Subcellular distribution of the anticancer drug mitoxantrone in human and drug-resistant murine cells analyzed by flow cytometry and confocal microscopy and its relationship to the induction of DNA damage.

Flow cytometry and laser scanning confocal imaging have been used to analyze the uptake of the anticancer topoisomerase II poison mitoxantrone by intact mammalian cells and the results correlated with the induction of DNA damage. Unlike Adriamycin, mitoxantrone displays only minimal levels of red fluorescence when excited at 514 wavelength. However, using these excitation and emission conditions, flow cytometry could detect low levels of fluorescence in human transformed fibroblasts exposed to high concentrations (5-20 microM) of mitoxantrone for 1 h. Over this dose range whole cell fluorescence was a function of cell size and increased with drug concentration while drug-induced DNA-protein cross-linking showed saturation. Confocal microscopy revealed the time- and dose-dependent appearance of fluorescence, interpreted here as reflecting the disposition of drug molecules, preferentially within the cytoplasm, nuclear membrane, and nucleoli. This pattern contrasted with the intense intranuclear fluorescence observed in Adriamycin-treated human cells. Loss of the nuclear membrane during mitosis resulted in an apparent increase in chromatin-associated fluorescence. Photon counting procedures revealed a predominantly cytoplasmic, possibly lysosomal, location for fluorescence from human cells exposed for 1 h to a low but cytotoxic concentration (0.1 microM, yielding approximately 90% cell kill) of mitoxantrone. At this low concentration, human cells displayed minimal levels of DNA strand cleavage or DNA-protein cross-linking. Murine cells, displaying mitoxantrone resistance as part of the P-glycoprotein-mediated multidrug resistance phenotype, showed specific extinction of mitoxantrone-associated fluorescence from inside nuclei but not from within extranuclear compartments. The study demonstrates the feasibility of high resolution studies on the intracellular distribution of mitoxantrone in intact living cells. We suggest a mechanism by which cytoplasmic sequestration of mitoxantrone may be important in determining the response of normal and multidrug-resistant cells as they attempt to progress through mitosis.

[1]  J. Gervasoni,et al.  Subcellular distribution of daunorubicin in P-glycoprotein-positive and -negative drug-resistant cell lines using laser-assisted confocal microscopy. , 1991, Cancer research.

[2]  C. Hanstock,et al.  Interactions of the antitumor agents mitoxantrone and bisantrene with deoxyribonucleic acids studied by electron microscopy. , 1984, Molecular pharmacology.

[3]  L. Liu,et al.  DNA topoisomerase poisons as antitumor drugs. , 1989, Annual review of biochemistry.

[4]  J. Endicott,et al.  The biochemistry of P-glycoprotein-mediated multidrug resistance. , 1989, Annual review of biochemistry.

[5]  L. Liu,et al.  DNA damage by antitumor acridines mediated by mammalian DNA topoisomerase II. , 1986, Cancer research.

[6]  W. Earnshaw,et al.  Differential expression of DNA topoisomerases I and II during the eukaryotic cell cycle. , 1988, Proceedings of the National Academy of Sciences of the United States of America.

[7]  M. Melamed,et al.  Interactions of a new antitumor agent, 1,4-dihydroxy-5,8-bis[[2-[(2-hydroxyethyl)amino]-ethyl]amino]-9,10-anthracenedione, with nucleic acids. , 1981, Biochemical pharmacology.

[8]  S. Crooke,et al.  Studies on the fluorescence labeling of human red blood cell membrane ghosts with 4'-(9-acridinylamino)methanesulfon-m-anisidide. , 1985, Biochemical pharmacology.

[9]  S. Yanovich,et al.  Differences in daunomycin retention in sensitive and resistant P388 leukemic cells as determined by digitized video fluorescence microscopy. , 1983, Cancer research.

[10]  C. W. Lin,et al.  Lysosomal localization and mechanism of uptake of Nile blue photosensitizers in tumor cells. , 1991, Cancer research.

[11]  R. Epstein,et al.  Estrogen-induced potentiation of DNA damage and cytotoxicity in human breast cancer cells treated with topoisomerase II-interactive antitumor drugs. , 1988, Cancer research.

[12]  D. V. Von Hoff,et al.  Mitoxantrone: a new anticancer drug with significant clinical activity. , 1986, Annals of internal medicine.

[13]  H. Preisler Alteration of binding of the supravital dye Hoechst 33342 to human leukemic cells by adriamycin. , 1978, Cancer treatment reports.

[14]  P. Smith,et al.  Mitoxantrone-DNA binding and the induction of topoisomerase II associated DNA damage in multi-drug resistant small cell lung cancer cells. , 1990, Biochemical pharmacology.

[15]  Sutton Sm,et al.  DNA sequence preferences for the anti-cancer drug mitoxanthrone and related anthraquinones revealed by DNase I footprinting , 1986 .

[16]  A. Gottlieb Therapy of adult acute lymphocytic leukemia. , 1984, Seminars in oncology.

[17]  M. Fordham,et al.  An evaluation of confocal versus conventional imaging of biological structures by fluorescence light microscopy , 1987, The Journal of cell biology.

[18]  J. R. Brown,et al.  Evidence for human liver mediated free-radical formation by doxorubicin and mitozantrone. , 1985, Anti-cancer drug design.

[19]  Z. Darżynkiewicz,et al.  Relationship between the pharmacological activity of antitumor drugs Ametantrone and mitoxantrone (Novatrone) and their ability to condense nucleic acids. , 1986, Proceedings of the National Academy of Sciences of the United States of America.

[20]  G. Koch,et al.  Chemosensitisation by verapamil and cyclosporin A in mouse tumour cells expressing different levels of P-glycoprotein and CP22 (sorcin). , 1990, British Journal of Cancer.

[21]  R. Wallace,et al.  Development of resistance and characteristics of a human colon carcinoma subline resistant to mitoxantrone in vitro. , 1987, Cancer investigation.

[22]  D. Henry,et al.  Experimental antitumor activity of aminoanthraquinones. , 1979, Cancer treatment reports.

[23]  L. Liu,et al.  Nonintercalative antitumor drugs interfere with the breakage-reunion reaction of mammalian DNA topoisomerase II. , 1984, The Journal of biological chemistry.

[24]  P. Smith,et al.  Reduced nuclear binding of a DNA minor groove ligand (Hoechst 33342) and its impact on cytotoxicity in drug resistant murine cell lines. , 1990, British Journal of Cancer.

[25]  M. Waring,et al.  DNA sequence preferences for the anti‐cancer drug mitoxanthrone and related anthraquinones revealed by DNase I footprinting , 1986, FEBS letters.

[26]  J. Watson A linear transform of the multi-target survival curve. , 1978, The British journal of radiology.

[27]  P. Tofilon,et al.  Chromatin modifications associated with N-methylformamide-induced radiosensitization of clone A cells. , 1988, Cancer research.

[28]  P. Rabbitts,et al.  Collateral resistance to verapamil in multidrug-resistant mouse tumor cells. , 1989, Journal of the National Cancer Institute.

[29]  D. Alberts,et al.  Comparative molecular pharmacology in leukemic L1210 cells of the anthracene anticancer drugs mitoxantrone and bisantrene. , 1985, Cancer research.

[30]  N. Bachur,et al.  Cytofluorescence localization of adriamycin and daunorubicin. , 1974, Cancer Research.

[31]  C. Hanstock,et al.  High field 1H-NMR analysis of the 1:1 intercalation complex of the antitumor agent mitoxantrone and the DNA duplex [d(CpGpCpG)]. , 1985, Journal of biomolecular structure & dynamics.

[32]  W. Dalton,et al.  Persistent intracellular binding of mitoxantrone in a human colon carcinoma cell line. , 1989, Biochemical pharmacology.

[33]  P. Smith,et al.  Cellular consequences of overproduction of DNA topoisomerase II in an ataxia-telangiectasia cell line. , 1989, Cancer research.

[34]  M. Melamed,et al.  Selective displacement of nuclear proteins by antitumor drugs having affinity for nucleic acids. , 1989, Proceedings of the National Academy of Sciences of the United States of America.

[35]  N. Bleehen,et al.  The in vitro effects and cross-resistance patterns of some novel anthracyclines. , 1986, British Journal of Cancer.

[36]  P. Workman,et al.  Improved cellular accumulation is characteristic of anthracyclines which retain high activity in multidrug resistant cell lines, alone or in combination with verapamil or cyclosporin A. , 1989, Biochemical pharmacology.

[37]  Z. Darżynkiewicz,et al.  Interactions of antitumor agents Ametantrone and Mitoxantrone (Novatrone) with double-stranded DNA. , 1985, Biochemical pharmacology.

[38]  P. Smith,et al.  Long-term inhibition of DNA synthesis and the persistence of trapped topoisomerase II complexes in determining the toxicity of the antitumor DNA intercalators mAMSA and mitoxantrone. , 1990, Cancer research.

[39]  M. Dietel,et al.  Membrane vesicle formation due to acquired mitoxantrone resistance in human gastric carcinoma cell line EPG85-257. , 1990, Cancer research.