukemias such as acute myelogenous leukemia (AML) since its synthesis in the late 1950s. Ara-C resembles endogenous deoxycytidine in all respects save for the position of the 2’-hydroxyl group on the sugar moiety which is in the arabinose configuration to distinguish it from cytidine. After administration, ara-C is primarily transported into cells via the human equilibrative transporter (hENT1) although this is thought to be concentration dependent [69]. At high concentrations, however, ara-C may enter the cell by passive diffusion [70]. Once inside the cell ara-C is phosphorylated in sequence to the triphosphate form by dCK and pyrimidine nucleotide kinases [71]. As with other nucleoside analogs, phosphorylation can serve as a means to retain ara-C in the cell however, it has been noted that phosphorylated ara-C can be effluxed from the cell by multi-drug resistance proteins (MRPs) 5 and 7 [72]. The importance of dCK mediated phosphorylation is supported by studies that show resistance to ara-C in cells lacking dCK [73, 74]. Like other nucleoside analogs discussed thus far, it is the triphosphate form that is incorporated into DNA. Incorporation of ara-CTP into DNA occurs in competition with endogenous deoxycytidine tri-phosphate (dCTP) and once incorporated, the hydroxyl group on the ribose makes ara-CTP a poor substrate for chain extension. Ara-C appears to induce cell death via activation of the apoptotic program and inhibition of Bcl-2 expression re-sensitizes AML blasts to cell killing by Ara-C [75]. The mechanism of cell death may involve generation of reactive oxygen species [76]. As with ara-G, there is currently no evidence that ara-C can act as a radiosensitizer. However, because dCK is important for the activation of ara-C, there is a potential for chemosensitization by IR as discussed below.

Gemcitabine

The deoxycytidine analog, gemcitabine, is employed in the treatment of metastatic breast cancers, locally advanced or metastatic non-small cell lung cancers, pancreatic cancers, and relapsed ovarian cancers [7784]. To address metabolic limitations of ara-C, gemcitabine was structurally modified through the addition of two fluorine atoms in lieu of a hydroxyl group on the 2’-position of the ribose. Similar to ara-C, gemcitabine is internalized into target cells via the human equilbrative nucleoside transporter 1 (hENT1) although other nucleoside transporters appear to also play an important role in uptake [24, 85, 86]. Also, like ara-C, gemcitabine is phosphorylated to the monophosphate form by dCK [87]. Gemcitabine-monophosphate is converted to the di-phosphate and tri-phosphate form in succession by pyrimidine nucleotide kinases [88, 89]. Heinemann et al. demonstrated that, in contrast to ara-C tri-phosphate, gemcitabine tri-phosphate enters cells more rapidly, has a higher affinity for dCK and a slower elimination rate, leading to prolonged inhibition of DNA synthesis [87]. It has been documented that MRP 5 and 7 can pump gemcitabine out of the cell following internalization as a resistance mechanism [72]. Following phosphorylation to the triphosphate form, gemcitabine-triphosphate is incorporated into DNA leading to inhibition of DNA synthesis [90]. Interestingly, after the incorporation of gemcitabine-triphosphate into DNA a single, normal nucleotide is added to the 3’-hydroxyl of its ribose, shielding gemcitabine from DNA repair mechanisms including base excision repair [91]. The diphosphate form of gemcitabine is also a potent inhibitor of ribonucleotide reductase which leads to inhibition of DNA synthesis via depletion of deoxynucleotides [91]. As a consequence of this action, declining levels of dCTP de-inhibit dCK, increasing its activity favoring the generation of additional gemcitabine triphosphate [92]. Although the precise manner by which cell death is executed remains unclear, it is most likely mitochondrially mediated and caspase dependent [93, 94].

Mechanisms of synergism between nucleoside analogs and ionizing radiation

The clinical use of radiation in combination with chemotherapeutic agents gained significant momentum in the 1970s although many of the original studies date back to the early 1960s [95]. The underlying goal of these early efforts was a simple one, to synergistically increase tumor cell killing and improve patient outcomes [95]. The most studied means of synergism has been radiosensitization of cancer cells with nucleoside analogs. At the present time, gemcitabine, fludarabine, and clofarabine are employed clinically as radiosensitizers [96100]. Observations from these and other studies have revealed mechanistic commonalities between these agents that contribute to radiosensitization including inhibition of DNA repair and modulation of nucleotide synthesis/availability. Ultimately, it is thought that these effects culminate in cell cycle redistribution/arrest and inhibition of DNA synthesis. While many questions remain unanswered concerning how these mechanisms work together to achieve radiosensitization this topic has been reviewed extensively elsewhere [100, 101]. An alternative explanation of the synergism between radiation and nucleoside analogs, that remains underexplored, is IR-mediated chemosensitization. As discussed, not all nucleoside analogs act as radiosensitizers. Indeed, Nelarabine (ara-G), cytarabine (ara-C) and cladribine are not known to function as radiosensitizers despite having significant similarities in mechanism of action and metabolism to the radiosensitizing nucleoside analogs noted above. However, the ability to chemosensitize cells to these agents could represent an important strategy for synergism. Here we will briefly review the involvement of DNA repair inhibition in radiosensitization and contrast this mechanism with a recently identified pathway involving the ATM kinase and dCK that may lead to synergism through chemosensitization.

Radiosensitization through inhibition of DNA repair

Inhibition of DNA repair is one method for nucleoside analog induced radiosensitization. Indeed, the inhibition of DNA repair pathways is a logical means by which these drugs could sensitize cancer cells to the DNA damaging actions of ionizing radiation. However, while DNA repair pathways remain an attractive target, there are few published examples of this type of inhibition by nucleoside analogs. The nature of the interaction between the DNA repair machinery and nucleoside analogs that leads to enhanced radiosensitization remains poorly described. For example, Wachters et al. used cells that were deficient in either XRCC2 or XRCC3 to show that gemcitabine interferes with homologous recombination (HR) repair pathways possibly by inhibiting Rad51 [102]. These same authors had previously reported that gemcitabine radiosensitization was not dependent on NHEJ and in fact radiosensitization was enhanced in the absence of an intact NHEJ system [103]. These results correlate well with cell cycle studies of gemcitabine that demonstrate maximal radiosensitization occurs in cells that have progressed into S phase when HR would be most active [104, 105]. However, it is known that cells in S phase are more radioresistant compared to cells in other phases of the cell cycle. Also, there does not appear to be any significant increase in double strand break formation or repair with gemcitabine and radiation combination in tissue culture models [105, 106]. This is in contrast with more recent studies demonstrating increased γ-H2A.X formation (a marker of double strand breaks) by gemcitabine and clofarabine in cells with siRNA silenced Neil1 [107]. Neil1 is a key glycosylase that initiates BER. Thus, the importance of gemcitabine mediated inhibition of HR in promoting radiosensitization and the explanation behind enhancement of radiosensitization in S phase remain to be fully elucidated. In the case of fludarabine, several publications have shown that it can inhibit BER [106, 108]. A more recent study by Bulger et al. showed that the BER associated glycosylase UDG is upregulated in response to fludarabine in the leukemic cell line, HL60 [109]. Nevertheless, the consequences of fludarabines effects on BER in terms of radiosensitization remain unclear. Finally, a recent study by Stackhouse et al. examined the combination treatment with clofarabine and radiation where cell lines from several solid tumors were pre-treated with clofarabine for 1 hour followed by low dose IR treatment [56]. The most profound responses to this combination were seen in the head and neck cancer cell line SR475, the pancreatic cancer cell line PANC-1, and the colon cancer cell line HCT-116. The explanation as to how clofarabine radiosensitizes may relate to its ability to interfere with the DNA damage response and inhibit DNA repair [96]. Indeed, it has been reported that incorporation of clofarabine monophosphate into DNA may serve to inhibit DNA repair [41]. In general more studies are needed to validate the importance of DNA repair inhibition in mediating radiosensitization by nucleoside analogs. As a case in point, cladribine is known to inhibit DNA repair but it is a poor radiosensitizer.

Role of IR-induced activation of deoxycytidine kinase in chemo- and radiosensitization

A number of publications have demonstrated that radiation alone can enhance the activity of dCK [110112]. Interestingly, Csapo et al. show that increased dCK activity following low dose IR treatment is not a result of changes in dCK protein levels but rather due to post-translational modifications such as phosphorylation [110]. Given the critical role dCK plays in phosphorylating and activating agents such as gemcitabine, fludarabine, clofarabine, cladribine, nelarabine (ara-G), and cytarabine (ara-C) one would predict that cells with higher dCK activity (either intrinsically or via IR induction) would accumulate higher levels of active drug. This in turn would lead to enhanced cell cycle arrest, DNA damage by means of DNA repair inhibition or depletion of deoxynucleotide pools depending on the actions of the nucleoside analog in question. In support of this idea, Gregoire et al. went one step further by showing that increases in dCK activity directly correlate with radiosensitization exhibited by gemcitabine [97]. While these authors were able to demonstrate a tight correlation between the mRNA levels of dCK and radiosensitization the correlation between protein levels and radiosensitization was less robust again suggesting a role for post-transcriptional or post-translational modifications in dCK function and activity. The importance of post-translational modification of dCK was subsequently outlined in an eloquent series of studies by Bontemps and colleagues in which they demonstrated the role of phosphorylation in activating and stabilizing dCK kinase activity [50, 113115]. In these studies a number of candidate phosphorylation sites were identified including Thr-3, Ser-11, Ser-15, and most importantly, Ser-74 [50]. How dCK activation mediates the synergism between nucleoside analogs and radiation remained unclear until only recently. Our group has established that dCK can be phosphorylated by the DNA damage responsive kinase ATM on Ser-74, thereby directly linking radiation and dCK activation [51] (Figure 1). We further show that phosphorylated dCK can interact with and inhibit cyclin dependent kinase 1 (cdk1) which participates in governing the transition of cells from the G2 to M phase [116]. Thus increasing dCK activity via IR could potentiate synergism by creating a cellular environment favoring increased phosphorylation and activation of some, but not all nucleoside analogs [117]. Thus chemosensitization would occur as a result of the enhanced ability of nucleoside analogs to alter nucleotide synthesis and availability, cell cycle synchronization, and DNA repair processes. However, it is important to note that this synergism would not occur uniformly with all nucleoside analogs as evidenced by documented substrate preferences for S74 phosphorylated dCK or the S74E mutant [117, 118]. Additionally, given the observation that maximal radiosensitization occurs when nucleoside analogs are administered prior to radiation activation of dCK by IR may represent a secondary event that propagates synergism rather than initiate it. Future studies are needed to fully understand the role that this signaling pathway plays in chemo- and radiosensitization and, ultimately, its clinical utility.

Figure 1
figure 1

Activation of dCK after radiation contributes to enhanced therapeutic effects of nucleoside analogs. Treatment with ionizing radiation causes double strand breaks in DNA and activates the serine/threonine kinase Ataxia-telangiectasia mutated (ATM) which, in turn, phosphorylates deoxycytidine kinase, leading to a synergistic effect with nucleoside analogs. For example, gemcitabine enters the cell via the hENT1 transporter and is rapidly phosphorylated to the monophosphate form by active, serine 74 phosphorylated dCK prior to incorporation into DNA.

Clinical application of deoxycytidine kinase as a biomarker and drug target

As noted above, many genes involved in DNA repair, DNA damage response, and activating nucleoside analogs have been determined to mediate the synergism between nucleoside analogs and radiation. Assessing how these genes and their resultant proteins are altered in cancer or in response to treatment offers the promise of identifying biomarkers to predict the potential susceptibility of individual patients to combination chemotherapy and radiotherapy. Additionally, gaining understanding of how these genes function to promote or impair chemosensitization or radiosensitization could yield insight into how to therapeutically enhance these processes using small molecule or gene therapy approaches. Focusing on deoxycytidine kinase, we will review the active efforts to identify variants of dCK that can drive the activation of nucleoside analogs and then follow this by a discussion of work towards establishing high-throughput screening methods for identification of therapeutic modulators of dCK.

Several groups have used pharmacogenomic approaches to identify genetic variants of dCK. Lamba et al. conducted an extensive examination of dCK single nucleotide polymorphisms (SNPs) in both European and African populations [119]. They identified a total of 64 genetic variants of which 3, I24V, A119G, and P122S, were nonsynonymous changes in the coding region. Further analysis of these variants revealed that I24V, A119G, and P122S exhibited significantly reduced ability to phosphorylate cladribine as compared to wild type dCK. The expression of these variants was examined clinically in patients with AML receiving ara-C either in short infusions or continuously. However, due to the low numbers of patients with these nonsynonymous polymorphisms their clinical significance remains unclear and further study is needed. A subsequent study by Kocabas et al. confirmed these results and analyzed the implications of these and other single amino acid changes (I24V, A119G, and P122S) on the structural conformation of dCK [120]. They note that these amino acid substitutions could alter the local flexibility and destabilize the conformation of dCK, however, the overall effect on dCK activity or as a phosphorylation target itself remain unclear. Li et al. identified an additional SNP in dCK (rs4308342) located in an intron that appears to be associated with altered sensitivity of lymphoblastoid cells from ethnically diverse populations to gemcitabine and ara-C [121]. These studies and others have helped pry open the door to discerning the relative contribution of individual amino acids in the function of dCK though it is evident that not all mutations in dCK have prognostic value as biomarkers [122]. Interestingly, none of the mutants identified in these screens had alterations in either the active site or phosphorylation sites of dCK such as serine-74. However, several publications have demonstrated that loss or attenuation of dCK activity can have profound implications on the activation of gemcitabine and ara-C. Indeed, independent studies by Saiki et al. and Ohmine et al. used matched pancreatic cell lines that were either sensitive or resistant to gemcitabine and then used gene expression and proteomic analysis approaches to define the role of dCK in gemcitabine resistance [123, 124]. While it is becoming clear that dCK kinase activity is necessary for activation and efficacy of nucleoside analogs it is unclear if dCK phosphorylation site mutants are viable biomarkers or potential drug targets. However, phosphorylated dCK may be useful as a biomarker to gauge the functionality of dCK following radiotherapy but prior to administration of nucleoside analogs. Nevertheless, by gaining a more in depth understanding of how mutations in dCK alter its conformation and its ability to serve as a target for phosphorylation, it might be possible to screen or design small molecules to stimulate activation of dCK. However, key questions emerge such as: What types of dCK mutations might activate dCK? Can identification of dCK mutants with enhanced activity serve as a basis for small molecule drug design or gene therapy approaches?

To properly address these questions a high-throughput platform for identifying dCK mutants that have altered activity is needed. One such approach has been developed and tested by Rossolillo et al. who tested a retrovirus based system for generating screening libraries of gene mutants [125]. To validate their system they generated and identified mutant versions of dCK which, when over-expressed in cancer cells, alter susceptibility to gemcitabine. The most exciting mutant to emerge from this study is G12 a triple mutant that is altered at amino acids 171, 247, and 249 (E171K, E247K, and L249M). They demonstrate that although G12 phosphorylates gemcitabine as efficiently as wild type dCK, the G12 mutant exhibits significantly diminished ability to phosphorylate the endogenous dCK target, deoxycytidine as compared to wild-type dCK. Thus the G12 mutant is less likely to interfere with normal nucleotide synthesis catalyzed by dCK and instead is more directed towards gemcitabine activation. This suggests that the potential exists to modulate dCK to enhance its ability to phosphorylate nucleoside analog pro-drugs to active form. They also demonstrate that G12 has superior phosphorylation kinetics for gemcitabine compared to either the S74E mutant or the A100V, R104M, D133A triple mutant which also has altered substrate specificity [9, 118, 126]. They posit that because the E247 and L249 are located in the base sensing loop, which is thought to govern folding of dCK following binding of ATP or UTP, their mutation may explain the shift in substrate specificity seen with G12. Furthermore, the E171 residue located in alpha helix-7, which is involved in dCK dimer formation, may abrogate dCK dimer formation thus impairing its activity. Therefore this approach relies on generation of dCK mutants and validation of their activity prior to structural analysis.

An alternative approach that merits discussion uses insight gained from structural and functional studies of dCK to guide the design of dCK variants that can be expressed in cancer cells using gene therapy technology [127]. In this study, by Neschadim et al. dCK cDNA mutants were generated that exhibit altered activity and substrate specificity, and they were packaged them into lentiviral vectors for delivery to lymphoma or glioblastoma cells lines (Jurkat and MOLT-4 or U87mg, respectively). Mutation of dCK at arginine-104 and aspartic acid-133 have been previously demonstrated to alter the substrate specificity of dCK to include thymidine and deoxyuridine [126]. Still other studies have demonstrated that substitution of a glutamic acid residue in lieu of serine-74 leads to enhanced dCK activity by mimicking S74 phosphorylation [50, 118, 126]. For example, Hazra et al. sought to ascertain if expressing dCK double (R104M and D133A) or a triple (R104M, D133A, and S74E) mutants in cancer cells would increase the sensitivity of dCK to non-natural substrates like pro-drugs bromovinyl-deoxyuridine (BVdU) or L-thymidine (LdT) [128]. The cells transduced with triple mutants were most sensitive to cell death in response to treatment with both BVdU and LdT. The glioblastoma cell line U87mg was most sensitive followed by both lymphoma cell lines. These studies, therefore, offer proof of concept that dCK can serve as a potential biomarker or target for small molecule development.

Conclusions

In summary, a number of prominent nucleoside analogs have been shown to have a synergistic effect when used in combination with radiation. The underlying mechanisms behind this synergism remain poorly understood but may result from inhibition of DNA repair machinery, inhibition of DNA synthesis, cell cycle redistribution, or activation of nucleoside kinases such as dCK. Indeed, many of the currently used nucleoside analogs that have exhibited synergistic activity with radiotherapy are activated by dCK. It is well recognized that hematological malignancies, including many leukemias and lymphomas, express higher than normal levels of dCK and that this makes them more “sensitive” to nucleoside analog induced cell death. However, solid tumors do not exhibit a clear dCK expression pattern and in many cases they have low dCK expression levels. Thus, by leveraging recent developments in our understanding of dCK function and activation it may be possible to develop pharmacologic or genetic therapeutic approaches to increase the susceptibility of these tumors to radiation and antimetabolite combination therapy. Additionally, new insight on the function of dCK and mechanism of activation has applicability to nucleoside analogs in the pipeline currently such as thiarabine and sapacitabine.

References

  1. Montgomery JA, Hewson K: Nucleosides of 2-fluoroadenine. J Med Chem 1969,12(3):498-504. 10.1021/jm00303a605

    CAS  PubMed  Google Scholar 

  2. Huang P, Plunkett W: Phosphorolytic cleavage of 2-fluoroadenine from 9-beta-D-arabinofuranosyl-2-fluoroadenine by Escherichia coli. A pathway for 2-fluoro-ATP production. Biochem Pharmacol 1987,36(18):2945-2950. 10.1016/0006-2952(87)90207-3

    CAS  PubMed  Google Scholar 

  3. Pugmire MJ, Ealick SE: Structural analyses reveal two distinct families of nucleoside phosphorylases. Biochem J 2002,361(Pt 1):1-25.

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Brockman RW, Schabel FM Jr, Montgomery JA: Biologic activity of 9-beta-D-arabinofuranosyl-2-fluoroadenine, a metabolically stable analog of 9-beta-D-arabinofuranosyladenine. Biochem Pharmacol 1977,26(22):2193-2196. 10.1016/0006-2952(77)90275-1

    CAS  PubMed  Google Scholar 

  5. Danhauser L, et al.: 9-beta-D-arabinofuranosyl-2-fluoroadenine 5'-monophosphate pharmacokinetics in plasma and tumor cells of patients with relapsed leukemia and lymphoma. Cancer Chemother Pharmacol 1986,18(2):145-152.

    CAS  PubMed  Google Scholar 

  6. Gandhi V, Plunkett W: Cellular and clinical pharmacology of fludarabine. Clin Pharmacokinet 2002,41(2):93-103. 10.2165/00003088-200241020-00002

    CAS  PubMed  Google Scholar 

  7. Shewach DS, Reynolds KK, Hertel L: Nucleotide specificity of human deoxycytidine kinase. Mol Pharmacol 1992,42(3):518-524.

    CAS  PubMed  Google Scholar 

  8. Krenitsky TA, et al.: Deoxycytidine kinase from calf thymus. Substrate and inhibitor specificity. J Biol Chem 1976,251(13):4055-4061.

    CAS  PubMed  Google Scholar 

  9. Sabini E, et al.: Structure of human dCK suggests strategies to improve anticancer and antiviral therapy. Nat Struct Biol 2003,10(7):513-519. 10.1038/nsb942

    CAS  PubMed  Google Scholar 

  10. Arner ES, Eriksson S: Mammalian deoxyribonucleoside kinases. Pharmacol Ther 1995,67(2):155-186. 10.1016/0163-7258(95)00015-9

    CAS  PubMed  Google Scholar 

  11. Grever M, et al.: A comprehensive phase I and II clinical investigation of fludarabine phosphate. Semin Oncol 1990,17(5 Suppl 8):39-48.

    CAS  PubMed  Google Scholar 

  12. Nitsche M, et al.: Fludarabine combined with radiotherapy in patients with locally advanced NSCLC lung carcinoma: a phase I study. J Cancer Res Clin Oncol 2012,138(7):1113-1120. 10.1007/s00432-012-1185-3

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Consoli U, et al.: Differential induction of apoptosis by fludarabine monophosphate in leukemic B and normal T cells in chronic lymphocytic leukemia. Blood 1998,91(5):1742-1748.

    CAS  PubMed  Google Scholar 

  14. Sandoval A, Consoli U, Plunkett W: Fludarabine-mediated inhibition of nucleotide excision repair induces apoptosis in quiescent human lymphocytes. Clin Cancer Res 1996,2(10):1731-1741.

    CAS  PubMed  Google Scholar 

  15. Piro LD, et al.: Lasting remissions in hairy-cell leukemia induced by a single infusion of 2-chlorodeoxyadenosine. N Engl J Med 1990,322(16):1117-1121. 10.1056/NEJM199004193221605

    CAS  PubMed  Google Scholar 

  16. Santana VM, et al.: Complete hematologic remissions induced by 2-chlorodeoxyadenosine in children with newly diagnosed acute myeloid leukemia. Blood 1994,84(4):1237-1242.

    CAS  PubMed  Google Scholar 

  17. Saven A, et al.: 2-Chlorodeoxyadenosine activity in patients with untreated, indolent non-Hodgkin's lymphoma. Blood 1995,86(5):1710-1716.

    CAS  PubMed  Google Scholar 

  18. Saven A, Piro LD: 2-Chlorodeoxyadenosine: a new nucleoside agent effective in the treatment of lymphoid malignancies. Leuk Lymphoma 1993,10(Suppl):43-49.

    PubMed  Google Scholar 

  19. Cass CE, Au-Yeung TH: Enhancement of 9-beta-d-arabinofuranosyladenine cytotoxicity to mouse leukemia L1210 in vitro by 2'-deoxycoformycin. Cancer Res 1976,36(4):1486-1491.

    CAS  PubMed  Google Scholar 

  20. Carson DA, et al.: Biochemical basis for the enhanced toxicity of deoxyribonucleosides toward malignant human T cell lines. Proc Natl Acad Sci U S A 1979,76(5):2430-2433. 10.1073/pnas.76.5.2430

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Lotfi K, Juliusson G, Albertioni F: Pharmacological basis for cladribine resistance. Leuk Lymphoma 2003,44(10):1705-1712. 10.1080/1042819031000099698

    CAS  PubMed  Google Scholar 

  22. King KM, et al.: A comparison of the transportability, and its role in cytotoxicity, of clofarabine, cladribine, and fludarabine by recombinant human nucleoside transporters produced in three model expression systems. Mol Pharmacol 2006,69(1):346-353.

    CAS  PubMed  Google Scholar 

  23. Leung GP, Tse CM: The role of mitochondrial and plasma membrane nucleoside transporters in drug toxicity. Expert Opin Drug Metab Toxicol 2007,3(5):705-718. 10.1517/17425255.3.5.705

    CAS  PubMed  Google Scholar 

  24. Mackey JR, et al.: Functional nucleoside transporters are required for gemcitabine influx and manifestation of toxicity in cancer cell lines. Cancer Res 1998,58(19):4349-4357.

    CAS  PubMed  Google Scholar 

  25. Wang L, et al.: Substrate specificity of mitochondrial 2'-deoxyguanosine kinase. Efficient phosphorylation of 2-chlorodeoxyadenosine. J Biol Chem 1993,268(30):22847-22852.

    CAS  PubMed  Google Scholar 

  26. Hatzis P, et al.: The intracellular localization of deoxycytidine kinase. J Biol Chem 1998,273(46):30239-30243. 10.1074/jbc.273.46.30239

    CAS  PubMed  Google Scholar 

  27. Johansson M, Brismar S, Karlsson A: Human deoxycytidine kinase is located in the cell nucleus. Proc Natl Acad Sci U S A 1997,94(22):11941-11945. 10.1073/pnas.94.22.11941

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Kawasaki H, et al.: Relationship of deoxycytidine kinase and cytoplasmic 5'-nucleotidase to the chemotherapeutic efficacy of 2-chlorodeoxyadenosine. Blood 1993,81(3):597-601.

    CAS  PubMed  Google Scholar 

  29. Hentosh P, Koob R, Blakley RL: Incorporation of 2-halogeno-2'-deoxyadenosine 5-triphosphates into DNA during replication by human polymerases alpha and beta. J Biol Chem 1990,265(7):4033-4040.

    CAS  PubMed  Google Scholar 

  30. Parker WB, et al.: Interaction of 2-halogenated dATP analogs (F, Cl, and Br) with human DNA polymerases, DNA primase, and ribonucleotide reductase. Mol Pharmacol 1988,34(4):485-491.

    CAS  PubMed  Google Scholar 

  31. Pettitt AR, Sherrington PD, Cawley JC: Role of poly(ADP-ribosyl)ation in the killing of chronic lymphocytic leukemia cells by purine analogues. Cancer Res 2000,60(15):4187-4193.

    CAS  PubMed  Google Scholar 

  32. Robertson LE, et al.: Induction of apoptotic cell death in chronic lymphocytic leukemia by 2-chloro-2'-deoxyadenosine and 9-beta-D-arabinosyl-2-fluoroadenine. Blood 1993,81(1):143-150.

    CAS  PubMed  Google Scholar 

  33. Seto S, et al.: Mechanism of deoxyadenosine and 2-chlorodeoxyadenosine toxicity to nondividing human lymphocytes. J Clin Invest 1985,75(2):377-383. 10.1172/JCI111710

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Pettitt AR, et al.: Purine analogues kill resting lymphocytes by p53-dependent and -independent mechanisms. Br J Haematol 1999,105(4):986-988. 10.1046/j.1365-2141.1999.01448.x

    CAS  PubMed  Google Scholar 

  35. Griffig J, Koob R, Blakley RL: Mechanisms of inhibition of DNA synthesis by 2-chlorodeoxyadenosine in human lymphoblastic cells. Cancer Res 1989,49(24 Pt 1):6923-6928.

    CAS  PubMed  Google Scholar 

  36. Fabianowska-Majewska K, et al.: The influence of 2-chloro-2'-deoxyadenosine on metabolism of deoxyadenosine in human primary CNS lymphoma. Biochem Pharmacol 1995,50(9):1379-1383. 10.1016/0006-2952(95)02018-7

    CAS  PubMed  Google Scholar 

  37. Genini D, et al.: Deoxyadenosine analogs induce programmed cell death in chronic lymphocytic leukemia cells by damaging the DNA and by directly affecting the mitochondria. Blood 2000,96(10):3537-3543.

    CAS  PubMed  Google Scholar 

  38. Hentosh P, Tibudan M: 2-Chloro-2'-deoxyadenosine, an antileukemic drug, has an early effect on cellular mitochondrial function. Mol Pharmacol 1997,51(4):613-619.

    CAS  PubMed  Google Scholar 

  39. Cheson BD: Perspectives on purine analogues. Hematol Cell Ther 1996,38(Suppl 2):S109-S116.

    CAS  PubMed  Google Scholar 

  40. Saven A, Piro LD: 2-Chlorodeoxyadenosine: a newer purine analog active in the treatment of indolent lymphoid malignancies. Ann Intern Med 1994,120(9):784-791. 10.7326/0003-4819-120-9-199405010-00010

    CAS  PubMed  Google Scholar 

  41. Bonate PL, et al.: Discovery and development of clofarabine: a nucleoside analogue for treating cancer. Nat Rev Drug Discov 2006,5(10):855-863. 10.1038/nrd2055

    CAS  PubMed  Google Scholar 

  42. Avramis VI, Plunkett W: 2-fluoro-ATP: a toxic metabolite of 9-beta-D-arabinosyl-2-fluoroadenine. Biochem Biophys Res Commun 1983,113(1):35-43. 10.1016/0006-291X(83)90428-X

    CAS  PubMed  Google Scholar 

  43. Montgomery JA, et al.: Synthesis and biologic activity of 2'-fluoro-2-halo derivatives of 9-beta-D-arabinofuranosyladenine. J Med Chem 1992,35(2):397-401. 10.1021/jm00080a029

    CAS  PubMed  Google Scholar 

  44. Galmarini CM, Mackey JR, Dumontet C: Nucleoside analogues and nucleobases in cancer treatment. Lancet Oncol 2002,3(7):415-424. 10.1016/S1470-2045(02)00788-X

    CAS  PubMed  Google Scholar 

  45. Zhang Y, Secrist JA 3rd, Ealick SE: The structure of human deoxycytidine kinase in complex with clofarabine reveals key interactions for prodrug activation. Acta Crystallogr D Biol Crystallogr 2006,62(Pt 2):133-139.

    PubMed  Google Scholar 

  46. Zhenchuk A, et al.: Mechanisms of anti-cancer action and pharmacology of clofarabine. Biochem Pharmacol 2009,78(11):1351-1359. 10.1016/j.bcp.2009.06.094

    CAS  PubMed  Google Scholar 

  47. Lotfi K, et al.: Biochemical pharmacology and resistance to 2-chloro-2'-arabino-fluoro-2'-deoxyadenosine, a novel analogue of cladribine in human leukemic cells. Clin Cancer Res 1999,5(9):2438-2444.

    CAS  PubMed  Google Scholar 

  48. Parker WB, et al.: Comparison of the mechanism of cytotoxicity of 2-chloro-9-(2-deoxy-2- fluoro-beta-D-arabinofuranosyl)adenine, 2-chloro-9-(2-deoxy-2-fluoro- beta-D-ribofuranosyl)adenine, and 2-chloro-9-(2-deoxy-2,2-difluoro- beta-D-ribofuranosyl)adenine in CEM cells. Mol Pharmacol 1999,55(3):515-520.

    CAS  PubMed  Google Scholar 

  49. Nagai S, et al.: Deoxycytidine kinase modulates the impact of the ABC transporter ABCG2 on clofarabine cytotoxicity. Cancer Res 2011,71(5):1781-1791. 10.1158/0008-5472.CAN-10-1919

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Smal C, et al.: Identification of in vivo phosphorylation sites on human deoxycytidine kinase. Role of Ser-74 in the control of enzyme activity. J Biol Chem 2006,281(8):4887-4893.

    CAS  PubMed  Google Scholar 

  51. Yang C, et al.: Deoxycytidine kinase regulates the G2/M checkpoint through interaction with cyclin-dependent kinase 1 in response to DNA damage. Nucleic Acids Res 2012,40(19):9621-9632. 10.1093/nar/gks707

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Parker WB, et al.: Effects of 2-chloro-9-(2-deoxy-2-fluoro-beta-D-arabinofuranosyl)adenine on K562 cellular metabolism and the inhibition of human ribonucleotide reductase and DNA polymerases by its 5'-triphosphate. Cancer Res 1991,51(9):2386-2394.

    CAS  PubMed  Google Scholar 

  53. Xie C, Plunkett W: Metabolism and actions of 2-chloro-9-(2-deoxy-2-fluoro-beta-D- arabinofuranosyl)-adenine in human lymphoblastoid cells. Cancer Res 1995,55(13):2847-2852.

    CAS  PubMed  Google Scholar 

  54. Genini D, et al.: Nucleotide requirements for the in vitro activation of the apoptosis protein-activating factor-1-mediated caspase pathway. J Biol Chem 2000,275(1):29-34. 10.1074/jbc.275.1.29

    CAS  PubMed  Google Scholar 

  55. Kantarjian HM, et al.: Phase I clinical and pharmacology study of clofarabine in patients with solid and hematologic cancers. J Clin Oncol 2003,21(6):1167-1173. 10.1200/JCO.2003.04.031

    CAS  PubMed  Google Scholar 

  56. Stackhouse MA, et al.: Preclinical combination therapy of clofarabine plus radiation. Nucleosides Nucleotides Nucleic Acids 2012,31(9):692-705. 10.1080/15257770.2012.723770

    CAS  PubMed  Google Scholar 

  57. Gandhi V, Plunkett W: Clofarabine and nelarabine: two new purine nucleoside analogs. Curr Opin Oncol 2006,18(6):584-590. 10.1097/01.cco.0000245326.65152.af

    CAS  PubMed  Google Scholar 

  58. Lambe CU, et al.: 2-Amino-6-methoxypurine arabinoside: an agent for T-cell malignancies. Cancer Res 1995,55(15):3352-3356.

    CAS  PubMed  Google Scholar 

  59. Cohen A, et al.: Deoxyguanosine triphosphate as a possible toxic metabolite in the immunodeficiency associated with purine nucleoside phosphorylase deficiency. J Clin Invest 1978,61(5):1405-1409. 10.1172/JCI109058

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Cohen A, et al.: The expression of deoxyguanosine toxicity in T lymphocytes at different stages of maturation. J Immunol 1980,125(4):1578-1582.

    CAS  PubMed  Google Scholar 

  61. Gelfand EW, Lee JJ, Dosch HM: Selective toxicity of purine deoxynucleosides for human lymphocyte growth and function. Proc Natl Acad Sci U S A 1979,76(4):1998-2002. 10.1073/pnas.76.4.1998

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Giblett ER, et al.: Nucleoside-phosphorylase deficiency in a child with severely defective T-cell immunity and normal B-cell immunity. Lancet 1975,1(7914):1010-1013.

    CAS  PubMed  Google Scholar 

  63. Rodriguez CO Jr, Stellrecht CM, Gandhi V: Mechanisms for T-cell selective cytotoxicity of arabinosylguanine. Blood 2003,102(5):1842-1848. 10.1182/blood-2003-01-0317

    CAS  PubMed  Google Scholar 

  64. Cohen A, Lee JW, Gelfand EW: Selective toxicity of deoxyguanosine and arabinosyl guanine for T-leukemic cells. Blood 1983,61(4):660-666.

    CAS  PubMed  Google Scholar 

  65. Prus KL, Averett DR, Zimmerman TP: Transport and metabolism of 9-beta-D-arabinofuranosylguanine in a human T-lymphoblastoid cell line: nitrobenzylthioinosine-sensitive and -insensitive influx. Cancer Res 1990,50(6):1817-1821.

    CAS  PubMed  Google Scholar 

  66. Rodriguez CO Jr, et al.: Arabinosylguanine is phosphorylated by both cytoplasmic deoxycytidine kinase and mitochondrial deoxyguanosine kinase. Cancer Res 2002,62(11):3100-3105.

    CAS  PubMed  Google Scholar 

  67. Fyrberg A, et al.: Induction of fetal hemoglobin and ABCB1 gene expression in 9-beta-D-arabinofuranosylguanine-resistant MOLT-4 cells. Cancer Chemother Pharmacol 2011,68(3):583-591. 10.1007/s00280-010-1524-5

    CAS  PubMed  Google Scholar 

  68. Rodriguez CO Jr, Gandh V: Arabinosylguanine-induced apoptosis of T-lymphoblastic cells: incorporation into DNA is a necessary step. Cancer Res 1999,59(19):4937-4943.

    CAS  PubMed  Google Scholar 

  69. Sundaram M, et al.: Topology of a human equilibrative, nitrobenzylthioinosine (NBMPR)-sensitive nucleoside transporter (hENT1) implicated in the cellular uptake of adenosine and anti-cancer drugs. J Biol Chem 2001,276(48):45270-45275. 10.1074/jbc.M107169200

    CAS  PubMed  Google Scholar 

  70. Capizzi RL, et al.: Alteration of the pharmacokinetics of high-dose ara-C by its metabolite, high ara-U in patients with acute leukemia. J Clin Oncol 1983,1(12):763-771.

    CAS  PubMed  Google Scholar 

  71. Verhoef V, Sarup J, Fridland A: Identification of the mechanism of activation of 9-beta-D-arabinofuranosyladenine in human lymphoid cells using mutants deficient in nucleoside kinases. Cancer Res 1981,41(11 Pt 1):4478-4483.

    CAS  PubMed  Google Scholar 

  72. Chen ZS, Tiwari AK: Multidrug resistance proteins (MRPs/ABCCs) in cancer chemotherapy and genetic diseases. FEBS J 2011,278(18):3226-3245. 10.1111/j.1742-4658.2011.08235.x

    CAS  PubMed  PubMed Central  Google Scholar 

  73. Bhalla K, Nayak R, Grant S: Isolation and characterization of a deoxycytidine kinase-deficient human promyelocytic leukemic cell line highly resistant to 1-beta-D- arabinofuranosylcytosine. Cancer Res 1984,44(11):5029-5037.

    CAS  PubMed  Google Scholar 

  74. Owens JK, et al.: Resistance to 1-beta-D-arabinofuranosylcytosine in human T-lymphoblasts mediated by mutations within the deoxycytidine kinase gene. Cancer Res 1992,52(9):2389-2393.

    CAS  PubMed  Google Scholar 

  75. Keith FJ, et al.: Inhibition of bcl-2 with antisense oligonucleotides induces apoptosis and increases the sensitivity of AML blasts to Ara-C. Leukemia 1995,9(1):131-138.

    CAS  PubMed  Google Scholar 

  76. Iacobini M, et al.: Involvement of oxygen radicals in cytarabine-induced apoptosis in human polymorphonuclear cells. Biochem Pharmacol 2001,61(8):1033-1040. 10.1016/S0006-2952(01)00548-2

    CAS  PubMed  Google Scholar 

  77. Seidman AD: Gemcitabine as single-agent therapy in the management of advanced breast cancer. Oncology (Williston Park) 2001,15(2 Suppl 3):11-14.

    CAS  Google Scholar 

  78. Carmichael J, et al.: Advanced breast cancer: a phase II trial with gemcitabine. J Clin Oncol 1995,13(11):2731-2736.

    CAS  PubMed  Google Scholar 

  79. Pollera CF, et al.: Weekly gemcitabine in advanced bladder cancer: a preliminary report from a phase I study. Ann Oncol 1994,5(2):182-184.

    CAS  PubMed  Google Scholar 

  80. Lund B, et al.: Phase II study of gemcitabine (2',2'-difluorodeoxycytidine) in previously treated ovarian cancer patients. J Natl Cancer Inst 1994,86(20):1530-1533. 10.1093/jnci/86.20.1530

    CAS  PubMed  Google Scholar 

  81. Anderson H, et al.: Single-agent activity of weekly gemcitabine in advanced non-small-cell lung cancer: a phase II study. J Clin Oncol 1994,12(9):1821-1826.

    CAS  PubMed  Google Scholar 

  82. Abratt RP, et al.: Efficacy and safety profile of gemcitabine in non-small-cell lung cancer: a phase II study. J Clin Oncol 1994,12(8):1535-1540.

    CAS  PubMed  Google Scholar 

  83. Casper ES, et al.: Phase II trial of gemcitabine (2,2'-difluorodeoxycytidine) in patients with adenocarcinoma of the pancreas. Invest New Drugs 1994,12(1):29-34. 10.1007/BF00873232

    CAS  PubMed  Google Scholar 

  84. Moore M: Activity of gemcitabine in patients with advanced pancreatic carcinoma. A review. Cancer 1996,78(3 Suppl):633-638.

    CAS  PubMed  Google Scholar 

  85. Rauchwerger DR, et al.: Equilibrative-sensitive nucleoside transporter and its role in gemcitabine sensitivity. Cancer Res 2000,60(21):6075-6079.

    CAS  PubMed  Google Scholar 

  86. Damaraju VL, et al.: Nucleoside anticancer drugs: the role of nucleoside transporters in resistance to cancer chemotherapy. Oncogene 2003,22(47):7524-7536. 10.1038/sj.onc.1206952

    CAS  PubMed  Google Scholar 

  87. Heinemann V, et al.: Comparison of the cellular pharmacokinetics and toxicity of 2',2'-difluorodeoxycytidine and 1-beta-D-arabinofuranosylcytosine. Cancer Res 1988,48(14):4024-4031.

    CAS  PubMed  Google Scholar 

  88. Galmarini CM, Mackey JR, Dumontet C: Nucleoside analogues: mechanisms of drug resistance and reversal strategies. Leukemia 2001,15(6):875-890. 10.1038/sj.leu.2402114

    CAS  PubMed  Google Scholar 

  89. Plunkett W, et al.: Gemcitabine: metabolism, mechanisms of action, and self-potentiation. Semin Oncol 1995,22(4 Suppl 11):3-10.

    CAS  PubMed  Google Scholar 

  90. Huang P, et al.: Action of 2',2'-difluorodeoxycytidine on DNA synthesis. Cancer Res 1991,51(22):6110-6117.

    CAS  PubMed  Google Scholar 

  91. Huang P, Plunkett W: Induction of apoptosis by gemcitabine. Semin Oncol 1995,22(4 Suppl 11):19-25.

    PubMed  Google Scholar 

  92. Heinemann V, et al.: Inhibition of ribonucleotide reduction in CCRF-CEM cells by 2',2'-difluorodeoxycytidine. Mol Pharmacol 1990,38(4):567-572.

    CAS  PubMed  Google Scholar 

  93. Chandler NM, Canete JJ, Callery MP: Caspase-3 drives apoptosis in pancreatic cancer cells after treatment with gemcitabine. J Gastrointest Surg 2004,8(8):1072-1078. 10.1016/j.gassur.2004.09.054

    PubMed  Google Scholar 

  94. Nabhan C, et al.: Caspase activation is required for gemcitabine activity in multiple myeloma cell lines. Mol Cancer Ther 2002,1(13):1221-1227.

    CAS  PubMed  Google Scholar 

  95. Bernier J, Hall EJ, Giaccia A: Radiation oncology: a century of achievements. Nat Rev Cancer 2004,4(9):737-747.

    CAS  PubMed  Google Scholar 

  96. Cariveau MJ, et al.: Clofarabine acts as radiosensitizer in vitro and in vivo by interfering with DNA damage response. Int J Radiat Oncol Biol Phys 2008,70(1):213-220. 10.1016/j.ijrobp.2007.09.012

    CAS  PubMed  Google Scholar 

  97. Gregoire V, et al.: Role of deoxycytidine kinase (dCK) activity in gemcitabine's radioenhancement in mice and human cell lines in vitro. Radiother Oncol 2002,63(3):329-338. 10.1016/S0167-8140(02)00106-8

    CAS  PubMed  Google Scholar 

  98. Latourette HB, Lawton RL: Combined Radiation and Chemotherapy. JAMA 1963, 186: 1057-1060. 10.1001/jama.1963.03710120039008

    CAS  PubMed  Google Scholar 

  99. Shewach DS, et al.: Metabolism of 2',2'-difluoro-2'-deoxycytidine and radiation sensitization of human colon carcinoma cells. Cancer Res 1994,54(12):3218-3223.

    CAS  PubMed  Google Scholar 

  100. Shewach DS, Lawrence TS: Antimetabolite radiosensitizers. J Clin Oncol 2007,25(26):4043-4050. 10.1200/JCO.2007.11.5287

    CAS  PubMed  Google Scholar 

  101. Ewald B, Sampath D, Plunkett W: Nucleoside analogs: molecular mechanisms signaling cell death. Oncogene 2008,27(50):6522-6537. 10.1038/onc.2008.316

    CAS  PubMed  Google Scholar 

  102. Wachters FM, et al.: Selective targeting of homologous DNA recombination repair by gemcitabine. Int J Radiat Oncol Biol Phys 2003,57(2):553-562. 10.1016/S0360-3016(03)00503-0

    CAS  PubMed  Google Scholar 

  103. van Putten JWG, et al.: End-joining deficiency and radiosensitization induced by gemcitabine. Cancer Res 2001,61(4):1585-1591.

    CAS  PubMed  Google Scholar 

  104. Latz D, et al.: Radiosensitizing potential of gemcitabine (2',2'-difluoro-2'-deoxycytidine) within the cell cycle in vitro. Int J Radiat Oncol Biol Phys 1998,41(4):875-882. 10.1016/S0360-3016(98)00105-9

    CAS  PubMed  Google Scholar 

  105. Lawrence TS, et al.: Delayed radiosensitization of human colon carcinoma cells after a brief exposure to 2',2'-difluoro-2'-deoxycytidine (Gemcitabine). Clin Cancer Res 1997,3(5):777-782.

    CAS  PubMed  Google Scholar 

  106. Gregoire V, et al.: Radiosensitization of mouse sarcoma cells by fludarabine (F-ara-A) or gemcitabine (dFdC), two nucleoside analogues, is not mediated by an increased induction or a repair inhibition of DNA double-strand breaks as measured by pulsed-field gel electrophoresis. Int J Radiat Biol 1998,73(5):511-520. 10.1080/095530098142059

    CAS  PubMed  Google Scholar 

  107. Taricani L, et al.: Phenotypic enhancement of thymidylate synthetase pathway inhibitors following ablation of Neil1 DNA glycosylase/lyase. Cell Cycle 2010,9(24):4876-4883. 10.4161/cc.9.24.14155

    CAS  PubMed  Google Scholar 

  108. Yamauchi T, et al.: DNA repair initiated in chronic lymphocytic leukemia lymphocytes by 4-hydroperoxycyclophosphamide is inhibited by fludarabine and clofarabine. Clin Cancer Res 2001,7(11):3580-3589.

    CAS  PubMed  Google Scholar 

  109. Bulgar AD, et al.: Targeting base excision repair suggests a new therapeutic strategy of fludarabine for the treatment of chronic lymphocytic leukemia. Leukemia 2010,24(10):1795-1799. 10.1038/leu.2010.166

    CAS  PubMed  Google Scholar 

  110. Csapo Z, et al.: Activation of deoxycytidine kinase by gamma-irradiation and inactivation by hyperosmotic shock in human lymphocytes. Biochem Pharmacol 2003,65(12):2031-2039. 10.1016/S0006-2952(03)00182-5

    CAS  PubMed  Google Scholar 

  111. Pauwels B, et al.: The relation between deoxycytidine kinase activity and the radiosensitising effect of gemcitabine in eight different human tumour cell lines. BMC Cancer 2006, 6: 142. 10.1186/1471-2407-6-142

    PubMed  PubMed Central  Google Scholar 

  112. Sigmond J, et al.: Enhanced activity of deoxycytidine kinase after pulsed low dose rate and single dose gamma irradiation. Nucleosides Nucleotides Nucleic Acids 2006,25(9–11):1177-1180.

    CAS  PubMed  Google Scholar 

  113. Smal C, et al.: Activation of deoxycytidine kinase by protein kinase inhibitors and okadaic acid in leukemic cells. Biochem Pharmacol 2004,68(1):95-103. 10.1016/j.bcp.2004.02.031

    CAS  PubMed  Google Scholar 

  114. Smal C, et al.: Positive regulation of deoxycytidine kinase activity by phosphorylation of Ser-74 in B-cell chronic lymphocytic leukaemia lymphocytes. Cancer Lett 2007,253(1):68-73. 10.1016/j.canlet.2007.01.013

    CAS  PubMed  Google Scholar 

  115. Smal C, et al.: Casein kinase 1delta activates human recombinant deoxycytidine kinase by Ser-74 phosphorylation, but is not involved in the in vivo regulation of its activity. Arch Biochem Biophys 2010,502(1):44-52. 10.1016/j.abb.2010.07.009

    CAS  PubMed  Google Scholar 

  116. Ubersax JA, et al.: Targets of the cyclin-dependent kinase Cdk1. Nature 2003,425(6960):859-864. 10.1038/nature02062

    CAS  PubMed  Google Scholar 

  117. Amsailale R, et al.: Phosphorylation of deoxycytidine kinase on Ser-74: impact on kinetic properties and nucleoside analog activation in cancer cells. Biochem Pharmacol 2012,84(1):43-51. 10.1016/j.bcp.2012.03.022

    CAS  PubMed  Google Scholar 

  118. McSorley T, et al.: Mimicking phosphorylation of Ser-74 on human deoxycytidine kinase selectively increases catalytic activity for dC and dC analogues. FEBS Lett 2008,582(5):720-724. 10.1016/j.febslet.2008.01.048

    CAS  PubMed  PubMed Central  Google Scholar 

  119. Lamba JK, et al.: Pharmacogenetics of deoxycytidine kinase: identification and characterization of novel genetic variants. J Pharmacol Exp Ther 2007,323(3):935-945. 10.1124/jpet.107.128595

    CAS  PubMed  Google Scholar 

  120. Kocabas NA, et al.: Gemcitabine pharmacogenomics: deoxycytidine kinase and cytidylate kinase gene resequencing and functional genomics. Drug Metab Dispos 2008,36(9):1951-1959. 10.1124/dmd.108.020925

    CAS  PubMed  PubMed Central  Google Scholar 

  121. Li L, et al.: Gemcitabine and arabinosylcytosin pharmacogenomics: genome-wide association and drug response biomarkers. PLoS One 2009,4(11):e7765. 10.1371/journal.pone.0007765

    PubMed  PubMed Central  Google Scholar 

  122. Ryu JS, et al.: Lack of association of genetic variations of deoxycytidine kinase with toxicity or survival of non-small-cell lung cancer patients treated with gemcitabine plus cisplatin. Oncol Res 2012,20(1):25-30. 10.3727/096504012X13425470196137

    PubMed  Google Scholar 

  123. Ohmine K, et al.: Attenuation of phosphorylation by deoxycytidine kinase is key to acquired gemcitabine resistance in a pancreatic cancer cell line: targeted proteomic and metabolomic analyses in PK9 cells. Pharm Res 2012,29(7):2006-2016. 10.1007/s11095-012-0728-2

    CAS  PubMed  Google Scholar 

  124. Saiki Y, et al.: DCK is frequently inactivated in acquired gemcitabine-resistant human cancer cells. Biochem Biophys Res Commun 2012,421(1):98-104. 10.1016/j.bbrc.2012.03.122

    CAS  PubMed  Google Scholar 

  125. Rossolillo P, et al.: Retrovolution: HIV-driven evolution of cellular genes and improvement of anticancer drug activation. PLoS Genet 2012,8(8):e1002904. 10.1371/journal.pgen.1002904

    CAS  PubMed  PubMed Central  Google Scholar 

  126. Hazra S, et al.: Post-translational phosphorylation of serine 74 of human deoxycytidine kinase favors the enzyme adopting the open conformation making it competent for nucleoside binding and release. Biochemistry 2011,50(14):2870-2880. 10.1021/bi2001032

    CAS  PubMed  PubMed Central  Google Scholar 

  127. Neschadim A, et al.: Cell fate control gene therapy based on engineered variants of human deoxycytidine kinase. Mol Ther 2012,20(5):1002-1013. 10.1038/mt.2011.298

    CAS  PubMed  PubMed Central  Google Scholar 

  128. Hazra S, et al.: Extending thymidine kinase activity to the catalytic repertoire of human deoxycytidine kinase. Biochemistry 2009,48(6):1256-1263. 10.1021/bi802062w

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgement

We thank Dr. Wei Xie for preparing the literature. This work was supported by NIH grants R01CA133093 and R01ES016354 to Bo Xu.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Michael W Lee or Bo Xu.

Additional information

Competing interests

Southern Research Institute holds the patent on clofarabine and has received royalties from its sale. William B. Parker has financially benefitted from these payments. Bo Xu has a US Patent Application (US2008/0262003A1) related to clofarabine.

Authors’ contributions

BX and MWL organized the structure of the manuscript. ML, WBP and BX wrote the manuscript. All authors read and approved the final manuscript.

Authors’ original submitted files for images

Below are the links to the authors’ original submitted files for images.

Authors’ original file for figure 1

Rights and permissions

This article is published under license to BioMed Central Ltd. This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Lee, M.W., Parker, W.B. & Xu, B. New insights into the synergism of nucleoside analogs with radiotherapy. Radiat Oncol 8, 223 (2013). https://doi.org/10.1186/1748-717X-8-223

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1748-717X-8-223

Keywords