Selective anti-tumor activity of the novel fluoropyrimidine polymer F10 towards G48a orthotopic GBM tumors

William H Gmeiner, Carla Lema-Tome, Denise Gibo, Jamie Jennings-Gee, Carol Milligan, Waldemar Debinski, William H Gmeiner, Carla Lema-Tome, Denise Gibo, Jamie Jennings-Gee, Carol Milligan, Waldemar Debinski

Abstract

F10 is a novel anti-tumor agent with minimal systemic toxicity in vivo and which displays strong cytotoxicity towards glioblastoma (GBM) cells in vitro. Here we investigate the cytotoxicity of F10 towards GBM cells and evaluate the anti-tumor activity of locally-administered F10 towards an orthotopic xenograft model of GBM. The effects of F10 on thymidylate synthase (TS) inhibition and Topoisomerase 1 (Top1) cleavage complex formation were evaluated using TS activity assays and in vivo complex of enzyme bioassays. Cytotoxicity of F10 towards normal brain was evaluated using cortices from embryonic (day 18) mice. F10 displays minimal penetrance of the blood-brain barrier and was delivered by intra-cerebral (i.c.) administration and prospective anti-tumor response towards luciferase-expressing G48a human GBM tumors in nude mice was evaluated using IVIS imaging. Histological examination of tumor and normal brain tissue was used to assess the selectivity of anti-tumor activity. F10 is cytotoxic towards G48a, SNB-19, and U-251 MG GBM cells through dual targeting of TS and Top1. F10 is not toxic to murine primary neuronal cultures. F10 is well-tolerated upon i.c. administration and induces significant regression of G48a tumors that is dose-dependent. Histological analysis from F10-treated mice revealed tumors were essentially completely eradicated in F10-treated mice while vehicle-treated mice displayed substantial infiltration into normal tissue. F10 displays strong efficacy for GBM treatment with minimal toxicity upon i.c. administration establishing F10 as a promising drug-candidate for treating GBM in human patients.

Figures

Fig. 1
Fig. 1
F10 is cytotoxic to GBM cells through induction of thymineless death. a Chemical structure of F10; b Metabolic conversion of F10 produces FdUMP and initiates Topoisomerase 1-mediated DNA double strand breaks; c Cytotoxicity of F10 to G48 cells; d TS inhibitory activity of F10 towards SNB-19 (top), G48a (middle), and U-251 (bottom) cells at 10−6 (asterisk) and 10−5 M (double asterisk) and for raltitrexed at 10−6 M. e Western blots (for three independent samples) evaluating TS expression in G48a, U-251, and SNB-19 cells relative to HL60 cells that are sensitive to F10 at nM concentrations. Overall sensitivity to F10 inversely correlates with TS expression
Fig. 2
Fig. 2
F10 is not toxic to primary neuronal cultures and does not damage normal brain upon i.c. administration. a Viability of primary neuronal cells grown in tissue culture following treatment with F10 or 5-FU at the indicated doses. 5-FU, but not F10, significantly decreased neuronal survival at 1 μM relative to control (p < 0.05). ANOVA followed by Newman–Keuls multiple comparison test: control versus 1 μM 5-FU; p ≤ 0.05; 1 μm 5-FU versus 1 μM F10; p ≤ 0.05. b H&E stained section from the brain of mice treated with F10 at 120 mg/kg
Fig. 3
Fig. 3
F10 treatment results in significant and dose-responsive regression of G48a orthotopic xenografts. a Luciferase signal from F10- and vehicle-treated nude mice at 80 and 120 mg kg doses. Treatment with F10 results in marked decrease in luciferase-signal for treated mice. b Mean tumor luminescence calculated from the luciferase images. F10 treatment results in regression of G48a xenografts that is highly significant (p < 0.01) relative to vehicle for the 120 mg/kg treatment
Fig. 4
Fig. 4
F10 treatment results in selective eradication of G48a orthotopic brain tumors in nude mice. H&E staining from brain sections obtained from nude mice bearing G48a xenografts following treatment with vehicle (a, d, g) or F10 at 80 (b, e, h) or 120 mg/kg (c, f, i). F10, or vehicle, was administered i.c. over 7 days using an osmotic pump. Brain tissue from mice treated with vehicle-only revealed extensive infiltration of GBM cells into non-malignant tissue and a significant tumor mass (arrows point to tumor borders). Treatment with F10 at either 80 or 120 mg/kg resulted in extensive necrosis selectively for malignant cells with no apparent damage for non-malignant cells in the contralateral side at the level of the hippocampus. Invasive island of cells were observed in contralateral hippocampus of vehicle-treated animals (arrows in g) (Scale bar 100 μm in c and i, 50 μm in f)
Fig. 5
Fig. 5
F10 treatment results in selective eradication of EphA2-stained G48a cells. EphA2 staining from brain sections obtained from nude mice bearing G48a xenografts following treatment with vehicle only (a, d) or F10 at 80 (b, e) or 120 mg/kg (c, f). Strong EphA2 staining is observed for tumor tissue in vehicle-only treated mice with no EphA2 staining in adjacent non-malignant tissue except in isolated invasive cells (arrows indicate tumor borders and point to invasive cells away from tumor core). EphA2 staining is greatly diminished in region of the residual tumor (indicated by arrows) from mice treated with F10 at 80 mg/kg and is absent in mice treated with F10 at 120 mg/kg (Scale bar 50 μm)

References

    1. Krex D, Klink B, Hartmann C, von Deimling A, Pietsch T, Simon M, Sabel M, Steinbach JP, Heese O, Reifenberger G, Weller M, Schackert G. Long-term survival with glioblastoma multiforme. Brain. 2007;130(Pt 10):2596–2606. doi: 10.1093/brain/awm204.
    1. Hulleman E, Helin K. Molecular mechanisms in gliomagenesis. Adv Cancer Res. 2005;94:1–27. doi: 10.1016/S0065-230X(05)94001-3.
    1. Henriksson R, Asklund T, Poulsen HS. Impact of therapy on quality of life, neurocognitive function and their correlates in glioblastoma multiforme: a review. J Neurooncol. 2011;104(3):639–646. doi: 10.1007/s11060-011-0565-x.
    1. Zuber J, Radtke I, Pardee TS, Zhao Z, Rappaport AR, Luo W, McCurrach ME, Yang MM, Dolan ME, Kogan SC, Downing JR, Lowe SW. Mouse models of human AML accurately predict chemotherapy response. Genes Dev. 2009;23(7):877–889. doi: 10.1101/gad.1771409.
    1. Pardee TS, Gomes E, Jennings-Gee J, Caudell D, Gmeiner WH. Unique dual targeting of thymidylate synthase and topoisomerase 1 by FdUMP[10] results in high efficacy against AML and low toxicity. Blood. 2012;119:3561–3570. doi: 10.1182/blood-2011-06-362442.
    1. Pardee T, Jennings-Gee J, Stadelman K, Caudell D, Gmeiner WH (2013) The poison oligonucleotide F10 is efficiently taken up by, and highly effective against, acute lymphoblastic leukemia while sparing normal hematopoietic cells. Blood 122:2674
    1. Feldman EJ. Too much ara-C? Not enough daunorubicin? Blood. 2011;117(8):2299–2300. doi: 10.1182/blood-2011-01-328633.
    1. Gmeiner WH, Reinhold WC, Pommier Y. Genome-wide mRNA and microRNA profiling of the NCI 60 cell-line screen and comparison of FdUMP[10] with fluorouracil, floxuridine, and topoisomerase 1 poisons. Mol Cancer Ther. 2010;9(12):3105–3114. doi: 10.1158/1535-7163.MCT-10-0674.
    1. Debinski W, Gibo DM. Fos-related antigen 1 modulates malignant features of glioma cells. Mol Cancer Res. 2005;3(4):237–249.
    1. Wykosky J, Gibo DM, Debinski W. A novel, potent, and specific ephrinA1-based cytotoxin against EphA2 receptor expressing tumor cells. Mol Cancer Ther. 2007;6(12 Pt 1):3208–3218. doi: 10.1158/1535-7163.MCT-07-0200.
    1. Liao ZY, Sordet O, Zhang HL, Kohlhagen G, Antony S, Gmeiner WH, Pommier Y. A novel polypyrimidine antitumor agent FdUMP[10] induces thymineless death with topoisomerase I-DNA complexes. Cancer Res. 2005;65(11):4844–4851. doi: 10.1158/0008-5472.CAN-04-1302.
    1. Gmeiner WH, Trump E, Wei C. Enhanced DNA-directed effects of FdUMP[10] compared to 5FU. Nucleosides Nucleotides Nucleic Acids. 2004;23(1–2):401–410. doi: 10.1081/NCN-120028336.
    1. Wang W, McLeod HL, Cassidy J, Collie-Duguid ES. Mechanisms of acquired chemoresistance to 5-fluorouracil and tomudex: thymidylate synthase dependent and independent networks. Cancer Chemother Pharmacol. 2007;59(6):839–845. doi: 10.1007/s00280-006-0384-5.
    1. Sharpe R, Pearson A, Herrera-Abreu MT, Johnson D, Mackay A, Welti JC, Natrajan R, Reynolds AR, Reis-Filho JS, Ashworth A, Turner NC. FGFR signaling promotes the growth of triple-negative and basal-like breast cancer cell lines both in vitro and in vivo. Clin Cancer Res. 2011;17(16):5275–5286. doi: 10.1158/1078-0432.CCR-10-2727.
    1. Mroz P, Bhaumik J, Dogutan DK, Aly Z, Kamal Z, Khalid L, Kee HL, Bocian DF, Holten D, Lindsey JS, Hamblin MR. Imidazole metalloporphyrins as photosensitizers for photodynamic therapy: role of molecular charge, central metal and hydroxyl radical production. Cancer Lett. 2009;282(1):63–76. doi: 10.1016/j.canlet.2009.02.054.
    1. Hockfield S, Carlson S, Carol Evans P, Pintar LJ, Silberstein L. Molecular probes of the nervous system. Plainview: Cold Spring Harbor Laboratory Press; 1993. Primary cultures from the central nervous system; pp. 34–55.
    1. Graham-Cole CL, Thomas HD, Taylor GA, Newell DR, Melton RG, Hesp R, Boddy AV. An evaluation of thymidine phosphorylase as a means of preventing thymidine rescue from the thymidylate synthase inhibitor raltitrexed. Cancer Chemother Pharmacol. 2007;59(2):197–206. doi: 10.1007/s00280-006-0258-x.
    1. Han R, Yang YM, Dietrich J, Luebke A, Mayer-Proschel M, Noble M. Systemic 5-fluorouracil treatment causes a syndrome of delayed myelin destruction in the central nervous system. J Biol. 2008;7(4):12. doi: 10.1186/jbiol69.
    1. Wykosky J, Debinski W. The EphA2 receptor and ephrinA1 ligand in solid tumors: function and therapeutic targeting. Mol Cancer Res. 2008;6(12):1795–1806. doi: 10.1158/1541-7786.MCR-08-0244.
    1. Debinski W, Tatter S. Convection-enhanced delivery for the treatment of brain tumors. Expert Rev Neurother. 2009;9(10):1519–1527. doi: 10.1586/ern.09.99.
    1. Debinski W, Tatter S. Convection-enhanced delivery to achieve widespread distribution of viral vectors: predicting clinical implementation. Curr Opin Mol Ther. 2010;12(6):647–653.
    1. Mueller S, Poller MY, Lee B, Kunwer S, Pedain C, Wembacher-Schroder E, Mittermeyer S, Westphal M, Sampson JH, Vogelbaum MA, Croteau D, Chang SM. Effects of imaging and catheter characteristics on clinical outcomes for patients in the PRECISE study. J Neurooncol. 2011;101(2):267–277. doi: 10.1007/s11060-010-0255-0.
    1. Pritchard DM, Watson AJ, Potten CS, Jackman AL, Hickman JA. Inhibition by uridine but not thymidine of p53-dependent intestinal apoptosis initiated by 5-fluorouracil: evidence for the involvement of RNA perturbation. Proc Natl Acad Sci USA. 1997;94(5):1795–1799. doi: 10.1073/pnas.94.5.1795.
    1. Yamashita K, Yada H, Ariyoshi T. Neurotoxic effects of alpha-fluoro-beta-alanine (FBAL) and fluoroacetic acid (FA) on dogs. J Toxicol Sci. 2004;29(2):155–166. doi: 10.2131/jts.29.155.
    1. Miller CR, Williams CR, Buchsbaum DJ, Gillespie GY. Intratumoral 5-fluorouracil produced by cytosine deaminase/5-fluorocytosine gene therapy is effective for experimental human glioblastomas. Cancer Res. 2002;62(3):773–780.

Source: PubMed

3
Abonnieren