Review

High-mobility group box 1 is a promising diagnostic and therapeutic monitoring biomarker in Cancers: A review

  • Received: 20 July 2018 Accepted: 26 October 2018 Published: 30 November 2018
  • High-mobility group box 1 (HMGB1) is universally found in the nuclei and cytosols of mammalian cells. Excessive secretion of HMGB1 is concomitant with unrestricted replicative capabilities, angiogenesis, apoptosis, self-reliance in growth signals, neutrality to growth inhibitors, inflammation, tissue invasion and metastasis which are the cardinal indicators of cancer. HMGB1 levels are markedly elevated in cancers like colon, breast, lung, prostate, cervical, gastric, hepatocellular, leukemia, lymphoma and melanoma as compare to normal controls. Currently, clinicians rely on history, examination and radiological findings to diagnose as well as follow-up patients on treatment. This orthodox method has led to a “diagnostic gap” in several cancers. Therefore, this “diagnostic gap” could be filled with circulating biomarkers. Biomarkers can effortlessly be obtained via venipuncture, cost-effective and permit an appraisal of current transformations in cancer or therapy response via serial testing. This review focuses on the potential biomarker role of HMGB1 in predicting the prognosis as well as monitor therapy in various human cancers.

    Citation: Seidu A. Richard. High-mobility group box 1 is a promising diagnostic and therapeutic monitoring biomarker in Cancers: A review[J]. AIMS Molecular Science, 2018, 5(4): 183-241. doi: 10.3934/molsci.2018.4.183

    Related Papers:

  • High-mobility group box 1 (HMGB1) is universally found in the nuclei and cytosols of mammalian cells. Excessive secretion of HMGB1 is concomitant with unrestricted replicative capabilities, angiogenesis, apoptosis, self-reliance in growth signals, neutrality to growth inhibitors, inflammation, tissue invasion and metastasis which are the cardinal indicators of cancer. HMGB1 levels are markedly elevated in cancers like colon, breast, lung, prostate, cervical, gastric, hepatocellular, leukemia, lymphoma and melanoma as compare to normal controls. Currently, clinicians rely on history, examination and radiological findings to diagnose as well as follow-up patients on treatment. This orthodox method has led to a “diagnostic gap” in several cancers. Therefore, this “diagnostic gap” could be filled with circulating biomarkers. Biomarkers can effortlessly be obtained via venipuncture, cost-effective and permit an appraisal of current transformations in cancer or therapy response via serial testing. This review focuses on the potential biomarker role of HMGB1 in predicting the prognosis as well as monitor therapy in various human cancers.


    加载中
    [1] Mcguire S (2016) World cancer report 2014. Geneva, Switzerland: World Health Organization, international agency for research on cancer, WHO Press, 2015. Adv Nutr 7: 418–419.
    [2] Pilzweger C, Holdenrieder S (2015) Circulating HMGB1 and RAGE as clinical biomarkers in malignant and autoimmune diseases. Diagnostics 5: 219–253. doi: 10.3390/diagnostics5020219
    [3] Therasse P, Arbuck SG, Eisenhauer EA, et al. (2000) New guidelines to evaluate the response to treatment in solid tumors. J Natl Cancer I 92: 205–216. doi: 10.1093/jnci/92.3.205
    [4] Brambs HJ, Claussen C (1993) Pancreatic and ampullary carcinoma. Ultrasound, computed tomography, magnetic resonance imaging and angiography. Endoscopy 25: 58–68.
    [5] Gress F, Gottlieb K, Sherman S, et al. (2001) Endoscopic ultrasonography-guided fine-needle aspiration biopsy of suspected pancreatic cancer. Ann Intern Med 134: 459–464. doi: 10.7326/0003-4819-134-6-200103200-00010
    [6] Yang X, Liu J, Ren Y, et al. (2017) Isolated intracranial Rosai-Dorfman disease mimicking petroclival meningioma in a child: Case report and review of the literature. Medicine 96: e8754. doi: 10.1097/MD.0000000000008754
    [7] Gerlinger M, Rowan AJ, Horswell S, et al. (2012) Intratumor heterogeneity and branched evolution revealed by multiregion sequencing. New Eng J Med 366: 883–892. doi: 10.1056/NEJMoa1113205
    [8] Sturgeon C, Hill R, Hortin GL, et al. (2010) Taking a new biomarker into routine use-a perspective from the routine clinical biochemistry laboratory. Proteomics Clin Appl 4: 892–903. doi: 10.1002/prca.201000073
    [9] Haas M, Heinemann V, Kullmann F, et al. (2013) Prognostic value of CA 19–9, CEA, CRP, LDH and bilirubin levels in locally advanced and metastatic pancreatic cancer: Results from a multicenter, pooled analysis of patients receiving palliative chemotherapy. J Cancer Res Clin Oncol 139: 681–689. doi: 10.1007/s00432-012-1371-3
    [10] Behesnilian AS, Reiter RE (2015) Risk stratification of prostate cancer in the modern era. Curr Opin Urol 25: 246. doi: 10.1097/MOU.0000000000000164
    [11] Korse CM, Holdenrieder S, Zhi XY, et al. (2015) Multicenter evaluation of a new progastrin-releasing peptide (ProGRP) immunoassay across Europe and China. Clin Chim Acta 438: 388–395. doi: 10.1016/j.cca.2014.09.015
    [12] Molina R, Escudero JM, Augé JM, et al. (2011) HE4 a novel tumour marker for ovarian cancer: Comparison with CA 125 and ROMA algorithm in patients with gynaecological diseases. Tumor Biol 32: 1087. doi: 10.1007/s13277-011-0204-3
    [13] Ying S, Jiang Z, He X, et al. (2017) Serum HMGB1 as a potential biomarker for patients with asbestos-related diseases. Dis Markers 2017: 5756102.
    [14] Naglova H, Bucova M (2012) HMGB1 and its physiological and pathological roles. Bratisl Lek Listy 113: 163–171.
    [15] Richard SA, Jiang Y, Xiang LH, et al. (2017) Post-translational modifications of high mobility group box 1 and cancer. Am J Transl Res 9: 5181–5196.
    [16] Sims GP, Rowe DC, Rietdijk ST, et al. (2009) HMGB1 and RAGE in inflammation and cancer. Annu Rev Immunol 28: 367–388.
    [17] Richard SA, Min W, Su Z, et al. (2017) High Mobility Group Box 1 and Traumatic Brain Injury. J Behav Brain Sci 7: 50. doi: 10.4236/jbbs.2017.72006
    [18] Richard SA, Sackey M, Su Z, et al. (2017) Pivotal neuroinflammatory and therapeutic role of high mobility group box 1 in ischemic stroke. Biosci Rep, 37.
    [19] Richard SA, Xiang L, Yun J, et al. (2017) Carcinogenic and therapeutic role of High-Mobility Group Box 1 in Cancer: Is it a cancer facilitator, a cancer inhibitor or both? World Cancer Res J 4: e919.
    [20] Richard SA, Min W, Su Z, et al. (2017) Epochal neuroinflammatory role of high mobility group box 1 in central nervous system diseases. AIMS Mol Sci 4: 185–218. doi: 10.3934/molsci.2017.2.185
    [21] Stoetzer OJ, Fersching DM, Salat C, et al. (2013) Circulating immunogenic cell death biomarkers HMGB1 and RAGE in breast cancer patients during neoadjuvant chemotherapy. Tumor Biol 34: 81–90. doi: 10.1007/s13277-012-0513-1
    [22] Lotze MT, Tracey KJ (2005) High-mobility group box 1 protein (HMGB1): Nuclear weapon in the immune arsenal. Nat Rev Immunol 5: 331. doi: 10.1038/nri1594
    [23] Tesniere A, Panaretakis T, Kepp O, et al. (2008) Molecular characteristics of immunogenic cancer cell death. Cell Death Differ 15: 3. doi: 10.1038/sj.cdd.4402269
    [24] Bianchi ME (2009) HMGB1 loves company. J Leukocyte Biol 86: 573–576. doi: 10.1189/jlb.1008585
    [25] Park JS, Gamboni-Robertson F, He Q, et al. (2006) High mobility group box 1 protein interacts with multiple Toll-like receptors. Am J Physiol-cell Ph 290: C917–C924. doi: 10.1152/ajpcell.00401.2005
    [26] Kepp O, Tesniere A, Schlemmer F, et al. (2009) Immunogenic cell death modalities and their impact on cancer treatment. Apoptosis 14: 364–375. doi: 10.1007/s10495-008-0303-9
    [27] Apetoh L, Ghiringhelli F, Tesniere A, et al. (2007) Toll-like receptor 4-dependent contribution of the immune system to anticancer chemotherapy and radiotherapy. Nat Med 13: 1050–1059. doi: 10.1038/nm1622
    [28] Apetoh L, Tesniere A, Ghiringhelli F, et al. (2008) Molecular interactions between dying tumor cells and the innate immune system determine the efficacy of conventional anticancer therapies. Cancer Res 68: 4026–4030. doi: 10.1158/0008-5472.CAN-08-0427
    [29] Wu T, Zhang W, Yang G, et al. (2016) HMGB1 overexpression as a prognostic factor for survival in cancer: A meta-analysis and systematic review. Oncotarget 7: 50417.
    [30] Johann DJ, Mcguigan MD, Patel AR, et al. (2004) Clinical proteomics and biomarker discovery. Ann Ny Acad Sci 1022: 295–305. doi: 10.1196/annals.1318.045
    [31] Passalacqua M, Zicca A, Sparatore B, et al. (1997) Secretion and binding of HMG1 protein to the external surface of the membrane are required for murine erythroleukemia cell differentiation. Febs Lett 400: 275–279. doi: 10.1016/S0014-5793(96)01402-0
    [32] Wähämaa H, Vallerskog T, Qin S, et al. (2007) HMGB1-secreting capacity of multiple cell lineages revealed by a novel HMGB1 ELISPOT assay. J Leukocyte Biol 81: 129–136. doi: 10.1189/jlb.0506349
    [33] Dong XDE, Ito N, Lotze MT, et al. (2007) High mobility group box I (HMGB1) release from tumor cells after treatment: implications for development of targeted chemoimmunotherapy. J Immunother 30: 596–606. doi: 10.1097/CJI.0b013e31804efc76
    [34] Schiraldi M, Raucci A, Muñoz LM, et al. (2012) HMGB1 promotes recruitment of inflammatory cells to damaged tissues by forming a complex with CXCL12 and signaling via CXCR4. J Exp Med 209: 551–563. doi: 10.1084/jem.20111739
    [35] van Beijnum JR, Dings RP, van der Linden E, et al. (2006) Gene expression of tumor angiogenesis dissected: Specific targeting of colon cancer angiogenic vasculature. Blood 108: 2339–2348. doi: 10.1182/blood-2006-02-004291
    [36] van Beijnum J, Nowak-Sliwinska P, van den Boezem E, et al. (2013) Tumor angiogenesis is enforced by autocrine regulation of high-mobility group box 1. Oncogene 32: 363–374. doi: 10.1038/onc.2012.49
    [37] Taguchi A, Blood DC, del Toro G, et al. (2000) Blockade of RAGE-amphoterin signalling suppresses tumour growth and metastases. Nature 405: 354–360. doi: 10.1038/35012626
    [38] Zhao H, Guo L, Zhao H, et al. (2015) CXCR4 over-expression and survival in cancer: A system review and meta-analysis. Oncotarget 6: 5022.
    [39] Ueda M, Takahashi Y, Shinden Y, et al. (2014) Prognostic significance of high mobility group box 1 (HMGB1) expression in patients with colorectal cancer. Anticancer Res 34: 5357–5362.
    [40] Seidu RA, Wu M, Su Z, et al. (2017) Paradoxical Role of High Mobility Group Box 1 in Glioma: A Suppressor or a Promoter? Oncol Rev 11: 325.
    [41] Angelopoulou E, Piperi C, Adamopoulos C, et al. (2016) Pivotal role of high-mobility group box 1 (HMGB1) signaling pathways in glioma development and progression. J Mol Med 94: 867–874. doi: 10.1007/s00109-016-1435-y
    [42] Zhang J, Liu C, Hou R (2014) Knockdown of HMGB1 improves apoptosis and suppresses proliferation and invasion of glioma cells. Chin J Cancer Res 26: 658–668.
    [43] Seidu A, Richard YY, Hao Li, et al. (2018) Glioblastoma multiforme subterfuge as acute cerebral hemorrhage: A case report and literature review. Neurol Int 10: 7558.
    [44] Richard SA, Liang RF, Lei CF, et al. (2017) Diffuse leptomeningeal tuberculoma masquerading as leptomeningeal gliomatosis: A case report and review of literature. Infect Dis Trop Med 3: e406.
    [45] Wang X, Zhou S, Fu X, et al. (2015) Clinical and prognostic significance of high-mobility group box-1 in human gliomas. Exp Ther Med 9: 513–518. doi: 10.3892/etm.2014.2089
    [46] Stupp R, Brada M, van den Bent M, et al. (2014) High-grade glioma: ESMO Clinical Practice Guidelines for diagnosis, treatment and follow-up. Ann Oncol 25: iii93–iii101. doi: 10.1093/annonc/mdu050
    [47] Louis DN, Ohgaki H, Wiestler OD, et al. (2007) The 2007 WHO classification of tumours of the central nervous system. Acta Neuropathol 114: 97–109. doi: 10.1007/s00401-007-0243-4
    [48] Gupta P, Dixit D, Sen E (2012) Oncrasin targets the JNK-NF-κB axis to sensitize glioma cells to TNFα-induced apoptosis. Carcinogenesis 34: 388–396.
    [49] Feldkamp MM, Feldkamp MM, Lala P, et al. (1999) Expression of activated epidermal growth factor receptors, Ras-guanosine triphosphate, and mitogen-activated protein kinase in human glioblastoma multiforme specimens. Neurosurgery 45: 1442–1453. doi: 10.1097/00006123-199912000-00034
    [50] Norris JL, Baldwin AS (1999) Oncogenic Ras enhances NF-κB transcriptional activity through Raf-dependent and Raf-independent mitogen-activated protein kinase signaling pathways. J Biol Chem 274: 13841–13846. doi: 10.1074/jbc.274.20.13841
    [51] Lee H, Herrmann A, Deng JH, et al. (2009) Persistently activated Stat3 maintains constitutive NF-κB activity in tumors. Cancer Cell 15: 283–293. doi: 10.1016/j.ccr.2009.02.015
    [52] Guo W, Wu S, Liu J, et al. (2008) Identification of a small molecule with synthetic lethality for K-ras and protein kinase C iota. Cancer Res 68: 7403–7408. doi: 10.1158/0008-5472.CAN-08-1449
    [53] Kennedy NJ, Sluss HK, Jones SN, et al. (2003) Suppression of Ras-stimulated transformation by the JNK signal transduction pathway. Gene Dev 17: 629–637. doi: 10.1101/gad.1062903
    [54] Turner KM, Sun Y, Ji P, et al. (2015) Genomically amplified Akt3 activates DNA repair pathway and promotes glioma progression. P Natl Acad Sci Usa 112: 3421–3426. doi: 10.1073/pnas.1414573112
    [55] Altomare DA, Testa JR (2005) Perturbations of the AKT signaling pathway in human cancer. Oncogene 24: 7455. doi: 10.1038/sj.onc.1209085
    [56] Wang H, Wang H, Zhang W, et al. (2004) Analysis of the activation status of Akt, NF [kappa] B, and Stat3 in human diffuse gliomas. Lab Invest 84: 941. doi: 10.1038/labinvest.3700123
    [57] Wang LH, Cheng GZ, Park S, et al. (2008) Advances of AKT pathway in human oncogenesis and as a target for anti-cancer drug discovery. Curr Cancer Drug Targets 8: 2–6. doi: 10.2174/156800908783497159
    [58] Dillon RL, Muller WJ (2010) Distinct biological roles for the akt family in mammary tumor progression. Cancer Res 70: 4260–4264.
    [59] Maroulakou IG, Oemler W, Naber SP, et al. (2007) Akt1 ablation inhibits, whereas Akt2 ablation accelerates, the development of mammary adenocarcinomas in mouse mammary tumor virus (MMTV)-ErbB2/neu and MMTV-polyoma middle T transgenic mice. Cancer Res 67: 167–177.
    [60] Pan B, Chen D, Huang J, et al. (2014) HMGB1-mediated autophagy promotes docetaxel resistance in human lung adenocarcinoma. Mol Cancer 13: 135.
    [61] Tang D, Kang R, Livesey KM, et al. (2010) Endogenous HMGB1 regulates autophagy. J Cell Biol 190: 881–892. doi: 10.1083/jcb.200911078
    [62] Mizushima N, Yoshimori T, Levine B (2010) Methods in mammalian autophagy research. Cell 140: 313–326. doi: 10.1016/j.cell.2010.01.028
    [63] He C, Klionsky DJ (2009) Regulation mechanisms and signaling pathways of autophagy. Annu Rev Genet 43: 67. doi: 10.1146/annurev-genet-102808-114910
    [64] Curtin JF, Liu N, Candolfi M, et al. (2009) HMGB1 mediates endogenous TLR2 activation and brain tumor regression. PLoS Med 6: e1000010. doi: 10.1371/journal.pmed.1000010
    [65] Singh MK, Singh L, Pushker N, et al. (2015) Correlation of high mobility group Box-1 protein (HMGB1) with clinicopathological parameters in primary retinoblastoma. Pathol Oncol Res 21: 1237–1242. doi: 10.1007/s12253-015-9951-6
    [66] Jiao Y (2007) Growth suppression and radiosensitivity increase by HMGB1 in breast cancer1. Acta Pharmacol Sin 28: 1957–1967. doi: 10.1111/j.1745-7254.2007.00669.x
    [67] Wang L, Feng Y, Cheng Y (2017) Effect on proliferation and apoptosis of retinoblastoma cell by RNA inhibiting high mobility group protein box-1 expression. Int J Ophthalmol 10: 30.
    [68] Meng Q, Zhao J, Liu H, et al. (2014) HMGB1 promotes cellular proliferation and invasion, suppresses cellular apoptosis in osteosarcoma. Tumor Biol 35: 12265–12274. doi: 10.1007/s13277-014-2535-3
    [69] Wang L, Meng Q, Jiao Y, et al. (2012) High-mobility group boxes mediate cell proliferation and radiosensitivity via retinoblastoma-interaction-dependent and-independent mechanisms. Cancer Biother Radio 27: 329–335.
    [70] Liu K, Huang J, Xie M, et al. (2014) MIR34A regulates autophagy and apoptosis by targeting HMGB1 in the retinoblastoma cell. Autophagy 10: 442–452. doi: 10.4161/auto.27418
    [71] Bartel DP (2004) MicroRNAs: genomics, biogenesis, mechanism, and function. Cell 116: 281–297. doi: 10.1016/S0092-8674(04)00045-5
    [72] Li Q, Li J, Wen T, et al. (2014) Overexpression of HMGB1 in melanoma predicts patient survival and suppression of HMGB1 induces cell cycle arrest and senescence in association with p21 (Waf1/Cip1) up-regulation via a p53-independent, Sp1-dependent pathway. Oncotarget 5: 6387.
    [73] Huber R, Meier B, Otsuka A, et al. (2016) Tumour hypoxia promotes melanoma growth and metastasis via High Mobility Group Box-1 and M2-like macrophages. Sci Rep 6: 29914. doi: 10.1038/srep29914
    [74] Bald T, Quast T, Landsberg J, et al. (2014) Ultraviolet-radiation-induced inflammation promotes angiotropism and metastasis in melanoma. Nature 507: 109. doi: 10.1038/nature13111
    [75] Parker KH, Sinha P, Horn LA, et al. (2014) HMGB1 enhances immune suppression by facilitating the differentiation and suppressive activity of myeloid-derived suppressor cells. Cancer Res 74: 5723–5733. doi: 10.1158/0008-5472.CAN-13-2347
    [76] Kurte M, López M, Aguirre A, et al. (2004) A synthetic peptide homologous to functional domain of human IL-10 down-regulates expression of MHC class I and transporter associated with antigen processing 1/2 in human melanoma cells. J Immunol 173: 1731–1737. doi: 10.4049/jimmunol.173.3.1731
    [77] Carter NA, Rosser EC, Mauri C (2012) Interleukin-10 produced by B cells is crucial for the suppression of Th17/Th1 responses, induction of T regulatory type 1 cells and reduction of collagen-induced arthritis. Arthritis Res Ther 14: R32. doi: 10.1186/ar3736
    [78] Carter NA, Vasconcellos R, Rosser EC, et al. (2011) Mice lacking endogenous IL-10-producing regulatory B cells develop exacerbated disease and present with an increased frequency of Th1/Th17 but a decrease in regulatory T cells. J Immunol 186: 5569–5579. doi: 10.4049/jimmunol.1100284
    [79] Chen Q, Daniel V, Maher DW, et al. (1994) Production of IL-10 by melanoma cells: Examination of its role in immunosuppression mediated by melanoma. Int J Cancer 56: 755–760. doi: 10.1002/ijc.2910560524
    [80] Tang Q, Li J, Zhu H, et al. (2013) Hmgb1-IL-23-IL-17-IL-6-Stat3 axis promotes tumor growth in murine models of melanoma. Mediat Inflamm 2013: 713859.
    [81] Zhao M, Yang M, Yang L, et al. (2011) HMGB1 regulates autophagy through increasing transcriptional activities of JNK and ERK in human myeloid leukemia cells. Bmb Rep 44: 601–606. doi: 10.5483/BMBRep.2011.44.9.601
    [82] Liu L, Yang M, Kang R, et al. (2011) HMGB1-induced autophagy promotes chemotherapy resistance in leukemia cells. Leukemia 25: 23. doi: 10.1038/leu.2010.225
    [83] Kang R, Livesey KM, Zeh I, et al. (2010) HMGB1: A novel Beclin 1-binding protein active in autophagy. Autophagy 6: 1209–1211. doi: 10.4161/auto.6.8.13651
    [84] Jia L, Clear A, Liu FT, et al. (2014) Extracellular HMGB1 promotes differentiation of nurse-like cells in chronic lymphocytic leukemia. Blood 123: 1709–1719. doi: 10.1182/blood-2013-10-529610
    [85] Zhao T, Ren H, Wang X, et al. (2015) Rituximab-induced HMGB1 release is associated with inhibition of STAT3 activity in human diffuse large B-cell lymphoma. Oncotarget 6: 27816.
    [86] Apetoh L, Ghiringhelli F, Tesniere A, et al. (2007) The interaction between HMGB1 and TLR4 dictates the outcome of anticancer chemotherapy and radiotherapy. Immunol Rev 220: 47–59. doi: 10.1111/j.1600-065X.2007.00573.x
    [87] Sterenczak KA, Joetzke AE, Willenbrock S, et al. (2010) High-mobility group B1 (HMGB1) and receptor for advanced glycation end-products (RAGE) expression in canine lymphoma. Anticancer Res 30: 5043–5048.
    [88] Nieters A, Beckmann L, Deeg E, et al. (2006) Gene polymorphisms in Toll-like receptors, interleukin-10, and interleukin-10 receptor alpha and lymphoma risk. Genes Immun 7: 615. doi: 10.1038/sj.gene.6364337
    [89] Mao X, Wang G, Chen Z, et al. (2012) Expression of HMGB1 and its Clinical Significance in T-cell Lymphoma. Asian Pac J Cancer Prev 13: 5569–5571. doi: 10.7314/APJCP.2012.13.11.5569
    [90] Wang H, Yang H, Tracey K (2004) Extracellular role of HMGB1 in inflammation and sepsis. J Intern Med 255: 320–331. doi: 10.1111/j.1365-2796.2003.01302.x
    [91] Sohun M, Shen H (2016) The implication and potential applications of high-mobility group box 1 protein in breast cancer. Annals Transl Med 4.
    [92] Wang R, He W, Li Z, et al. (2014) Caveolin-1 functions as a key regulator of 17β-estradiol-mediated autophagy and apoptosis in BT474 breast cancer cells. Int J Mol Med 34: 822–827. doi: 10.3892/ijmm.2014.1836
    [93] He Q, Liang CH, Lippard SJ (2000) Steroid hormones induce HMG1 overexpression and sensitize breast cancer cells to cisplatin and carboplatin. Proc Nat Acad Sci 97: 5768–5772. doi: 10.1073/pnas.100108697
    [94] Su Z, Ni P, She P, et al. (2017) Bio-HMGB1 from breast cancer contributes to M-MDSC differentiation from bone marrow progenitor cells and facilitates conversion of monocytes into MDSC-like cells. Cancer Immunology, Immunotherapy 66: 391–401. doi: 10.1007/s00262-016-1942-2
    [95] Zhu X, Sun L, Wang Y (2016) High mobility group box 1 (HMGB1) is upregulated by the Epstein-Barr virus infection and promotes the proliferation of human nasopharyngeal carcinoma cells. Acta Oto-laryngol 136: 87–94. doi: 10.3109/00016489.2015.1082192
    [96] Wu D, Ding Y, Wang S, et al. (2008) Increased expression of high mobility group box 1 (HMGB1) is associated with progression and poor prognosis in human nasopharyngeal carcinoma. Journal Pathol 216: 167–175. doi: 10.1002/path.2391
    [97] Kuniyasu H, Chihara Y, Takahashi T (2003) Co-expression of receptor for advanced glycation end products and the ligand amphoterin associates closely with metastasis of colorectal cancer. Oncol Rep 10: 445–448.
    [98] Kuniyasu H, Oue N, Wakikawa A, et al. (2002) Expression of receptors for advanced glycation end-products (RAGE) is closely associated with the invasive and metastatic activity of gastric cancer. J Pathol 196: 163–170.
    [99] Huttunen HJ, Fages C, Kuja-Panula J, et al. (2002) Receptor for advanced glycation end products-binding COOH-terminal motif of amphoterin inhibits invasive migration and metastasis. Cancer Res 62: 4805–4811.
    [100] Ding J, Cui X, Liu Q (2016) Emerging role of HMGB1 in lung diseases: Friend or foe. J Cell Mol Med.
    [101] Campana L, Bosurgi L, Rovere-Querini P (2008) HMGB1: a two-headed signal regulating tumor progression and immunity. Curr Opin Immunol 20: 518–523. doi: 10.1016/j.coi.2008.04.012
    [102] Kusume A, Sasahira T, Luo Y, et al. (2009) Suppression of dendritic cells by HMGB1 is associated with lymph node metastasis of human colon cancer. Pathobiol 76: 155–162. doi: 10.1159/000218331
    [103] He Y, Zha J, Wang Y, et al. (2012) Tissue damage-associated 'danger signals' influence T cell responses that promote the progression of pre-neoplasia to cancer. Cancer Res 73: 629–639.
    [104] Wild CA, Brandau S, Lotfi R, et al. (2012) HMGB1 is overexpressed in tumor cells and promotes activity of regulatory T cells in patients with head and neck cancer. Oral Oncol 48: 409–416. doi: 10.1016/j.oraloncology.2011.12.009
    [105] Kang R, Zhang Q, Zeh HJ, et al. (2013) HMGB1 in cancer: good, bad, or both? Clin Cancer Res 19: 4046–4057. doi: 10.1158/1078-0432.CCR-13-0495
    [106] Kang R, Tang D, Schapiro NE, et al. (2010) The receptor for advanced glycation end products (RAGE) sustains autophagy and limits apoptosis, promoting pancreatic tumor cell survival. Cell Death Differ 17: 666–676. doi: 10.1038/cdd.2009.149
    [107] Kang R, Chen R, Zhang Q, et al. (2014) HMGB1 in health and disease. Mol Aspects Med 40: 1–116.
    [108] Štros M, Polanska E, Štruncová S, et al. (2009) HMGB1 and HMGB2 proteins up-regulate cellular expression of human topoisomerase IIα. Nucleic Acids Res 37: 2070–2086. doi: 10.1093/nar/gkp067
    [109] Chuangui C, Peng T, Zhentao Y (2012) The expression of high mobility group box 1 is associated with lymph node metastasis and poor prognosis in esophageal squamous cell carcinoma. Pathol Oncol Res 18: 1021–1027. doi: 10.1007/s12253-012-9539-3
    [110] Obeid M, Tesniere A, Ghiringhelli F, et al. (2007) Calreticulin exposure dictates the immunogenicity of cancer cell death. Nat Med 13: 54. doi: 10.1038/nm1523
    [111] Suzuki Y, Mimura K, Yoshimoto Y, et al. (2012) Immunogenic tumor cell death induced by chemoradiotherapy in patients with esophageal squamous cell carcinoma. Cancer Res 72: 3967–3976. doi: 10.1158/0008-5472.CAN-12-0851
    [112] Zhang J, Zhang R, Lu W, et al. (2014) Clinical significance of hmgb1 expression in human gastric cancer. Int J Immunopath Ph 27: 543–551. doi: 10.1177/039463201402700410
    [113] Nehil M, Paquette J, Tokuyasu T, et al. (2014) High mobility group box 1 promotes tumor cell migration through epigenetic silencing of semaphorin 3A. Oncogene 33: 5151. doi: 10.1038/onc.2013.459
    [114] Abe A, Kuwata T, Yamauchi C, et al. (2014) High Mobility Group Box1 (HMGB1) released from cancer cells induces the expression of pro-inflammatory cytokines in peritoneal fibroblasts. Pathol Int 64: 267–275. doi: 10.1111/pin.12167
    [115] Zhang Q, Wu L, Zhang T, et al. (2015) Autophagy-mediated HMGB1 release promotes gastric cancer cell survival via RAGE activation of extracellular signal-regulated kinases 1/2. Oncol Rep 33: 1630–1638. doi: 10.3892/or.2015.3782
    [116] Cheng P, Dai W, Wang F, et al. (2014) Ethyl pyruvate inhibits proliferation and induces apoptosis of hepatocellular carcinoma via regulation of the HMGB1–RAGE and AKT pathways. Biochem Bioph Res Co 443: 1162–1168. doi: 10.1016/j.bbrc.2013.12.064
    [117] Yan H, Wu H, Zhang H, et al. (2013) p53 promotes inflammation-associated hepatocarcinogenesis by inducing HMGB1 release. J Hepatol 59: 762–768. doi: 10.1016/j.jhep.2013.05.029
    [118] Zhang L, Han J, Wu H, et al. (2014) The association of HMGB1 expression with clinicopathological significance and prognosis in hepatocellular carcinoma: A meta-analysis and literature review. PloS one 9: e110626. doi: 10.1371/journal.pone.0110626
    [119] Takada M, Hirata K, Ajiki T, et al. (2004) Expression of receptor for advanced glycation end products (RAGE) and MMP-9 in human pancreatic cancer cells. Hepato-Gastroenterol 51: 928–930.
    [120] Kang R, Livesey KM, Zeh I, et al. (2011) Metabolic regulation by HMGB1-mediated autophagy and mitophagy. Autophagy 7: 1256–1258. doi: 10.4161/auto.7.10.16753
    [121] Kang R, Loux T, Tang D, et al. (2012) The expression of the receptor for advanced glycation endproducts (RAGE) is permissive for early pancreatic neoplasia. Proc Nat Acad Sci 109: 7031–7036.
    [122] Cebrian MJG, Bauden M, Andersson R, et al. (2016) Paradoxical Role of HMGB1 in Pancreatic Cancer: Tumor Suppressor or Tumor Promoter? Anticancer Res 36: 4381–4389. doi: 10.21873/anticanres.10981
    [123] Kang R, Tang D, Loze MT, et al. (2011) Apoptosis to autophagy switch triggered by the MHC class III-encoded receptor for advanced glycation endproducts (RAGE). Autophagy 7: 91–93. doi: 10.4161/auto.7.1.13852
    [124] Arumugam T, Ramachandran V, Gomez SB, et al. (2012) S100P-derived RAGE antagonistic peptide reduces tumor growth and metastasis. Clin Cancer Res 18: 4356–4364. doi: 10.1158/1078-0432.CCR-12-0221
    [125] Li J, Sun J, Rong R, et al. (2017) HMGB1 promotes myeloid-derived suppressor cells and renal cell carcinoma immune escape. Oncotarget 8: 63290–63298.
    [126] Messmer D, Yang H, Telusma G, et al. (2004) High mobility group box protein 1: An endogenous signal for dendritic cell maturation and Th1 polarization. J Immunol 173: 307–313. doi: 10.4049/jimmunol.173.1.307
    [127] Wagner KU, Schmidt JW (2011) The two faces of Janus kinases and their respective STATs in mammary gland development and cancer. J Carcinog 10: 32.
    [128] Gnanasekar M, Kalyanasundaram R, Zheng G, et al. (2013) HMGB1: A promising therapeutic target for prostate cancer. Prostate Cancer 2013: 157103.
    [129] Ishiguro H, Nakaigawa N, Miyoshi Y, et al. (2005) Receptor for advanced glycation end products (RAGE) and its ligand, amphoterin are overexpressed and associated with prostate cancer development. Prostate 64: 92–100. doi: 10.1002/pros.20219
    [130] Elangovan I, Thirugnanam S, Chen A, et al. (2012) Targeting receptor for advanced glycation end products (RAGE) expression induces apoptosis and inhibits prostate tumor growth. Biochem Bioph Res Co 417: 1133–1138. doi: 10.1016/j.bbrc.2011.12.060
    [131] Shiota M, Izumi H, Miyamoto N, et al. (2008) Ets regulates peroxiredoxin1 and 5 expressions through their interaction with the high‐mobility group protein B1. Cancer Sci 99: 1950–1959.
    [132] He Y, Zha J, Wang Y, et al. (2013) Tissue Damage-Associated "Danger Signals" Influence T-cell Responses That Promote the Progression of Preneoplasia to Cancer. Cancer Res 73: 629–639. doi: 10.1158/0008-5472.CAN-12-2704
    [133] Huang Z, Zhong Z, Zhang L, et al. (2015) Down-regulation of HMGB1 expression by shRNA constructs inhibits the bioactivity of urothelial carcinoma cell lines via the NF-κB pathway. Sci Rep 5: 12807. doi: 10.1038/srep12807
    [134] Musumeci D, Roviello GN, Montesarchio D (2014) An overview on HMGB1 inhibitors as potential therapeutic agents in HMGB1-related pathologies. Pharmacol Therapeut 141: 347–357. doi: 10.1016/j.pharmthera.2013.11.001
    [135] Shrivastava S, Mansure JJ, Almajed W, et al. (2016) The Role of HMGB1 in radioresistance of Bladder Cancer. Mol Cancer Ther 15: 471–479. doi: 10.1158/1535-7163.MCT-15-0581
    [136] Livesey KM, Kang R, Vernon P, et al. (2012) p53/HMGB1 complexes regulate autophagy and apoptosis. Cancer Res 72: 1996–2005. doi: 10.1158/0008-5472.CAN-11-2291
    [137] Huang J, Ni J, Liu K, et al. (2012) HMGB1 promotes drug resistance in osteosarcoma. Cancer Res 72: 230–238.
    [138] Exner R, Sachet M, Arnold T, et al. (2016) Prognostic value of HMGB1 in early breast cancer patients under neoadjuvant chemotherapy. Cancer Med 5: 2350–2358. doi: 10.1002/cam4.827
    [139] Song D, Li H, Li H, et al. (2015) Effect of human papillomavirus infection on the immune system and its role in the course of cervical cancer. Oncol Lett 10: 600–606. doi: 10.3892/ol.2015.3295
    [140] Huang L, Yao Y, Zhang L, et al. (2009) The effect of high-mobility group box 1 protein on activity of regulatory T cells after thermal injury in rats. Shock 31: 322–329. doi: 10.1097/SHK.0b013e3181834070
    [141] Pang X, Zhang Y, Wei H, et al. (2014) Expression and effects of high-mobility group box 1 in cervical cancer. Int J Mol Sci 15: 8699–8712. doi: 10.3390/ijms15058699
    [142] Zhao W, Chen Q (2014) Effects of HMGB1 on proliferation and apoptosis of human brain glioma CD133 cells. Bratisl Lek Listy 116: 480–485.
    [143] Kivelä T (2009) The epidemiological challenge of the most frequent eye cancer: Retinoblastoma, an issue of birth and death. Br J Ophthalmol 93: 1129–1131. doi: 10.1136/bjo.2008.150292
    [144] Knudson AG (1971) Mutation and cancer: statistical study of retinoblastoma. Proc Nat Acad Sci 68: 820–823. doi: 10.1073/pnas.68.4.820
    [145] Comings DE (1973) A general theory of carcinogenesis. Proc Nat Acad Sci 70: 3324–3328. doi: 10.1073/pnas.70.12.3324
    [146] Chen L, Su M, Ren Sg, et al. (2014) Analysis of current status and strategies of retinopathy of prematurity screening during 6 years in local regions of China: Implication and caution. J Ophthalmol : 756059.
    [147] Gould J (2014) Gene therapy: Genie in a vector. Nature 515: S160–S161. doi: 10.1038/515S160a
    [148] Calin GA, Croce CM (2006) MicroRNA signatures in human cancers. Nat Rev Cancer 6: 857. doi: 10.1038/nrc1997
    [149] Garzon R, Calin GA, Croce CM (2009) MicroRNAs in cancer. Annu Rev Med 60: 167–179. doi: 10.1146/annurev.med.59.053006.104707
    [150] Jo DH, Kim JH, Park WY, et al. (2011) Differential profiles of microRNAs in retinoblastoma cell lines of different proliferation and adherence patterns. Am J Pediatr Hematol/Oncol 33: 529–533. doi: 10.1097/MPH.0b013e318228280a
    [151] Dalgard CL, Gonzalez M, O'Brien JM (2009) Differential microRNA-34a expression and tumor suppressor function in retinoblastoma cells. Invest Ophth Vis Sci 50: 4542–4551. doi: 10.1167/iovs.09-3520
    [152] Liu K, Huang J, Xie M, et al. (2014) MIR34A regulates autophagy and apoptosis by targeting HMGB1 in the retinoblastoma cell. Autophagy 10: 442–452. doi: 10.4161/auto.27418
    [153] Leclerc J, Ballotti R, Bertolotto C (2017) Pathways from senescence to melanoma: Focus on MITF sumoylation. Oncogene.
    [154] Nguyen AH, Detty SQ, Agrawal DK (2017) Clinical Implications of High-mobility Group Box-1 (HMGB1) and the Receptor for Advanced Glycation End-products (RAGE) in Cutaneous Malignancy: A Systematic Review. Anticancer Res 37: 1–7. doi: 10.21873/anticanres.11282
    [155] Schneider SL, Ross AL, Grichnik JM (2015) Do inflammatory pathways drive melanomagenesis? Exp Dermatol 24: 86–90. doi: 10.1111/exd.12502
    [156] Margolin K (2014) Introduction to the role of the immune system in melanoma. Hematol Oncol Clin North Am 28: 537–558. doi: 10.1016/j.hoc.2014.02.005
    [157] Lee N, Zakka LR, Mihm MC, et al. (2016) Tumour-infiltrating lymphocytes in melanoma prognosis and cancer immunotherapy. Pathology 48: 177–187. doi: 10.1016/j.pathol.2015.12.006
    [158] Sherry HY, Bordeaux JS, Baron ED (2014) The immune system and skin cancer. Sunlight, Vitamin D Skin Cancer: 182–191.
    [159] Carucci JA (2013) Understanding dendritic cells and their role in cutaneous carcinoma and cancer immunotherapy. Clin Dev Immunol 2013: 624123.
    [160] Xiong Y, Hannon GJ, Zhang H, et al. (1991) p21 is a universal inhibitor of cyclin kinases. Ried 49: 370.
    [161] El-Deiry WS, Tokino T, Velculescu VE, et al. (1993) WAF1, a potential mediator of p53 tumor suppression. Cell 75: 817–825. doi: 10.1016/0092-8674(93)90500-P
    [162] Harper JW, Elledge SJ, Keyomarsi K, et al. (1995) Inhibition of cyclin-dependent kinases by p21. Mol Biol Cell 6: 387–400. doi: 10.1091/mbc.6.4.387
    [163] Warfel NA, El-Deiry WS (2013) p21WAF1 and tumourigenesis: 20 years after. Curr Opin Oncol 25: 52–58. doi: 10.1097/CCO.0b013e32835b639e
    [164] Datto MB, Li Y, Panus JF, et al. (1995) Transforming growth factor beta induces the cyclin-dependent kinase inhibitor p21 through a p53-independent mechanism. Proc Natl Acad Sci 92: 5545–5549. doi: 10.1073/pnas.92.12.5545
    [165] De Siervi A, Marinissen M, Diggs J, et al. (2004) Transcriptional Activation of p21waf1/cip1 by Alkylphospholipids. Cancer Res 64: 743–750. doi: 10.1158/0008-5472.CAN-03-2505
    [166] Berg JM (1992) Sp1 and the subfamily of zinc finger proteins with guanine-rich binding sites. Proc Natl Acad Sci 89: 11109–11110. doi: 10.1073/pnas.89.23.11109
    [167] Liu H, Kodumudi K, Weber A, et al. (2015) Intralesional Rose Bengal in melanoma elicits tumor immunity via HMGB1. J Immunother Cancer 3: P408. doi: 10.1186/2051-1426-3-S2-P408
    [168] Liu Z, Falo LD, You Z (2011) Knockdown of HMGB1 in tumor cells attenuates their ability to induce regulatory T cells and uncovers naturally acquired CD8 T cell-dependent antitumor immunity. J Immunol 187: 118–125. doi: 10.4049/jimmunol.1003378
    [169] Leclerc E, Heizmann CW, Vetter SW (2009) RAGE and S100 protein transcription levels are highly variable in human melanoma tumors and cells. Gen Physiol Biophys 28: F65–F75.
    [170] Wagner NB, Weide B, Reith M, et al. (2015) Diminished levels of the soluble form of RAGE are related to poor survival in malignant melanoma. Int J Cancer 137: 2607–2617. doi: 10.1002/ijc.29619
    [171] Ottaviani G, Jaffe N (2009) The epidemiology of osteosarcoma. Pediatr Adolesc Osteosarcoma: 3–13.
    [172] Li X, Wang S, Chen Y, et al. (2014) miR-22 targets the 3'UTR of HMGB1 and inhibits the HMGB1-associated autophagy in osteosarcoma cells during chemotherapy. Tumor Biol 35: 6021–6028. doi: 10.1007/s13277-014-1797-0
    [173] Zwaga T, Bovée JV, Kroon HM (2008) Osteosarcoma of the femur with skip, lymph node, and lung metastases. Radiographics 28: 277–283. doi: 10.1148/rg.281075015
    [174] Li B, Jiang S, Zheng X, et al. (2013) Expression of the inflammatory molecule hmgb1 in human osteosarcoma and its clinical relevance. Eur J Inflammation 11: 61–73. doi: 10.1177/1721727X1301100107
    [175] Chano T, Matsumoto K, Ishizawa M, et al. (1996) Analysis of the presence of osteocalcin, S-100 protein, and proliferating cell nuclear antigen in cells of various types of osteosarcomas. Eur J Histochem: EJH 40: 189–198.
    [176] Loukopoulos P, O'brien T, Ghoddusi M, et al. (2004) Characterisation of three novel canine osteosarcoma cell lines producing high levels of matrix metalloproteinases. Res Vet Sci 77: 131–141. doi: 10.1016/j.rvsc.2004.01.006
    [177] Cawston TE, Wilson AJ (2006) Understanding the role of tissue degrading enzymes and their inhibitors in development and disease. Best Practice Res Clin Rheumatol 20: 983–1002. doi: 10.1016/j.berh.2006.06.007
    [178] Hotfilder M, Sondermann P, Senss A, et al. (2005) PI3K/AKT is involved in mediating survival signals that rescue Ewing tumour cells from fibroblast growth factor 2-induced cell death. Brit J Cancer 92: 705. doi: 10.1038/sj.bjc.6602384
    [179] Tsubaki M, Yamazoe Y, Yanae M, et al. (2011) Blockade of the Ras/MEK/ERK and Ras/PI3K/Akt pathways by statins reduces the expression of bFGF, HGF, and TGF-β as angiogenic factors in mouse osteosarcoma. Cytokine 54: 100–107. doi: 10.1016/j.cyto.2011.01.005
    [180] Liu B, Shi Zl, Feng J, et al. (2008) Celecoxib, a cyclooxygenase-2 inhibitor, induces apoptosis in human osteosarcoma cell line MG-63 via down-regulation of PI3K/Akt. Cell Biol Int 32: 494–501. doi: 10.1016/j.cellbi.2007.10.008
    [181] Wu Z, Min L, Chen D, et al. (2011) Overexpression of BMI-1 promotes cell growth and resistance to cisplatin treatment in osteosarcoma. PloS One 6: e14648. doi: 10.1371/journal.pone.0014648
    [182] Richard SA, Zheng S, Su Z, et al. (2017) The Pivotal Neuroinflammatory, Therapeutic and Neuroprotective Role of Alpha-Mangostin. J Neurol Res 7: 67–79. doi: 10.14740/jnr455w
    [183] Charoonpatrapong K, Shah R, Robling AG, et al. (2006) HMGB1 expression and release by bone cells. J Cell Physiol 207: 480–490. doi: 10.1002/jcp.20577
    [184] Huang J, Liu K, Yu Y, et al. (2012) Targeting HMGB1-mediated autophagy as a novel therapeutic strategy for osteosarcoma. Autophagy: 8: 275–277. doi: 10.4161/auto.8.2.18940
    [185] Zhou Z, Han JY, Xi CX, et al. (2008) HMGB1 regulates RANKL‐induced osteoclastogenesis in a manner dependent on RAGE. J Bone Miner Res 23: 1084–1096. doi: 10.1359/jbmr.080234
    [186] Martinotti S, Patrone M, Manfredi M, et al. (2016) HMGB1 Osteo‐Modulatory Action on Osteosarcoma SaOS‐2 Cell Line: An Integrated Study From Biochemical and‐Omics Approaches. J Cell Biochem 117: 2559–2569. doi: 10.1002/jcb.25549
    [187] Kataoka K, Ono T, Murata H, et al. (2012) S100A7 promotes the migration and invasion of osteosarcoma cells via the receptor for advanced glycation end products. Oncology Lett 3: 1149–1153. doi: 10.3892/ol.2012.612
    [188] Yang D, Liu G, Wang K (2015) miR-203 acts as a tumor suppressor gene in osteosarcoma by regulating RAB22A. PloS One 10: e0132225. doi: 10.1371/journal.pone.0132225
    [189] Liu K, Huang J, Ni J, et al. (2017) MALAT1 promotes osteosarcoma development by regulation of HMGB1 via miR-142–3p and miR-129–5p. Cell Cycle 16: 578–587. doi: 10.1080/15384101.2017.1288324
    [190] Sehgal A, Konig H, Johnson D, et al. (2015) You eat what you are: Autophagy inhibition as a therapeutic strategy in leukemia. Leukemia 29: 517. doi: 10.1038/leu.2014.349
    [191] Burnett AK, Goldstone A, Hills RK, et al. (2013) Curability of patients with acute myeloid leukemia who did not undergo transplantation in first remission. J Clin Oncol 31: 1293–1301. doi: 10.1200/JCO.2011.40.5977
    [192] Döhner H, Estey EH, Amadori S, et al. (2010) Diagnosis and management of acute myeloid leukemia in adults: recommendations from an international expert panel, on behalf of the European LeukemiaNet. Blood 115: 453–474. doi: 10.1182/blood-2009-07-235358
    [193] Steinbach D, Legrand O (2007) ABC transporters and drug resistance in leukemia: Was P-gp nothing but the first head of the Hydra? Leukemia 21: 1172. doi: 10.1038/sj.leu.2404692
    [194] Leith CP, Kopecky KJ, Chen IM, et al. (1999) Frequency and clinical significance of the expression of the multidrug resistance proteins MDR1/P-glycoprotein, MRP1, and LRP in acute myeloid leukemia. A Southwest Oncology Group Study. Blood 94: 1086–1099.
    [195] Chen X, Wang Y, Liu J, et al. (2015) Synergistic effect of HMGB1 knockdown and cordycepin in the K562 human chronic myeloid leukemia cell line. Mol Med Rep 12: 4462–4468. doi: 10.3892/mmr.2015.3928
    [196] Druker BJ (2002) STI571 (Gleevec™) as a paradigm for cancer therapy. Trends Mol Med 8: S14–S18. doi: 10.1016/S1471-4914(02)02305-5
    [197] Baran Y, Salas A, Senkal CE, et al. (2007) Alterations of ceramide/sphingosine 1-phosphate rheostat involved in the regulation of resistance to imatinib-induced apoptosis in K562 human chronic myeloid leukemia cells. J Biol Chem 282: 10922–10934. doi: 10.1074/jbc.M610157200
    [198] Cagnetta A, Garuti A, Marani C, et al. (2013) Evaluating treatment response of chronic myeloid leukemia: Emerging science and technology. Curr Cancer Drug Targets 13: 779–790. doi: 10.2174/15680096113139990084
    [199] Ji J, Wang HS, Gao YY, et al. (2014) Synergistic anti-tumor effect of KLF4 and curcumin in human gastric carcinoma cell line. Asian Pac J Cancer Prev 15: 7747–7752. doi: 10.7314/APJCP.2014.15.18.7747
    [200] Xie M, Kang R, Yu Y, et al. (2008) Enhancive effect of HMGB1 gene silence on adriamycin-induced apoptosis in K562/A02 drug resistance leukemia cells. ZhongHua Xue Ye Xue Za Zhi 29: 549–552.
    [201] Yu Y, Xie M, He YL, et al. (2008) Role of high mobility group box 1 in adriamycin-induced apoptosis in leukemia K562 cells. Aizheng 27: 929–933.
    [202] Baris D, Zahm SH (2000) Epidemiology of lymphomas. Curr Opin Oncol 12: 383–394. doi: 10.1097/00001622-200009000-00002
    [203] Li J, Lei C, Seidu A. Richard (2017) Giant solitary primary intracranial lymphoma masquerading as meningioma: a case and review of literature. Pan Afr Med J 28.
    [204] Martinez‐Delgado B, Robledo M, Arranz E, et al. (1997) Correlation between mutations in p53 gene and protein expression in human lymphomas. Am J Hematol 55: 1–8. doi: 10.1002/(SICI)1096-8652(199705)55:1<1::AID-AJH1>3.0.CO;2-Y
    [205] Gaidano G, Ballerini P, Gong JZ, et al. (1991) p53 mutations in human lymphoid malignancies: association with Burkitt lymphoma and chronic lymphocytic leukemia. Proc Natl Acad Sci 88: 5413–5417. doi: 10.1073/pnas.88.12.5413
    [206] Doglioni C, Pelosio P, Mombello A, et al. (1991) Immunohistochemical evidence of abnormal expression of the antioncogene-encoded p53 phosphoprotein in Hodgkin's disease and CD30+ anaplastic lymphomas. Hematol Pathol 5: 67–73.
    [207] Li C-C, O'connell C, Beckwith M, et al. (1995) Detection of p53 mutations in B cell non-Hodgkin's lymphoma cell lines. Leukemia 9: 650–655.
    [208] Sakashita A, Hattori T, Miller C, et al. (1992) Mutations of the p53 gene in adult T-cell leukemia. Blood 79: 477–480.
    [209] Ott G, Rosenwald A, Campo E (2013) Understanding MYC-driven aggressive B-cell lymphomas: pathogenesis and classification. Blood 122: 3884–3891. doi: 10.1182/blood-2013-05-498329
    [210] Klapproth K, Wirth T (2010) Advances in the understanding of MYC‐induced lymphomagenesis. Brit J Haematol 149: 484–497. doi: 10.1111/j.1365-2141.2010.08159.x
    [211] Dang CV (2012) MYC on the path to cancer. Cell 149: 22–35. doi: 10.1016/j.cell.2012.03.003
    [212] Navarro A, Beà S, Fernández V, et al. (2009) MicroRNA expression, chromosomal alterations, and immunoglobulin variable heavy chain hypermutations in Mantle cell lymphomas. Cancer Res 69: 7071–7078. doi: 10.1158/0008-5472.CAN-09-1095
    [213] Laurent C, Do C, Gascoyne RD, et al. (2009) Anaplastic lymphoma kinase–positive diffuse large B-cell lymphoma: a rare clinicopathologic entity with poor prognosis. J Clin Oncol 27: 4211–4216. doi: 10.1200/JCO.2008.21.5020
    [214] Valera A, Colomo L, Martínez A, et al. (2013) ALK-positive large B-cell lymphomas express a terminal B-cell differentiation program and activated STAT3 but lack MYC rearrangements. Modern Pathol 26: 1329. doi: 10.1038/modpathol.2013.73
    [215] Valera A, Balagué O, Colomo L, et al. (2010) IG/MYC rearrangements are the main cytogenetic alteration in plasmablastic lymphomas. The Am J Surg Pathol 34: 1686.
    [216] Salven P, Orpana A, Teerenhovi L, et al. (2000) Simultaneous elevation in the serum concentrations of the angiogenic growth factors VEGF and bFGF is an independent predictor of poor prognosis in non-Hodgkin lymphoma: A single-institution study of 200 patients. Blood 96: 3712–3718.
    [217] Meyer A, Staratschek-Jox A, Springwald A, et al. (2008) Non-Hodgkin lymphoma expressing high levels of the danger-signalling protein HMGB1. Leukemia Lymphoma 49: 1184–1189. doi: 10.1080/10428190802064909
    [218] Tang D, Kang R, Cheh C-W, et al. (2010) HMGB1 release and redox regulates autophagy and apoptosis in cancer cells. Oncogene 29: 5299–5310. doi: 10.1038/onc.2010.261
    [219] Yamazaki T, Hannani D, Poirier-Colame V, et al. (2014) Defective immunogenic cell death of HMGB1-deficient tumors: Compensatory therapy with TLR4 agonists. Cell Death Differ 21: 69. doi: 10.1038/cdd.2013.72
    [220] Sun S, Zhang W, Cui Z, et al. (2015) High mobility group box-1 and its clinical value in breast cancer. OncoTargets Ther 8: 413.
    [221] Mantovani A (2009) Cancer: Inflaming metastasis. Nature 457: 36–37.
    [222] Salto-Tellez M, James JA, Hamilton PW (2014) Molecular pathology-the value of an integrative approach. Mol Oncol 8: 1163–1168. doi: 10.1016/j.molonc.2014.07.021
    [223] Zhang Y, Jiao J, Yang P, et al. (2014) Mass spectrometry-based N-glycoproteomics for cancer biomarker discovery. Clin Proteomics 11: 18. doi: 10.1186/1559-0275-11-18
    [224] Gion M, Mione R, Leon AE, et al. (1999) Comparison of the diagnostic accuracy of CA27. 29 and CA15. 3 in primary breast cancer. Clin Chem 45: 630–637.
    [225] Hou MF, Chen YL, Tseng TF, et al. (1999) Evaluation of serum CA27.29, CA15-3 and CEA in patients with breast cancer. Kaohsiung J Med Sci 15: 520–528.
    [226] Li J, Zhang Z, Rosenzweig J, et al. (2002) Proteomics and bioinformatics approaches for identification of serum biomarkers to detect breast cancer. Clin Chem 48: 1296–1304.
    [227] dos Anjos Pultz B, da Luz FAC, de Faria PR, et al. (2014) Far beyond the usual biomarkers in breast cancer: A review. J Cancer 5: 559. doi: 10.7150/jca.8925
    [228] Oñate SA, Prendergast P, Wagner JP, et al. (1994) The DNA-bending protein HMG-1 enhances progesterone receptor binding to its target DNA sequences. Mol Cell Biol 14: 3376–3391. doi: 10.1128/MCB.14.5.3376
    [229] Verrier CS, Roodi N, Yee CJ, et al. (1997) High-mobility group (HMG) protein HMG-1 and TATA-binding protein-associated factor TAFII30 affect estrogen receptor-mediated transcriptional activation. Mol Endocrinol 11: 1009–1019. doi: 10.1210/mend.11.8.9962
    [230] Huang JC, Zamble DB, Reardon JT, et al. (1994) HMG-domain proteins specifically inhibit the repair of the major DNA adduct of the anticancer drug cisplatin by human excision nuclease. Proc Natl Acad Sci 91: 10394–10398. doi: 10.1073/pnas.91.22.10394
    [231] Burstein HJ (2005) The distinctive nature of HER2-positive breast cancers. New Engl J Med 353: 1652–1654. doi: 10.1056/NEJMp058197
    [232] Tang D, Billiar TR, Lotze MT (2012) A Janus tale of two active high mobility group box 1 (HMGB1) redox states. Mol Med 18: 1360.
    [233] Ladoire S, Penault-Llorca F, Senovilla L, et al. (2015) Combined evaluation of LC3B puncta and HMGB1 expression predicts residual risk of relapse after adjuvant chemotherapy in breast cancer. Autophagy 11: 1878–1890. doi: 10.1080/15548627.2015.1082022
    [234] Jia WH, Huang QH, Liao J, et al. (2006) Trends in incidence and mortality of nasopharyngeal carcinoma over a 20–25 year period (1978/1983–2002) in Sihui and Cangwu counties in southern China. Bmc Cancer 6: 178. doi: 10.1186/1471-2407-6-178
    [235] Lung ML, Cheung AKL, Ko JMY, et al. (2014) The interplay of host genetic factors and Epstein-Barr virus in the development of nasopharyngeal carcinoma. Chinese J Cancer 33: 556. doi: 10.5732/cjc.014.10170
    [236] Ogino T, Moriai S, Ishida Y, et al. (2007) Association of immunoescape mechanisms with Epstein‐Barr virus infection in nasopharyngeal carcinoma. Int J Cancer 120: 2401–2410. doi: 10.1002/ijc.22334
    [237] Eliopoulos AG, Young LS (1998) Activation of the cJun N-terminal kinase (JNK) pathway by the Epstein-Barr virus-encoded latent membrane protein 1 (LMP1). Oncogene 16: 1731–1742. doi: 10.1038/sj.onc.1201694
    [238] Wu L, Nakano H, Wu Z (2006) The C-terminal activating region 2 of the Epstein-Barr virus-encoded latent membrane protein 1 activates NF-κB through TRAF6 and TAK1. J Biol Chem 281: 2162–2169. doi: 10.1074/jbc.M505903200
    [239] Huttunen HJ, Fages C, Kuja-Panula J, et al. (2002) Receptor for advanced glycation end products-binding COOH-terminal motif of amphoterin inhibits invasive migration and metastasis. Cancer Res 62: 4805–4811.
    [240] Feng A, Tu Z, Yin B (2016) The effect of HMGB1 on the clinicopathological and prognostic features of non-small cell lung cancer. Oncotarget 7: 20507.
    [241] DeSantis CE, Lin CC, Mariotto AB, et al. (2014) Cancer treatment and survivorship statistics, 2014. CA: A Cancer J Clin 64: 252–271. doi: 10.3322/caac.21235
    [242] Xia Q, Xu J, Chen H, et al. (2016) association between an elevated level of hMgB1 and non-small-cell lung cancer: A meta-analysis and literature review. OncoTargets Ther 9: 3917. doi: 10.2147/OTT.S104409
    [243] Jube S, Rivera ZS, Bianchi ME, et al. (2012) Cancer cell secretion of the DAMP protein HMGB1 supports progression in malignant mesothelioma. Cancer Res 72: 3290–3301. doi: 10.1158/0008-5472.CAN-11-3481
    [244] Celona B, Weiner A, Di Felice F, et al. (2011) Substantial histone reduction modulates genomewide nucleosomal occupancy and global transcriptional output. Plos Biol 9: e1001086. doi: 10.1371/journal.pbio.1001086
    [245] Polanská E, Dobšáková Z, Dvořáčková M, et al. (2012) HMGB1 gene knockout in mouse embryonic fibroblasts results in reduced telomerase activity and telomere dysfunction. Chromosoma 121: 419–431. doi: 10.1007/s00412-012-0373-x
    [246] Lange SS, Vasquez KM (2009) HMGB1: The jack‐of‐all‐trades protein is a master DNA repair mechanic. Mol Carcinogen 48: 571–580. doi: 10.1002/mc.20544
    [247] Zhao Z, Oh S, Li D, et al. (2012) A dual role for UVRAG in maintaining chromosomal stability independent of autophagy. Dev Cell 22: 1001–1016. doi: 10.1016/j.devcel.2011.12.027
    [248] Duff SE, Li C, Jeziorska M, et al. (2003) Vascular endothelial growth factors C and D and lymphangiogenesis in gastrointestinal tract malignancy. Br J Cancer 89: 426–430. doi: 10.1038/sj.bjc.6601145
    [249] Aurello P, Rossi Del Monte S, D'Angelo F, et al. (2009) Vascular endothelial growth factor C and microvessel density in gastric carcinoma: Correlation with clinicopathological factors. Our experience and review of the literature. Oncology Res Featuring Preclin Clin Cancer Ther 17: 405–411.
    [250] SHIMODA H, TAKAHASHI Y, KAJIWARA T, et al. (2003) Demonstration of the rat lymphatic vessels using immuohistochemistry and in situ hybridization for 5'-nucleotidase. Biomed Res 24: 51–57. doi: 10.2220/biomedres.24.51
    [251] Liu P, Chen W, Zhu H, et al. (2009) Expression of VEGF-C correlates with a poor prognosis based on analysis of prognostic factors in 73 patients with esophageal squamous cell carcinomas. Jpn J Clin Oncol 39: 644–650. doi: 10.1093/jjco/hyp079
    [252] Ishikura S, Nihei K, Ohtsu A, et al. (2003) Long-term toxicity after definitive chemoradiotherapy for squamous cell carcinoma of the thoracic esophagus. J Clin Oncol 21: 2697–2702. doi: 10.1200/JCO.2003.03.055
    [253] Tuomela J, Valta M, Seppänen J, et al. (2009) Overexpression of vascular endothelial growth factor C increases growth and alters the metastatic pattern of orthotopic PC-3 prostate tumors. Bmc Cancer 9: 362. doi: 10.1186/1471-2407-9-362
    [254] Chung HW, Lee SG, Kim H, et al. (2009) Serum high mobility group box-1 (HMGB1) is closely associated with the clinical and pathologic features of gastric cancer. J Transl Med 7: 38. doi: 10.1186/1479-5876-7-38
    [255] Correa P (1992) Human gastric carcinogenesis: a multistep and multifactorial process-first American Cancer Society award lecture on cancer epidemiology and prevention. Cancer Res 52: 6735–6740.
    [256] Lin MT, Lin BR, Chang CC, et al. (2007) IL‐6 induces AGS gastric cancer cell invasion via activation of the c‐Src/RhoA/ROCK signaling pathway. Int J Cancer 120: 2600–2608. doi: 10.1002/ijc.22599
    [257] Kim DK, Oh SY, Kwon HC, et al. (2009) Clinical significances of preoperative serum interleukin-6 and C-reactive protein level in operable gastric cancer. Bmc Cancer 9: 155. doi: 10.1186/1471-2407-9-155
    [258] Ko YB, Kim BR, Nam SL, et al. (2014) High-mobility group box 1 (HMGB1) protein regulates tumor-associated cell migration through the interaction with BTB domain. Cell Signal 26: 777–783. doi: 10.1016/j.cellsig.2013.12.018
    [259] Yin Y, Li W, Deng M, et al. (2014) Extracellular high mobility group box chromosomal protein 1 promotes drug resistance by increasing the expression of P‑glycoprotein expression in gastric adenocarcinoma cells. Mol Med Rep 9: 1439–1443. doi: 10.3892/mmr.2014.1961
    [260] Xiang YY, Wang DY, Tanaka M, et al. (1997) Expression of high-mobility group-1 mRNA in human gastrointestinal adenocarcinoma and corresponding non-cancerous mucosa. Int J Cancer 74: 1–6. doi: 10.1002/(SICI)1097-0215(19970220)74:1<1::AID-IJC1>3.0.CO;2-6
    [261] Fidler IJ (1990) Critical factors in the biology of human cancer metastasis: twenty-eighth GHA Clowes memorial award lecture. Cancer Res 50: 6130–6138.
    [262] Jemal A, Bray F, Center MM, et al. (2011) Global cancer statistics. CA: A Cancer J Clin 61: 69–90. doi: 10.3322/caac.20107
    [263] Bosch FX, Ribes J, Díaz M, et al. (2004) Primary liver cancer: worldwide incidence and trends. Gastroenterology 127: S5–S16. doi: 10.1053/j.gastro.2004.09.011
    [264] El-Serag HB, Marrero JA, Rudolph L, et al. (2008) Diagnosis and treatment of hepatocellular carcinoma. Gastroenterology 134: 1752–1763. doi: 10.1053/j.gastro.2008.02.090
    [265] Befeler AS, Di Bisceglie AM (2002) Hepatocellular carcinoma: Diagnosis and treatment. Gastroenterology 122: 1609–1619. doi: 10.1053/gast.2002.33411
    [266] Zhang B, Yang B, Tang Z (2004) Randomized controlled trial of screening for hepatocellular carcinoma. J Cancer Res Clin 130: 417–422.
    [267] Yoshida S, Kurokohchi K, Arima K, et al. (2002) Clinical significance of lens culinaris agglutinin-reactive fraction of serum α-fetoprotein in patients with hepatocellular carcinoma. Int J Oncol 20: 305–309.
    [268] Yamamoto K, Imamura H, Matsuyama Y, et al. (2010) AFP, AFP-L3, DCP, and GP73 as markers for monitoring treatment response and recurrence and as surrogate markers of clinicopathological variables of HCC. J Gastroenterol 45: 1272–1282. doi: 10.1007/s00535-010-0278-5
    [269] Nakamura S, Nouso K, Sakaguchi K, et al. (2006) Sensitivity and specificity of des-gamma-carboxy prothrombin for diagnosis of patients with hepatocellular carcinomas varies according to tumor size. Am J Gastroenterol 101: 2038. doi: 10.1111/j.1572-0241.2006.00681.x
    [270] Tian L, Wang Y, Xu D, et al. (2011) Serological AFP/Golgi protein 73 could be a new diagnostic parameter of hepatic diseases. Int J Cancer 129: 1923–1931. doi: 10.1002/ijc.25838
    [271] Qi J, Wang J, Katayama H, et al. (2013) Circulating microRNAs (cmiRNAs) as novel potential biomarkers for hepatocellular carcinoma. Neoplasma 60: 135.
    [272] Pinato DJ, Pirisi M, Maslen L, et al. (2014) Tissue biomarkers of prognostic significance in hepatocellular carcinoma. Adv Anat Pathol 21: 270–284. doi: 10.1097/PAP.0000000000000029
    [273] Xiao J, Ding Y, Huang J, et al. (2014) The association of HMGB1 gene with the prognosis of HCC. PloS One 9: e89097. doi: 10.1371/journal.pone.0089097
    [274] Dong Y, Cui L, Peng C, et al. (2013) Expression and clinical significance of HMGB1 in human liver cancer: Knockdown inhibits tumor growth and metastasis in vitro and in vivo. Oncol Rep 29: 87–94. doi: 10.3892/or.2012.2070
    [275] Liu F, Zhang Y, Peng Z, et al. (2012) High expression of high mobility group box 1 (hmgb1) predicts poor prognosis for hepatocellular carcinoma after curative hepatectomy. J Transl Med 10: 135. doi: 10.1186/1479-5876-10-135
    [276] Cheng B, Jia C, Liu C, et al. (2008) Serum high mobility group box chromosomal protein 1 is associated with clinicopathologic features in patients with hepatocellular carcinoma. Digest Liver Dis 40: 446–452. doi: 10.1016/j.dld.2007.11.024
    [277] Schnurr M, Duewell P (2014) Induction of immunogenic cell death by targeting RIG-I-like helicases in pancreatic cancer. Oncoimmunology 3: e955687. doi: 10.4161/21624011.2014.955687
    [278] Yang S, Wang X, Contino G, et al. (2011) Pancreatic cancers require autophagy for tumor growth. Gene Dev 25: 717–729. doi: 10.1101/gad.2016111
    [279] Kang R, Tang D (2012) Autophagy in pancreatic cancer pathogenesis and treatment. Am J Cancer Res 2: 383.
    [280] Wittwer C, Boeck S, Heinemann V, et al. (2013) Circulating nucleosomes and immunogenic cell death markers HMGB1, sRAGE and DNAse in patients with advanced pancreatic cancer undergoing chemotherapy. Int J Cancer 133: 2619–2630.
    [281] Tang D, Loze MT, Zeh I, et al. (2010) The redox protein HMGB1 regulates cell death and survival in cancer treatment. Autophagy 6: 1181–1183. doi: 10.4161/auto.6.8.13367
    [282] Rosenfeldt MT, O'prey J, Morton JP, et al. (2013) p53 status determines the role of autophagy in pancreatic tumour development. Nature 504: 296. doi: 10.1038/nature12865
    [283] Tarbé N, Evtimova V, Burtscher H, et al. (2001) Transcriptional profiling of cell lines derived from an orthotopic pancreatic tumor model reveals metastasis-associated genes. Anticancer Res 21: 3221–3228.
    [284] Nestl A, Von Stein OD, Zatloukal K, et al. (2001) Gene expression patterns associated with the metastatic phenotype in rodent and human tumors. Cancer Res 61: 1569–1577.
    [285] Huang Q, Wang G, Sun N, et al. (2004) Inhibitory effects of high mobility group box 1 antisense nucleotide on invasion of human pancreatic cancer cell line PCNA-1. Ai Zheng 23: 1036–1040.
    [286] Boeck S, Wittwer C, Heinemann V, et al. (2013) Cytokeratin 19-fragments (CYFRA 21-1) as a novel serum biomarker for response and survival in patients with advanced pancreatic cancer. Brit J Cancer 108: 1684. doi: 10.1038/bjc.2013.158
    [287] Yasuda T, Ueda T, Takeyama Y, et al. (2006) Significant increase of serum high-mobility group box chromosomal protein 1 levels in patients with severe acute pancreatitis. Pancreas 33: 359–363. doi: 10.1097/01.mpa.0000236741.15477.8b
    [288] Stoetzer OJ, Wittwer C, Lehner J, et al. (2012) Circulating nucleosomes and biomarkers of immunogenic cell death as predictive and prognostic markers in cancer patients undergoing cytotoxic therapy. Expert Opin Biol Th 12: S217–S224. doi: 10.1517/14712598.2012.689280
    [289] Bauden M, Pamart D, Ansari D, et al. (2015) Circulating nucleosomes as epigenetic biomarkers in pancreatic cancer. Clin Epigenet 7: 106. doi: 10.1186/s13148-015-0139-4
    [290] Wang J, Yu H, Bei S, et al. (2016) Association of HMGB1 Gene Polymorphisms with Risk of Colorectal Cancer in a Chinese Population. Med Sci Monit: Int Med J Exp Clin Res 22: 3419.
    [291] Schirrmacher V, Fournier P, Schlag P (2014) Autologous tumor cell vaccines for post-operative active-specific immunotherapy of colorectal carcinoma: Long-term patient survival and mechanism of function. Expert Rev Vaccines 13: 117–130. doi: 10.1586/14760584.2014.854169
    [292] Zhang S, Han Z, Jing Y, et al. (2012) CD133+ CXCR4+ colon cancer cells exhibit metastatic potential and predict poor prognosis of patients. BMC Med 10: 85. doi: 10.1186/1741-7015-10-85
    [293] Feigelson HS, Zeng C, Pawloski PA, et al. (2014) Does KRAS testing in metastatic colorectal cancer impact overall survival? A comparative effectiveness study in a population-based sample. PloS One 9: e94977.
    [294] Peng R, Wu X, Ding Y, et al. (2010) Co-expression of nuclear and cytoplasmic HMGB1 is inversely associated with infiltration of CD45RO+ T cells and prognosis in patients with stage IIIB colon cancer. Bmc Cancer 10: 496. doi: 10.1186/1471-2407-10-496
    [295] Wagner P, Koch M, Nummer D, et al. (2008) Detection and functional analysis of tumor infiltrating T-lymphocytes (TIL) in liver metastases from colorectal cancer. Ann Surg Oncol 15: 2310. doi: 10.1245/s10434-008-9971-5
    [296] Ohtani H (2007) Focus on TILs: Prognostic significance of tumor infiltrating lymphocytes in human colorectal cancer. Cancer Immunity Archive 7: 4.
    [297] Galon J, Costes A, Sanchez-Cabo F, et al. (2006) Type, density, and location of immune cells within human colorectal tumors predict clinical outcome. Science 313: 1960–1964. doi: 10.1126/science.1129139
    [298] Vezys V, Yates A, Casey KA, et al. (2009) Memory CD8 T-cell compartment grows in size with immunological experience. Nature 457: 196. doi: 10.1038/nature07486
    [299] Zhang Z, Wang M, Zhou L, et al. (2015) Increased HMGB1 and cleaved caspase-3 stimulate the proliferation of tumor cells and are correlated with the poor prognosis in colorectal cancer. J Exp Clin Canc Res 34: 51. doi: 10.1186/s13046-015-0166-1
    [300] Cheng J, Tian L, Ma J, et al. (2015) Dying tumor cells stimulate proliferation of living tumor cells via caspase-dependent protein kinase Cδ activation in pancreatic ductal adenocarcinoma. Mol oncol 9: 105–114. doi: 10.1016/j.molonc.2014.07.024
    [301] Huang Q, Li F, Liu X, et al. (2011) Caspase 3-mediated stimulation of tumor cell repopulation during cancer radiotherapy. Nat Med 17: 860–866. doi: 10.1038/nm.2385
    [302] Kurtova AV, Xiao J, Mo Q, et al. (2015) Blocking PGE2-induced tumour repopulation abrogates bladder cancer chemoresistance. Nature 517: 209. doi: 10.1038/nature14034
    [303] Shepherd N, Richman P, England J (1988) Ki‐67 derived proliferative activity in colorectal adenocarcinoma with prognostic correlations. J Pathol 155: 213–219. doi: 10.1002/path.1711550306
    [304] Livesey KM, Kang R, Zeh I, et al. (2012) Direct molecular interactions between HMGB1 and TP53 in colorectal cancer. Autophagy 8: 846–848. doi: 10.4161/auto.19891
    [305] Schlueter C, Weber H, Meyer B, et al. (2005) Angiogenetic signaling through hypoxia: HMGB1: an angiogenetic switch molecule. Am J Pathol 166: 1259–1263. doi: 10.1016/S0002-9440(10)62344-9
    [306] Srikrishna G, Freeze HH (2009) Endogenous damage-associated molecular pattern molecules at the crossroads of inflammation and cancer. Neoplasia 11: 615–628. doi: 10.1593/neo.09284
    [307] Qie G, Wang C, Chu Y, et al. (2015) Expression of HMGB1/RAGE protein in renal carcinoma and its clinical significance. Int J Clin Exp Pathol 8: 6262.
    [308] Zhu P, Xie L, Ding H, et al. (2013) High mobility group box 1 and kidney diseases. Int J Mol Med 31: 763–768. doi: 10.3892/ijmm.2013.1286
    [309] Sandim V, Pereira DA, Ornellas AA, et al. (2010) Renal cell carcinoma and proteomics. Urol Int 84: 373–377. doi: 10.1159/000296283
    [310] Nelson N, Szekeres K, Cooper D, et al. (2012) Preparation of myeloid derived suppressor cells (MDSC) from naive and pancreatic tumor-bearing mice using flow cytometry and automated magnetic activated cell sorting (AutoMACS). J Visualized Exp 18: e3875.
    [311] Ostrand-Rosenberg S, Sinha P, Beury DW, et al. (2012) Cross-talk between myeloid-derived suppressor cells (MDSC), macrophages, and dendritic cells enhances tumor-induced immune suppression. Semin Cancer Biol 22: 275–281. doi: 10.1016/j.semcancer.2012.01.011
    [312] Kowanetz M, Wu X, Lee J, et al. (2010) Granulocyte-colony stimulating factor promotes lung metastasis through mobilization of Ly6G+ Ly6C+ granulocytes. Proc Natl Acad Sci 107: 21248–21255. doi: 10.1073/pnas.1015855107
    [313] Qu P, Yan C, Blum JS, et al. (2011) Myeloid-specific expression of human lysosomal acid lipase corrects malformation and malfunction of myeloid-derived suppressor cells in lal−/− mice. J Immunol 187: 3854–3866. doi: 10.4049/jimmunol.1003358
    [314] Liu C, Yu S, Kappes J, et al. (2007) Expansion of spleen myeloid suppressor cells represses NK cell cytotoxicity in tumor-bearing host. Blood 109: 4336–4342. doi: 10.1182/blood-2006-09-046201
    [315] Aliper AM, Frieden-Korovkina VP, Buzdin A, et al. (2014) Interactome analysis of myeloid-derived suppressor cells in murine models of colon and breast cancer. Oncotarget 5: 11345.
    [316] Gabrilovich D (2004) Mechanisms and functional significance of tumour-induced dendritic-cell defects. Nat Rev Immunol 4: 941. doi: 10.1038/nri1498
    [317] Serafini P, Mgebroff S, Noonan K, et al. (2008) Myeloid-derived suppressor cells promote cross-tolerance in B-cell lymphoma by expanding regulatory T cells. Cancer Res 68: 5439–5449. doi: 10.1158/0008-5472.CAN-07-6621
    [318] Sinha P, Parker KH, Horn L, et al. (2012) Tumor‐induced myeloid‐derived suppressor cell function is independent of IFN‐γ and IL‐4Rα. Eur J Immunol 42: 2052–2059. doi: 10.1002/eji.201142230
    [319] Dumitriu IE, Bianchi ME, Bacci M, et al. (2007) The secretion of HMGB1 is required for the migration of maturing dendritic cells. J Leukocyte Biol 81: 84–91. doi: 10.1189/jlb.0306171
    [320] Lin L, Zhong K, Sun Z, et al. (2012) Receptor for advanced glycation end products (RAGE) partially mediates HMGB1-ERKs activation in clear cell renal cell carcinoma. J Cancer Res Clin 138: 11–22. doi: 10.1007/s00432-011-1067-0
    [321] Homma Y, Akaza H, Okada K, et al. (2004) Radical prostatectomy and adjuvant endocrine therapy for prostate cancer with or without preoperative androgen deprivation: Five‐year results. Int J Urol 11: 295–303. doi: 10.1111/j.1442-2042.2004.00795.x
    [322] De Marzo AM, Nakai Y, Nelson WG (2007) Inflammation, atrophy, and prostate carcinogenesis. Urol Oncol 5: 398–400.
    [323] Mettlin C (1997) Recent developments in the epidemiology of prostate cancer. Eur J Cancer 33: 340–347. doi: 10.1016/S0959-8049(97)89003-X
    [324] Wang Y, Kreisberg JI, Ghosh PM (2007) Cross-talk between the androgen receptor and the phosphatidylinositol 3-kinase/Akt pathway in prostate cancer. Curr Cancer Drug Targets 7: 591–604. doi: 10.2174/156800907781662248
    [325] Scher HI, Sawyers CL (2005) Biology of progressive, castration-resistant prostate cancer: directed therapies targeting the androgen-receptor signaling axis. J Clin Oncol 23: 8253–8261. doi: 10.1200/JCO.2005.03.4777
    [326] Brusa D, Migliore E, Garetto S, et al. (2009) Immunogenicity of 56 °C and UVC‐treated prostate cancer is associated with release of HSP70 and HMGB1 from necrotic cells. Prostate 69: 1343–1352. doi: 10.1002/pros.20981
    [327] Chen T, Guo J, Han C, et al. (2009) Heat shock protein 70, released from heat-stressed tumor cells, initiates antitumor immunity by inducing tumor cell chemokine production and activating dendritic cells via TLR4 pathway. J Immunol 182: 1449–1459.
    [328] Sanda MG, Restifo NP, Walsh JC, et al. (1995) Molecular characterization of defective antigen processing in human prostate cancer. J Natl Cancer Inst 87: 280–285. doi: 10.1093/jnci/87.4.280
    [329] Bronte V, Kasic T, Gri G, et al. (2005) Boosting antitumor responses of T lymphocytes infiltrating human prostate cancers. J Exp Med 201: 1257–1268. doi: 10.1084/jem.20042028
    [330] Korbelik M, Sun J, Cecic I (2005) Photodynamic therapy-induced cell surface expression and release of heat shock proteins: relevance for tumor response. Cancer Res 65: 1018–1026.
    [331] Bianchini D, Zivi A, Sandhu S, et al. (2010) Horizon scanning for novel therapeutics for the treatment of prostate cancer. Ann Oncol 21: vii43–vii55.
    [332] Shetty AV, Thirugnanam S, Dakshinamoorthy G, et al. (2011) 18α-glycyrrhetinic acid targets prostate cancer cells by down-regulating inflammation-related genes. Int J Oncol 39: 635–640.
    [333] Kuniyasu H, Chihara Y, Kondo H, et al. (2003) Amphoterin induction in prostatic stromal cells by androgen deprivation is associated with metastatic prostate cancer. Oncol Rep 10: 1863–1868.
    [334] Song D, Chen A, Guo F, et al. (2008) Differential proteomic analysis and function study of human prostate carcinoma cells with different osseous metastatic tendency. Zhonghua Yi Xue Za Zhi 88: 1197–1201.
    [335] Saenz R, da Silva Souza C, Huang CT, et al. (2010) HMGB1-derived peptide acts as adjuvant inducing immune responses to peptide and protein antigen. Vaccine 28: 7556–7562. doi: 10.1016/j.vaccine.2010.08.054
    [336] Babjuk M, Burger M, Zigeuner R, et al. (2013) EAU guidelines on non-muscle-invasive urothelial carcinoma of the bladder: Update 2013. Eur Urol 64: 639–653. doi: 10.1016/j.eururo.2013.06.003
    [337] Parekh DJ, Bochner BH, Dalbagni G (2006) Superficial and muscle-invasive bladder cancer: Principles of management for outcomes assessments. J Clin Oncol 24: 5519–5527. doi: 10.1200/JCO.2006.08.5431
    [338] Malkowicz SB, Van Poppel H, Mickisch G, et al. (2007) Muscle-invasive urothelial carcinoma of the bladder. Urology 69: 3–16. doi: 10.1016/S0090-4295(07)00617-6
    [339] Jiang H, Hu X, Zhang H, et al. (2017) Down-regulation of LncRNA TUG1 enhances radiosensitivity in bladder cancer via suppressing HMGB1 expression. Radiat Oncol 12: 65. doi: 10.1186/s13014-017-0802-3
    [340] Zhu Y, Luo M, Brooks M, et al. (2014) Biological and clinical significance of cancer stem cell plasticity. Clin Transl Med 3: 32. doi: 10.1186/s40169-014-0032-3
    [341] Prensner JR, Chinnaiyan AM (2011) The emergence of lncRNAs in cancer biology. Cancer Discovery 1: 391–407. doi: 10.1158/2159-8290.CD-11-0209
    [342] Qiu M, Hu J, Yin R, et al. (2013) Long noncoding RNA: An emerging paradigm of cancer research. Tumor Biol 34: 613–620. doi: 10.1007/s13277-013-0658-6
    [343] Jiang Y, Li Z, Zheng S, et al. (2016) The long non-coding RNA HOTAIR affects the radiosensitivity of pancreatic ductal adenocarcinoma by regulating the expression of Wnt inhibitory factor 1. Tumor Biol 37: 3957–3967. doi: 10.1007/s13277-015-4234-0
    [344] Iliev R, Kleinova R, Juracek J, et al. (2016) Overexpression of long non-coding RNA TUG1 predicts poor prognosis and promotes cancer cell proliferation and migration in high-grade muscle-invasive bladder cancer. Tumor Biol 37: 13385–13390. doi: 10.1007/s13277-016-5177-9
    [345] Tan J, Qiu K, Li M, et al. (2015) Double-negative feedback loop between long non-coding RNA TUG1 and miR-145 promotes epithelial to mesenchymal transition and radioresistance in human bladder cancer cells. Febs Lett 589: 3175–3181. doi: 10.1016/j.febslet.2015.08.020
    [346] Liao H, Xiao Y, Hu Y, et al. (2015) Suppression of cellular proliferation and invasion by HMGB1 knockdown in bladder urothelial carcinoma cells. Oncol Res Featuring Preclin Clin Cancer Ther 22: 235–245.
    [347] Yang G, Zhang L, Bo J, et al. (2012) Increased expression of HMGB1 is associated with poor prognosis in human bladder cancer. J Surg Oncol 106: 57–61. doi: 10.1002/jso.23040
    [348] Štros M, Cherny D, Jovin TM (2000) HMG1 protein stimulates DNA end joining by promoting association of DNA molecules via their ends. Eur J Biochem 267: 4088–4097. doi: 10.1046/j.1432-1327.2000.01450.x
    [349] O'Driscoll M, Jeggo PA (2006) The role of double-strand break repair-insights from human genetics. Nat Rev Genet 7: 45.
    [350] Wei Y, An Z, Zou Z, et al. (2015) The stress-responsive kinases MAPKAPK2/MAPKAPK3 activate starvation-induced autophagy through Beclin 1 phosphorylation. Elife 4: e05289. doi: 10.7554/eLife.05289
    [351] Trocoli A, Djavaheri-Mergny M (2011) The complex interplay between autophagy and NF-κB signaling pathways in cancer cells. Am J Cancer Res 1: 629.
    [352] Wang H, Li Z, Sun Y, et al. (2015) Relationship between high-mobility group box 1 overexpression in ovarian cancer tissue and serum: a meta-analysis. OncoTargets Ther 8: 3523.
    [353] Zhang W, Tian J, Hao Q (2014) HMGB1 combining with tumor-associated macrophages enhanced lymphangiogenesis in human epithelial ovarian cancer. Tumour Biol 35: 2175–2186. doi: 10.1007/s13277-013-1288-8
    [354] Zhang Y, Cheng Y, Ren X, et al. (2012) NAC1 Modulates Sensitivity of Ovarian Cancer Cells to Cisplatin via Altering the HMGB1-Mediated Autophagic Response. Oncogene 31: 1055. doi: 10.1038/onc.2011.290
    [355] Jinawath N, Vasoontara C, Yap KL, et al. (2009) NAC-1, a potential stem cell pluripotency factor, contributes to paclitaxel resistance in ovarian cancer through inactivating Gadd45 pathway. Oncogene 28: 1941. doi: 10.1038/onc.2009.37
    [356] Chen J, Liu X, Zhang J, et al. (2012) Targeting HMGB1 inhibits ovarian cancer growth and metastasis by lentivirus‐mediated RNA interference. J Cell Physiol 227: 3629–3638. doi: 10.1002/jcp.24069
    [357] Zhang L, Jiang F, Xu L, et al. (2012) The Role of Cyclin D1 Expression and Patient's Survival in Non-Small-Cell Lung Cancer: A Systematic Review With Meta-Analysis. Clin Lung Cancer 13: 188–195. doi: 10.1016/j.cllc.2011.10.003
    [358] Naryzhny S (2008) Proliferating cell nuclear antigen: a proteomics view. Cell Mol Life Sci 65: 3789–3808. doi: 10.1007/s00018-008-8305-x
    [359] Duluc D, Delneste Y, Tan F, et al. (2007) Tumor-associated leukemia inhibitory factor and IL-6 skew monocyte differentiation into tumor-associated macrophage-like cells. Blood 110: 4319–4330. doi: 10.1182/blood-2007-02-072587
    [360] Ungefroren H, Sebens S, Seidl D, et al. (2011) Interaction of tumor cells with the microenvironment. Cell Commun Signal 9: 18. doi: 10.1186/1478-811X-9-18
    [361] Kelly PN, Strasser A (2011) The role of Bcl-2 and its pro-survival relatives in tumourigenesis and cancer therapy. Cell Death Differ 18: 1414. doi: 10.1038/cdd.2011.17
    [362] Schmidt SV, Seibert S, Walch-Rückheim B, et al. (2015) RIPK3 expression in cervical cancer cells is required for PolyIC-induced necroptosis, IL-1α release, and efficient paracrine dendritic cell activation. Oncotarget 6: 8635.
    [363] Cheng H, Wang W, Zhang Y, et al. (2017) Expression levels and clinical significance of hepsin and HMGB1 proteins in cervical carcinoma. Oncol Lett 14: 159–164. doi: 10.3892/ol.2017.6116
    [364] Tanimoto H, Yan Y, Clarke J, et al. (1997) Hepsin, a cell surface serine protease identified in hepatoma cells, is overexpressed in ovarian cancer. Cancer Res 57: 2884–2887.
    [365] Lipari MT, Li W, Moran P, et al. (2012) Furin-cleaved proprotein convertase subtilisin/kexin type 9 (PCSK9) is active and modulates low density lipoprotein receptor and serum cholesterol levels. J Biol Chem 287: 43482–43491. doi: 10.1074/jbc.M112.380618
    [366] Goulet B, Chan G, Chambers AF, et al. (2011) An emerging role for the nuclear localization of maspin in the suppression of tumor progression and metastasis 1 1 This article is part of Special Issue entitled Asilomar Chromatin and has undergone the Journal's usual peer review process. Biochem Cell Biol 90: 22–38.
    [367] Secord AA, Darcy KM, Hutson A, et al. (2011) The regulation of MASPIN expression in epithelial ovarian cancer: Association with p53 status, and MASPIN promoter methylation: A gynecologic oncology group study. Gynecol Oncol 123: 314–319. doi: 10.1016/j.ygyno.2011.08.003
    [368] Niebler M, Qian X, Höfler D, et al. (2013) Post-translational control of IL-1β via the human papillomavirus type 16 E6 oncoprotein: A novel mechanism of innate immune escape mediated by the E3-ubiquitin ligase E6-AP and p53. Plos Pathog 9: e1003536. doi: 10.1371/journal.ppat.1003536
    [369] Kleine‐Lowinski K, Gillitzer R, Kühne‐Heid R, et al. (1999) Monocyte‐chemo‐attractant‐protein‐1 (mcp‐1)‐gene expression in cervical intra‐epithelial neoplasias and cervical carcinomas. Int J Cancer 82: 6–11. doi: 10.1002/(SICI)1097-0215(19990702)82:1<6::AID-IJC2>3.0.CO;2-3
    [370] Iglesias M, Yen K, Gaiotti D, et al. (1998) Human papillomavirus type 16 E7 protein sensitizes cervical keratinocytes to apoptosis and release of interleukin-1α. Oncogene 17.
    [371] Peralta-Zaragoza O, Bermúdez-Morales VH, Pérez-Plasencia C, et al. (2012) Targeted treatments for cervical cancer: A review. OncoTargets Ther 5: 315.
    [372] Panaretakis T, Kepp O, Brockmeier U, et al. (2009) Mechanisms of pre‐apoptotic calreticulin exposure in immunogenic cell death. EMBO J 28: 578–590.
    [373] Michaud M, Martins I, Sukkurwala AQ, et al. (2011) Autophagy-dependent anticancer immune responses induced by chemotherapeutic agents in mice. Science 334: 1573–1577. doi: 10.1126/science.1208347
    [374] Srivastava P (2002) Roles of heat-shock proteins in innate and adaptive immunity. Nat Rev Immunol 2: 185. doi: 10.1038/nri749
    [375] Jiang D, Wang H, Li Z, et al. (2017) MiR-142 inhibits the development of cervical cancer by targeting HMGB1. Oncotarget 8: 4001–4007.
    [376] Sheng X, Du X, Zhang X, et al. (2009) Clinical value of serum HMGB1 levels in early detection of recurrent squamous cell carcinoma of uterine cervix: comparison with serum SCCA, CYFRA21-1, and CEA levels. Croat Med J 50: 455–464. doi: 10.3325/cmj.2009.50.455
    [377] Du X, Fu X, Hao Q (2007) Expression and clinical significance of HMGB1 and RAGE in cervical carcinoma. Chin J Clin Oncol 4: 343–346. doi: 10.1007/s11805-007-0343-x
  • Reader Comments
  • © 2018 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(5002) PDF downloads(1804) Cited by(2)

Article outline

Figures and Tables

Figures(1)  /  Tables(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog