Low-dose decitabine-based chemoimmunotherapy for patients with refractory advanced solid tumors: a phase I/II report

Hui Fan, Xuechun Lu, Xiaohui Wang, Yang Liu, Bo Guo, Yan Zhang, Wenying Zhang, Jing Nie, Kaichao Feng, Meixia Chen, Yajing Zhang, Yao Wang, Fengxia Shi, Xiaobing Fu, Hongli Zhu, Weidong Han, Hui Fan, Xuechun Lu, Xiaohui Wang, Yang Liu, Bo Guo, Yan Zhang, Wenying Zhang, Jing Nie, Kaichao Feng, Meixia Chen, Yajing Zhang, Yao Wang, Fengxia Shi, Xiaobing Fu, Hongli Zhu, Weidong Han

Abstract

Aberrant DNA methylation is one of the main drivers of tumor initiation and progression. The reversibility of methylation modulation makes it an attractive target for novel anticancer therapies. Clinical studies have demonstrated that high-dose decitabine, a hypomethylating agent, results in some clinical benefits in patients with refractory advanced tumors; however, they are extremely toxic. Low doses of decitabine minimize toxicity while potentially improving the targeted effects of DNA hypomethylation. Based on these mechanisms, low-dose decitabine combined with chemoimmunotherapy may be a new treatment option for patients with refractory advanced tumors. We proposed the regimen of low-dose decitabine-based chemoimmunotherapy for patients with refractory advanced solid tumors. A favorable adverse event profile was observed in our trial that was highlighted by the finding that most of these adverse events were grades 1-2. Besides, the activity of our cohort was optimistic and the clinical benefit rate was up to 60%, and the median PFS was prolonged compared with PFS to previous treatment. We also identified a significant correlation between the PFS to previous treatment and clinical response. The low-dose DAC decitabine-based chemoimmunotherapy might be a promising protocol for improving the specificity and efficiency of patients with refractory advanced solid tumors. This trial is registered in the ClinicalTrials.gov database (identifier NCT01799083).

Figures

Figure 1
Figure 1
Study design. Chemo: chemotherapy; CIK: cytokine induced killer cells.
Figure 2
Figure 2
Trial profile. (∗) Reasons for not meeting the inclusion criteria: patient had a massive hemorrhage (n = 1). CIK: cytokine induced killer cells. DAC + Chemo: DAC combined with chemotherapy group; DAC + CIK: DAC combined with CIK group.
Figure 3
Figure 3
CT image for UPN 20 exhibited a partial response. A CT image showing the specific characteristics of the response of lung lesion following four cycles of decitabine treatment, assessing by the Response Evaluation Criteria in Solid Tumors (RECIST). The red arrows indicate the areas of measurable disease.
Figure 4
Figure 4
Swim plot showing the increase in progression-free survival (PFS) compared with the patients' previous therapies. The bars represent the progression-free survivals (PFSs) of the DAC (a) and DAC combined with chemo (b) or CIK (c) versus the PFSs following the patients' previous therapies. Four patients had died by the first assessment (due to disease progression) and were therefore not evaluated. ∗ The patients' progression-free survivals (PFSs) were significantly prolonged compared with those of previous therapies; P value <0.05 was considered significant. CIK: cytokine induced killer cells. # Patients with disease progression. ▲ Patients who died.
Figure 5
Figure 5
In vivo and in vitro biological activities of decitabine (DAC) were shown in human hepatocellular carcinoma HepG2 cell line and peripheral blood mononuclear cells (PBMCs). (a) Quantitative RT-PCR analyses of the mRNA levels of RASSF1A, MAGEA-3, MAGEA-1, p16, p15, and BRCA1 expression on human hepatocellular carcinoma HepG2 cell line; the cell lines were treated with 10 nM DAC for 72 h before collecting mRNA for analysis. Compared to untreated control, the mRNA expression levels of MAGEA-3, MAGEA-1, p16, and p15 were augmented significantly in treated cell line. (b, c) Quantitative RT-PCR analyses of the mRNA levels of MAGEA-3, MAGEA-1, p16, and p15 expression in peripheral blood mononuclear cells (PBMCs) from patients who exhibited prolonged disease stabilization following low-dose DAC treatment for the first cycle. Progressive increases in the mRNA expressions of MAGEA-3, MAGEA-1, p16, and p15 were observed in patient UNP 25; in contrast, the p16 and p15 mRNA expressions were reduced in patient UNP 14. (d) Methylation-specific PCR analyses of the changes in MAGEA-1 promoter methylation levels in peripheral blood mononuclear cells (PBMCs) collected from patients UNP 25 and UNP 14, during the first treatment cycle. The levels of MAGEA-1 promoter methylation of patients UNP 14 and UNP 25 were reduced and the levels of MAGEA-1 promoter unmethylation were increased at the same time. M: methylation; U: unmethylation. *P < 0.05, for the significance of the gene expressions differences between the DAC treatment sample and the pre-DAC sample. Error bars represent standard deviation of the measurements.

References

    1. Siegel R, Naishadham D, Jemal A. Cancer statistics, 2012. CA Cancer Journal for Clinicians. 2012;62(1):10–29.
    1. Longley DB, Johnston PG. Molecular mechanisms of drug resistance. Journal of Pathology. 2005;205(2):275–292.
    1. Berger SL, Kouzarides T, Shiekhattar R, Shilatifard A. An operational definition of epigenetics. Genes and Development. 2009;23(7):781–783.
    1. Sharma S, Kelly TK, Jones PA. Epigenetics in cancer. Carcinogenesis. 2009;31(1):27–36.
    1. Taby R, Issa J-P. Cancer epigenetics. CA Cancer Journal for Clinicians. 2010;60(6):376–392.
    1. Gomes A, Reis-Silva M, Alarcao A. Promoter hypermethylation of DNA repair genes MLH1 and MSH2 in adenocarcinomas and squamous cell carcinomas of the lung. Revista Portuguesa de Pneumologia. 2014;20(1):20–30.
    1. Peedicayil J. The role of DNA methylation in the pathogenesis and treatment of cancer. Current Clinical Pharmacology. 2012;7(4):333–340.
    1. Sang M, Wu X, Fan X, et al. Multiple MAGE-A genes as surveillance marker for the detection of circulating tumor cells in patients with ovarian cancer. Biomarkers. 2014;19(1):34–42.
    1. Gurion R, Vidal L, Gafter-Gvili A, Yeshurun YB, Raanani P, Shpilberg O. 5-Azacitidine prolongs overall survival in patients with myelodysplastic syndrome—a systematic review and meta-analysis. Haematologica. 2010;95(2):303–310.
    1. Cashen AF, Schiller GJ, O’Donnell MR, DiPersio JF. Multicenter, phase II study of decitabine for the first-line treatment of older patients with Acute Myeloid Leukemia. Journal of Clinical Oncology. 2010;28(4):556–561.
    1. Gnyszka A, Jastrzebski Z, Flis S. DNA methyltransferase inhibitors and their emerging role in epigenetic therapy of cancer. Anticancer Research. 2013;33(8):2989–2996.
    1. Kelly TK, de Carvalho DD, Jones PA. Epigenetic modifications as therapeutic targets. Nature Biotechnology. 2010;28(10):1069–1078.
    1. Blum W, Klisovic RB, Hackanson B, et al. Phase I study of decitabine alone or in combination with valproic acid in acute myeloid leukemia. Journal of Clinical Oncology. 2007;25(25):3884–3891.
    1. Wang L-X, Mei Z-Y, Zhou J-H, et al. Low dose decitabine treatment induces CD80 expression in cancer cells and stimulates tumor specific cytotoxic T Lymphocyte Responses. PLoS ONE. 2013;8(5)e62924
    1. Wijermans P, Lübbert M, Verhoef G, et al. Low-dose 5-Aza-2′-deoxycytidine, a DNA hypomethylating agent, for the treatment of high-risk myelodysplastic syndrome: a multicenter phase II study in elderly patients. Journal of Clinical Oncology. 2000;18(5):956–962.
    1. Kantarjian H, Oki Y, Garcia-Manero G, et al. Results of a randomized study of 3 schedules of low-dose decitabine in higher-risk myelodysplastic syndrome and chronic myelomonocytic leukemia. Blood. 2007;109(1):52–57.
    1. Chu BF, Karpenko MJ, Liu Z, et al. Phase I study of 5-aza-2′-deoxycytidine in combination with valproic acid in non-small-cell lung cancer. Cancer Chemotherapy and Pharmacology. 2013;71(1):115–121.
    1. Kopp LM, Ray A, Denman CJ, et al. Decitabine has a biphasic effect on natural killer cell viability, phenotype, and function under proliferative conditions. Molecular Immunology. 2013;54(3-4):296–301.
    1. Zhang Y, Wang J, Wang Y, et al. Autologous CIK cell immunotherapy in patients with renal cell carcinoma after radical nephrectomy. Clinical and Developmental Immunology. 2013;2013:12 pages.195691
    1. Esteller M, Corn PG, Baylin SB, Herman JG. A gene hypermethylation profile of human cancer. Cancer Research. 2001;61(8):3225–3229.
    1. Lotem J, Sachs L. Epigenetics and the plasticity of differentiation in normal and cancer stem cells. Oncogene. 2006;25(59):7663–7672.
    1. Hirata H, Hinoda Y, Nakajima K, et al. Wnt antagonist DKK1 acts as a tumor suppressor gene that induces apoptosis and inhibits proliferation in human renal cell carcinoma. International Journal of Cancer. 2011;128(8):1793–1803.
    1. Chun SG, Zhou W, Yee NS. Combined targeting of histone deacetylases and hedgehog signaling enhances cytoxicity in pancreatic cancer. Cancer Biology & Therapy. 2009;8(14):1328–1339.
    1. Lundqvist A, Abrams SI, Schrump DS, et al. Bortezomib and depsipeptide sensitize tumors to tumor necrosis factor-related apoptosis-inducing ligand: a novel method to potentiate natural killer cell tumor cytotoxicity. Cancer Research. 2006;66(14):7317–7325.
    1. Fonsatti E, Nicolay HJM, Sigalotti L, et al. Functional up-regulation of human leukocyte antigen class I antigens expression by 5-aza-2′-deoxycytidine in cutaneous melanoma: immunotherapeutic implications. Clinical Cancer Research. 2007;13(11):3333–3338.
    1. Samlowski WE, Leachman SA, Wade M, et al. Evaluation of a 7-day continuous intravenous infusion of decitabine: inhibition of promoter-specific and global genomic DNA methylation. Journal of Clinical Oncology. 2005;23(17):3897–3905.
    1. Vigil CE, Martin-Santos T, Garcia-Manero G. Safety and efficacy of azacitidine in myelodysplastic syndromes. Drug Design, Development and Therapy. 2010;4:221–229.
    1. Stathis A, Hotte SJ, Chen EX, et al. Phase I study of decitabine in combination with vorinostat in patients with advanced solid tumors and non-Hodgkin’s lymphomas. Clinical Cancer Research. 2011;17(6):1582–1590.
    1. George RE, Lahti JM, Adamson PC, et al. Phase I study of decitabine with doxorubicin and cyclophosphamide in children with neuroblastoma and other solid tumors: a children’s oncology group study. Pediatric Blood and Cancer. 2010;55(4):629–638.
    1. Appleton K, Mackay HJ, Judson I, et al. Phase I and pharmacodynamic trial of the DNA methyltransferase inhibitor decitabine and carboplatin in solid tumors. Journal of Clinical Oncology. 2007;25(29):4603–4609.
    1. Yoo CB, Jones PA. Epigenetic therapy of cancer: past, present and future. Nature Reviews Drug Discovery. 2006;5(1):37–50.
    1. Issa J-PJ. DNA methylation as a therapeutic target in cancer. Clinical Cancer Research. 2007;13(6):1634–1637.
    1. Jabbour E, Issa J-P, Garcia-Manero G, Kantarjian H. Evolution of decitabine development: accomplishments, ongoing investigations, and future strategies. Cancer. 2008;112(11):2341–2351.
    1. Karahoca M, Momparler RL. Pharmacokinetic and pharmacodynamic analysis of 5-aza-2′-deoxycytidine (decitabine) in the design of its dose-schedule for cancer therapy. Clin Epigenetics. 2013;5(1, article 3)
    1. Bauman J, Verschraegen C, Belinsky S, et al. A phase i study of 5-azacytidine and erlotinib in advanced solid tumor malignancies. Cancer Chemotherapy and Pharmacology. 2012;69(2):547–554.
    1. Schrump DS, Fischette MR, Nguyen DM, et al. Phase I study of decitabine-mediated gene expression in patients with cancers involving the lungs, esophagus, or pleura. Clinical Cancer Research. 2006;12(19):5777–5785.
    1. Krishnadas DK, Bao L, Bai F, et al. Decitabine facilitates immune recognition of sarcoma cells by upregulating CT antigens, MHC molecules, and ICAM-1. Tumor Biology. 2014
    1. Santini V, Kantarjian HM, Issa J-P. Changes in DNA methylation in neoplasia: pathophysiology and therapeutic implications. Annals of Internal Medicine. 2001;134(7):573–586.

Source: PubMed

3
Subskrybuj