Skip to main content

Advertisement

Log in

The role of the microenvironment in regulation of CSPG-driven invasive and non-invasive tumor growth in glioblastoma

Brain tumor growth

  • Original Paper
  • Area 3
  • Published:
Japan Journal of Industrial and Applied Mathematics Aims and scope Submit manuscript

Abstract

Glioblastoma (GBM) is one of the most lethal type of brain cancer with poor survival time. GBM is characterized by infiltration of the cancer cells through the brain tissue while lower grade gliomas and other non-neural metastatic types form self-contained non-invasive lesions. Glycosylated chondroitin sulfate proteoglycans (CSPGs), acting as critical regulators of the tumor microenvironment, dramatically govern the spatiotemporal status of resident reactive astrocytes and activation of tumor-associated microglia. In this paper we develop a mathematical model to investigate the effect of the CSPG distribution on regulation of a fundamental switch between two distinct patterns: invasive and non-invasive tumors. We show that the model’s predictions agree with experimental results for a spherical glioma. The model specifically predicts that non-invasive tumor lesions are highly associated with a thick extracellular matrix (ECM) containing rich CSPGs, while the absence of glycosylated CSPGs results in diffusely infiltrative tumors. It is also shown that heavy CSPGs can drive the exodus of resident reactive astrocytes from the main tumor mass, and direct inhibition of tumor invasion by the astrogliotic capsule, leading to encapsulation of non-invasive lesions. However, stable residence of reactive astrocytes from GBM in the absence or low level of CSPGs presents a microenvironment favorable to diffuse infiltration due to loss of the primary (CSPG-induced cell-ECM bonding) and secondary (astrogliotic capsule) inhibitors. The mathematical model presents the clear role of the key tumor microenvironment in brain tumor invasion.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10

Similar content being viewed by others

References

  1. Aguzzi, A., Barres, B.A., Bennett, M.L.: Microglia: scapegoat, saboteur, or something else? Science 339(6116), 156–161 (2013)

    Article  Google Scholar 

  2. Arslan, F., Bosserhoff, A., Nickl-Jockschat, T., Doerfelt, A., Bogdahn, U., Hau, P.: The role of versican isoforms V0/V1 in glioma migration mediated by transforming growth factor-\(\beta \)2. Br. J. Cancer 96(10), 1560–1568 (2007)

    Article  Google Scholar 

  3. Aubert, M., Badoual, M., Freol, S., Christov, C., Grammaticos, B.: A cellular automaton model for the migration of glioma cells. Phys. Biol. 3(2), 93–100 (2006)

    Article  Google Scholar 

  4. Badoual, M., Deroulers, C., Aubert, M., Grammaticos, B.: Modelling intercellular communication and its effects on tumour invasion. Phys. Biol. 7(4), 046,013 (2010)

    Article  Google Scholar 

  5. Badri, H., Pitter, K., Holland, E., Michor, F., Leder, K.: Optimization of radiation dosing schedules for proneural glioblastoma. J. Math. Biol. 1–36 (2015). doi:10.1007/s00285-015-0908-x

  6. Barcellos-Hoff, M., Newcomb, E., Zagzag, D., Narayana, A.: Therapeutic targets in malignant glioblastoma microenvironment. Semin. Radiat. Oncol. 19(3), 163–170 (2009)

    Article  Google Scholar 

  7. Baron, R., Babcock, A.A., Nemirovsky, A., Finsen, B., Monsonego, A.: Accelerated microglial pathology is associated with ab plaques in mouse models of alzheimer’s disease. Aging Cell 13, 584–595 (2014)

    Article  Google Scholar 

  8. Bellail, A., Hunter, S., Brat, D., Tan, C., Meir, E.V.: Microregional extracellular matrix heterogeneity in brain modulates glioma cell invasion. Int. J. Biochem. Cell Biol. 36, 1046–1069 (2004)

    Article  Google Scholar 

  9. Bignami, A., Hosley, M., Dahl, D.: Hyaluronic acid and hyaluronic acid-binding proteins in brain extracellular matrix. Anat. Embryol. 188(5), 419–433 (1993)

    Article  Google Scholar 

  10. Black, P., Agarwal, P., Dinney, C.: Targeted therapies in bladder cancer—an update. Urol. Oncol. 25(5), 433–438 (2007)

    Article  Google Scholar 

  11. Brittis, P.A., Canning, D.R., Silver, J.: Chondroitin sulfate as a regulator of neuronal patterning in the retina. Science 255(5045), 733–736 (1992)

    Article  Google Scholar 

  12. Bruckner, G., Bringmann, A., Hartig, W., Koppe, G., Delpech, B., Brauer, K.: Acute and long-lasting changes in extracellular-matrix chondroitin-sulphate proteoglycans induced by injection of chondroitinase abc in the adult rat brain. Exp. Brain Res. 121, 300–310 (1998)

    Article  Google Scholar 

  13. Bush, T.G., Puvanachandra, N., Horner, C.H., Polito, A., Ostenfeld, T., Svendsen, C.N., Mucke, L., Johnson, M.H., Sofroniew, M.V.: Leukocyte infiltration, neuronal degeneration, and neurite outgrowth after ablation of scar-forming, reactive astrocytes in adult transgenic mice. Neuron 23(2), 297–308 (1999)

    Article  Google Scholar 

  14. Casciari, J., Sotirchos, S., Sutherland, R.: Mathematical modelling of microenvironment and growth in EMT6/Ro multicellular tumour spheroids. Cell Prolif. 25(1), 1–22 (1992)

    Article  Google Scholar 

  15. Charles, N., HOLLAND, E., Gilbertson, R., GLASS, R., Kettenmann, H.: The brain tumor microenvironment. Glia 59, 1169–1180 (2011)

    Article  Google Scholar 

  16. Chau, C., Shum, D., Li, H., Pei, J., Lui, Y., Wirthlin, L.: Chondroitinase ABC enhances axonal regrowth through Schwann cell-seeded guidance channels after spinal cord injury. FASEB J. 18, 194–196 (2004)

    Google Scholar 

  17. Chauviere, A., Preziosi, L., Byrne, H.: A model of cell migration within the extracellular matrix based on a phenotypic switching mechanism. Math. Med. Biol. 27(3), 255–281 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  18. Chicoine, M., Madsen, C., Silbergeld, D.: Modification of human glioma locomotion in vitro by cytokines EGF, bFGF, PDGFbb, NGF, and TNF alpha. Neurosurgery 36(6), 1165–1170 (1995). (disbrain cancer)

    Article  Google Scholar 

  19. Chintala, S., Tonn, J., Rao, J.: Matrix metalloproteinases and their biological function in human gliomas. Int. J. Dev. Neurosci. 17(5–6), 495–502 (1999)

    Article  Google Scholar 

  20. Choucair, A., Levin, V., Gutin, P., Davis, R., Silver, P., Edwards, M., Wilson, C.: Development of multiple lesions during radiation therapy and chemotherapy in patients with gliomas. J. Neurosurg. 65(5), 654–658 (1986)

    Article  Google Scholar 

  21. Cooper, N., Steindler, D.: Lectins demarcate the barrel subfield in the somatosensory cortex of the early postnatal mouse. J. Comp. Neurol. 249(2), 157–169 (1986)

    Article  Google Scholar 

  22. Corwin, D., Holdsworth, C., Rockne, R., Trister, A., Mrugala, M., Rockhill, J., Stewart, R., Phillips, M., Swanson, K.: Toward patient-specific, biologically optimized radiation therapy plans for the treatment of glioblastoma. PLoS One 8(11) (2013)

  23. Crespo, D., Asher, R.A., Lin, R., Rhodes, K.E., Fawcett, J.W.: How does chondroitinase promote functional recovery in the damaged CNS? Exp. Neurol. 206(2), 159–171 (2007)

    Article  Google Scholar 

  24. Davis, F., McCarthy, B.: Current epidemiological trends and surveillance issues in brain tumors. Expert Rev. Anticancer Ther. 1(3), 395–401 (2001)

    Article  Google Scholar 

  25. DeAngelis, L.M.: Brain tumors. N. Engl. J. Med. 344(2), 114–123 (2001)

    Article  MathSciNet  Google Scholar 

  26. Deepa, S., Carulli, D., Galtrey, C., Rhodes, K., Fukuda, J., Mikami, T., Sugahara, K., Fawcett, J.: Composition of perineuronal net extracellular matrix in rat brain: a different disaccharide composition for the net-associated proteoglycans. J. Biol. Chem. 281(26), 17789–17800 (2006)

    Article  Google Scholar 

  27. Deisboeck, T., Berens, M., Kansal, A., Torquato, S., Stemmer-Rachamimov, A., Chiocca, E.: Pattern of self-organization in tumour systems: complex growth dynamics in a novel brain tumour spheroid model. Cell Prolif. 34, 115–134 (2001)

    Article  Google Scholar 

  28. Demuth, T., Berens, M.: Molecular mechanisms of glioma cell migration and invasion. J Neurooncol. 70(2), 217–228 (2004)

    Article  Google Scholar 

  29. Demuth, T., Hopf, N., Kempski, O., Sauner, D., Herr, M., Giese, A., Perneczky, A.: Migratory activity of human glioma cell lines in vitro assessed by continuous single cell observation. Clin. Exp. Metastasis 18(7), 589–597 (2000)

    Article  Google Scholar 

  30. Dmitrieva, N., Yu, L., Viapiano, M., Cripe, T., Chiocca, E., Glorioso, J., Kaur, B.: Choindroitinase ABC I-mediated enhancement of oncolytic virus spread and antitumor efficacy. Clin. Cancer Res. 17(6), 1362–1372 (2011)

    Article  Google Scholar 

  31. Dutt, S., Kléber, M., Matasci, M., Sommer, L., Zimmermann, D.R.: Versican v0 and v1 guide migratory neural crest cells. J. Biol. Chem. 281(17), 12123–12131 (2006)

    Article  Google Scholar 

  32. Dutt, S., Matasci, M., Sommer, L., Zimmermann, D.R.: Guidance of neural crest cell migration: the inhibitory function of the chondroitin sulfate proteoglycan, versican. Sci. World J. 6, 1114–1117 (2006)

    Article  Google Scholar 

  33. Edwards, L.A., Woolard, K., Son, M.J., Li, A., Lee, J., Ene, C., Mantey, S.A., Maric, D., Song, H., Belova, G., et al.: Effect of brain-and tumor-derived connective tissue growth factor on glioma invasion. J. Natl. Cancer Inst. 103(15), 1162–1178 (2011)

    Article  Google Scholar 

  34. Faissner, A., Steindler, D.: Boundaries and inhibitory molecules in developing neural tissues. Glia 13(4), 233–254 (1995)

    Article  Google Scholar 

  35. Farin, A., Suzuki, S.O., Weiker, M., Goldman, J.E., Bruce, J.N., Canoll, P.: Transplanted glioma cells migrate and proliferate on host brain vasculature: a dynamic analysis. Glia 53(8), 799–808 (2006). doi:10.1002/glia.20334

    Article  Google Scholar 

  36. Freyer, J., Sutherland, R.: A reduction in the in situ rates of oxygen and glucose consumption of cells in EMT6/Ro spheroids during growth. J. Cell Physiol. 124(3), 516–524 (1985)

    Article  Google Scholar 

  37. Friedman, A., Tian, J., Fulci, G., Chiocca, E., Wang, J.: Glioma virotherapy: effects of innate immune suppression and increased viral replication capacity. Cancer Res. 66(4), 2314–2319 (2006)

    Article  Google Scholar 

  38. Fryer, H., Kelly, G., Molinaro, L., Hockfield, S.: The high molecular weight cat-301 chondroitin sulfate proteoglycan from brain is related to the large aggregating proteoglycan from cartilage, aggrecan. J. Biol. Chem. 267(14), 9874–9883 (1992)

    Google Scholar 

  39. Furnari, F., Fenton, T., Bachoo, R., Mukasa, A., Stommel, J., Stegh, A., Hahn, W., Ligon, K., Louis, D., Brennan, C., Chin, L., DePinho, R., Cavenee, W.: Malignant astrocytic glioma: genetics, biology, and paths to treatment. Genes Dev. 21(21), 2683–2710 (2007)

    Article  Google Scholar 

  40. Gerin, C., Pallud, J., Grammaticos, B., Mandonnet, E., Deroulers, C., Varlet, P., Capelle, L., Taillandier, L., Bauchet, L., Duffau, H., et al.: Improving the time-machine: estimating date of birth of grade II gliomas. Cell Prolif. 45(1), 76–90 (2012)

    Article  Google Scholar 

  41. Gerlee, P., Nelander, S.: The impact of phenotypic switching on glioblastoma growth and invasion. PLoS Comput. Biol. 8(6), e1002556 (2012)

    Article  MathSciNet  Google Scholar 

  42. Gladson, C.: The extracellular matrix of gliomas: modulation of cell function. J. Neuropathol. Exp. Neurol. 58, 1029–1040 (1999)

    Article  Google Scholar 

  43. Godlewski, J., Bronisz, A., Nowicki, M., Chiocca, E., Lawler, S.: microRNA-451: A conditional switch controlling glioma cell proliferation and migration. Cell Cycle 9(14), 2742–2748 (2010)

    Article  Google Scholar 

  44. Godlewski, J., Nowicki, M., Bronisz, A., Palatini, G.N.J., Lay, M.D., Brocklyn, J., Ostrowski, M., Chiocca, E., Lawler, S.: MircroRNA-451 regulates LKB1/AMPK signaling and allows adaptation to metabolic stress in glioma cells. Mol. Cell 37, 620–632 (2010)

    Article  Google Scholar 

  45. Golding, J.P., Tidcombe, H., Tsoni, S., Gassmann, M.: Chondroitin sulphate-binding molecules may pattern central projections of sensory axons within the cranial mesenchyme of the developing mouse. Dev. Biol. 216(1), 85–97 (1999)

    Article  Google Scholar 

  46. Gotoh, N.: Regulation of growth factor signaling by FRS2 family docking/scaffold adaptor proteins. Cancer Sci. 99(7), 1319–1325 (2008)

    Article  Google Scholar 

  47. Green, N., Morrison, J., Hale, S., Briggs, S., Stevenson, M., Subr, V., Ulbrich, K., Chandler, L., Mautner, V., Seymour, L., Fisher, K.: Retargeting polymer-coated adenovirus to the FGF receptor allows productive infection and mediates efficacy in a peritoneal model of human ovarian cancer. J. Gene. Med. 10(3), 280–289 (2008)

    Article  Google Scholar 

  48. Grumet, M., Friedlander, D., Sakurai, T.: Functions of brain chondroitin sulfate proteoglycans during developments: interactions with adhesion molecules. Perspect. Dev. Neurobiol. 3(4), 319–330 (1996)

    Google Scholar 

  49. Gu, W., Fu, S., Wang, Y., Li, Y., Lu, H., Xu, X., Lu, P.: Chondroitin sulfate proteoglycans regulate the growth, differentiation and migration of multipotent neural precursor cells through the integrin signaling pathway. BMC Neurosci. 10(128), 1–15 (2009)

    Google Scholar 

  50. Harpold, H., EC, J., Swanson, K.: The evolution of mathematical modeling of glioma proliferation and invasion. J. Neuropathol. Exp. Neurol. 66(1), 1–9 (2007)

    Article  Google Scholar 

  51. Hatzikirou, H., Basanta, D., Simon, M., Schaller, K., Deutsch, A.: ‘go or grow’: the key to the emergence of invasion in tumour progression? Math. Med. Biol. 27, 255–281 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  52. Hegedus, B., Zach, J., Czirok, A., Lovey, J., Vicsek, T.: Irradiation and taxol treatment result in non-monotonous, dose-dependent changes in the motility of glioblastoma cells. J. Neurooncol. 67(1–2), 147–157 (2004)

    Article  Google Scholar 

  53. Heinegard, D., Paulsson, M., Inerot, S., Carlstrom, C.: A novel low-molecular weight chondroitin sulphate proteoglycan isolated from cartilage. Biochem. J. 197, 355–366 (1981)

    Article  Google Scholar 

  54. Helmlinger, G., Netti, P., Lichtenbeld, H., Melder, R., Jain, R.: Solid stress inhibits the growth of multicellular tumor spheroids. Nat. Biotechnol. 15(8), 778–783 (1997)

    Article  Google Scholar 

  55. Heyman, I., Faissner, A., Lumsden, A.: Cell and matrix specialisations of rhombomere boundaries. Dev. Dyn. 204(3), 301–315 (1995)

    Article  Google Scholar 

  56. Hines, D., Hines, R., Mulligan, S., Macvicar, B.: Microglia processes block the spread of damage in the brain and require functional chloride channels. Glia 57, 1610–1618 (2009)

    Article  Google Scholar 

  57. von Holst, A., Sirko, S., Faissner, A.: The unique 473HD-chondroitin sulfate epitope is expressed by radial glia and involved in neural precursor cell proliferation. J. Neurosci. 26(15), 4082–4094 (2006)

    Article  Google Scholar 

  58. Jacobs, V., Valdes, P., Hickey, W., Leo, J.D.: current review of in vivo GBM rodent models: emphasis on the CNS-1 tumour model. ASN Neuro 3(3), e00063 (2011)

    Article  Google Scholar 

  59. Jain, R.: Transport of molecules in the tumor interstitium: a review. Cancer Res. 47(12), 3039–3051 (1987)

    Google Scholar 

  60. Jhaveri, S.: Midline glia of the tectum: a barrier for developing retinal axons. Perspect. Dev. Neurobiol. 1(4), 237–243 (1993)

    Google Scholar 

  61. Jin, Y., Ketschek, A., Jiang, Z., Smith, G., Fischer, I.: Chondroitinase activity can be transduced by a lentiviral vector in vitro and in vivo. J. Neurosci. Methods 199(2), 208–213 (2011)

    Article  Google Scholar 

  62. Jue, T., Hovey, E., Davis, S., Carleton, O., McDonald, K.: Incorporation of biomarkers in phase II studies of recurrent glioblastoma. Tumour Biol. 36(1), 153–162 (2015)

    Article  Google Scholar 

  63. Kaufman, L., Brangwynne, C., Kasza, K., Filippidi, E., Gordon, V., Deisboeck, T., Weitz, D.: Glioma expansion in collagen I matrices: analyzing collagen concentration-dependent growth and motility patterns. Biophys. J. BioFAST 89, 635–650 (2005)

    Article  Google Scholar 

  64. Kettenmann, H., Kirchhoff, F., Verkhratsky, A.: Microglia: new roles for the synaptic stripper. Neuron 77(1), 10–18 (2013)

    Article  Google Scholar 

  65. Kettenmann, H., Verkhratsky, A.: Neuroglia-living nerve glue. Fortschritte der Neurologie. Psychiatrie 79, 588–597 (2011)

  66. Khain, E., Sander, L.: Dynamics and pattern formation in invasive tumor growth. Phys. Rev. Lett. 96(18), 188,103 (2006)

    Article  Google Scholar 

  67. Kim, J., Kim, P., Kim, S., Yun, C.: Enhancing the therapeutic efficacy of adenovirus in combination with biomaterials. Biomaterials 33(6), 1838–1850 (2012)

    Article  Google Scholar 

  68. Kim, J., Lee, Y., Kim, H., Huang, J., Yoon, A., Yun, C.: Relaxin expression from tumor targeting adenoviruses and its intra tumoral spread, apoptosis induction, and efficacy. Gene Ther. 98, 1482–1493 (2006)

    Google Scholar 

  69. Kim, Y.: Regulation of cell proliferation and migration in glioblastoma: new therapeutic approach. Front. Mol. Cell. Oncol. 3, 53 (2013)

    Google Scholar 

  70. Kim, Y., Kang, H., Powathil, G., Kim, H., Trucu, D., Lee, W., Lawler, S., Chaplain, M.: Microrna regulation of a cancer network in glioblastoma: the role of miR-451-AMPK-mTOR in regulation of cell proliferation and infiltration. J. R. Soc. Interface (2015). (submitted)

  71. Kim, Y., Lawler, S., Nowicki, M., Chiocca, E., Friedman, A.: A mathematical model of brain tumor: pattern formation of glioma cells outside the tumor spheroid core. J. Theor. Biol. 260, 359–371 (2009)

    Article  MathSciNet  Google Scholar 

  72. Kim, Y., Lee, H., Dmitrieva, N., Kim, J., Kaur, B., Friedman, A.: Choindroitinase ABC I-mediated enhancement of oncolytic virus spread and anti-tumor efficacy: a mathematical model. PLoS One 9(7), e102,499 (2014)

    Article  Google Scholar 

  73. Kim, Y., Powathil, G., Kang, H., Trucu, D., Kim, H., Lawler, S., Chaplain, M.: Strategies of eradicating glioma cells: a multi-scale mathematical model with miR-451-AMPK-mTOR control. PLoS One 10(1), e0114370 (2015)

    Article  Google Scholar 

  74. Kim, Y., Roh, S.: A hybrid model for cell proliferation and migration in glioblastoma. Discret. Contin. Dyn. Syst. B 18(4), 969–1015 (2013)

    Article  MATH  MathSciNet  Google Scholar 

  75. Kim, Y., Roh, S., Lawler, S., Friedman, A.: miR451 and AMPK/MARK mutual antagonism in glioma cells migration and proliferation. PLoS One 6(12), e28,293 (2011)

    Article  Google Scholar 

  76. Kim, Y., Stolarska, M., Othmer, H.: A hybrid model for tumor spheroid growth in vitro I: theoretical development and early results. Math. Model. Methods. Appl. Sci. 17, 1773–1798 (2007)

    Article  MATH  MathSciNet  Google Scholar 

  77. Kim, Y., Stolarska, M., Othmer, H.: The role of the microenvironment in tumor growth and invasion. Prog. Biophys. Mol. Biol. 106, 353–379 (2011)

    Article  Google Scholar 

  78. Lalancette-Hebert, M., Gowing, G., Simard, A., Weng, Y., Kriz, J.: Selective ablation of proliferating microglial cells exacerbates ischemic injury in the brain. J. Neurosci. 27, 2596–2605 (2007)

    Article  Google Scholar 

  79. Lang, B., Cregg, J., Weng, Y., Li, S., Silver, J.: The LAR family of prosynaptic proteins help mediate glial scar induced axonal regeneration failure following spinal cord injury. Soc. Neurosci. Abstr. 37, 892.15 (2011)

    Google Scholar 

  80. Le, D.M., Besson, A., Fogg, D.K., Choi, K.S., Waisman, D.M., Goodyer, C.G., Rewcastle, B., Yong, V.W.: Exploitation of astrocytes by glioma cells to facilitate invasiveness: a mechanism involving matrix metalloproteinase-2 and the urokinase-type plasminogen activator-plasmin cascade. J. Neurosci. 23(10), 4034–4043 (2003)

    Google Scholar 

  81. Leder, K., Pitter, K., Laplant, Q., Hambardzumyan, D., Ross, B., Chan, T., Holland, E., Michor, F.: Mathematical modeling of PDGF-driven glioblastoma reveals optimized radiation dosing schedules. Cell 156(3), 603–616 (2014)

    Article  Google Scholar 

  82. Lee, H., McKeon, R., Bellamkonda, R.: Sustained delivery of thermostabilized ChABC enhances axonal sprouting and functional recovery after spinal cord injury. PNAS 107, 3340–3345 (2010)

    Article  Google Scholar 

  83. Lin, R., Kwok, J., Crespo, D., Fawcett, J.: Chondroitinase ABC has a long-lasting effect on chondroitin sulphate glycosaminoglycan content in the injured rat brain. J. Neurochem. 104, 400–408 (2008)

    Article  Google Scholar 

  84. Lowengrub, J., Frieboes, H., Jin, F., Chuang, Y., Li, X., Macklin, P., Wise, S., Cristini, V.: Nonlinear modelling of cancer: bridging the gap between cells and tumours. Nonlinearity 23, R1 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  85. Machado, C., Schenka, A., Vassallo, J., Tamashiro, W., Gonalves, E., Genari, S., Verinaud, L.: Morphological characterization of a human glioma cell line. Cancer Cell Int. 5(1), 13 (2005)

    Article  Google Scholar 

  86. Markert, J., Liechty, P., Wang, W., Gaston, S., Braz, E., Karrasch, M., Nabors, L., Markiewicz, M., Lakeman, A., Palmer, C., Parker, J., Whitley, R., Gillespie, G.: Phase Ib trial of mutant herpes simplex virus G207 inoculated pre-and post-tumor resection for recurrent GBM. Mol. Ther. 17, 199–207 (2009)

    Article  Google Scholar 

  87. Markovic, D., Vinnakota, K., Chirasani, S., Synowitz, M., Raguet, H., Stock, K., Sliwa, M., Lehmann, S., Kalin, R., van Rooijen, N., Holmbeck, K., Heppner, F., Kiwit, J., Matyash, V., Lehnardt, S., Kaminska, B., Glass, R., Kettenmann, H.: Gliomas induce and exploit microglial MT1-MMP expression for tumor expansion. Proc. Natl. Acad. Sci. USA 106, 12530–12535 (2009)

    Article  Google Scholar 

  88. Mok, W., Boucher, Y., Jain, R.: Matrix metalloproteinases-1 and -8 improve the distribution and efficacy of an oncolytic virus. Cancer Res. 67, 10664–10668 (2007)

    Article  Google Scholar 

  89. Mok, W., Stylianopoulos, T., Boucher, Y., Jain, R.: Mathematical modeling of herpes simplex virus distribution in solid tumors: implications for cancer gene therapy. Clin. Cancer Res. 15(7), 2352–2360 (2009)

    Article  Google Scholar 

  90. Moon, L., Asher, R., Fawcett, J.: Limited growth of severed CNS axons after treatment of adult rat brain with hyaluronidase. J. Neurosci. Res. 71, 23–37 (2003)

    Article  Google Scholar 

  91. Morrison, J., Briggs, S., Green, N., Fisher, K., Subr, V., Ulbrich, K., Kehoe, S., Seymour, L.: Virotherapy of ovarian cancer with polymer-cloaked adenovirus retargeted to the epidermal growth factor receptor. Mol. Ther. 16(2), 244–251 (2008)

    Article  Google Scholar 

  92. Myer, D., Gurkoff, G., Lee, S., Hovda, D., Sofroniew, M.: Essential protective roles of reactive astrocytes in traumatic brain injury. Brain 129(10), 2761–2772 (2006)

    Article  Google Scholar 

  93. Nakajima, M., Nakasu, S., Morikawa, S., Inubushi, T.: Estimation of volume doubling time and cell loss in an experimental rat glioma model in vivo. Acta Neurochir. (Wien) 140(6), 607–612 (1998)

    Article  Google Scholar 

  94. Nimmerjahn, A., Kirchhoff, F., Helmchen, F.: Resting microglial cells are highly dynamic surveillants of brain parenchyma in vivo. Science 308, 1314–1318 (2005)

    Article  Google Scholar 

  95. ODonoghue, J., Bardies, M., Wheldon, T.: Relationships between tumor size and curability for uniformly targeted therapy with beta-emitting radionuclides. J. Nucl. Med. 36(10), 1902–1909 (1995)

    Google Scholar 

  96. Paolicelli, R.C., Bolasco, G., Pagani, F., Maggi, L., Scianni, M., Panzanelli, P., Giustetto, M., Ferreira, T.A., Guiducci, E., Dumas, L., et al.: Synaptic pruning by microglia is necessary for normal brain development. Science 333(6048), 1456–1458 (2011)

    Article  Google Scholar 

  97. Pasquier, B., Pasquier, D., N’Golet, A., Panh, M., Couderc, P.: Extraneural metastases of astrocytomas and glioblastomas: clinicopathological study of two cases and review of literature. Cancer 45(1), 112–25 (1980)

    Article  Google Scholar 

  98. Pham, K., Chauviere, A., Hatzikirou, H., Li, X., Byrne, H., Cristini, V., Lowengrub, J.: Density-dependent quiescence in glioma invasion: instability in a simple reaction-diffusion model for the migration/proliferation dichotomy. J. Biol. Dyn. (2011). doi:10.1080/17513758.2011.590610

  99. Reichert, M., Steinbach, J., Supra, P., Weller, M.: Modulation of growth and radiochemosensitivity of human malignant glioma cells by acidosis. Cancer 95(5), 1113–9 (2002)

    Article  Google Scholar 

  100. Rejniak, K., Anderson, A.: Hybrid models of tumor growth. WIRES Syst. Biol. Med. 3, 115–125 (2011)

    Article  Google Scholar 

  101. Reyes, A.L., Jung, E., Kim, Y.: Optimal control strategies of eradicating invisible glioblastoma cells after conventional surgery. J. R. Soc. Interface 12, 20141,392 (2015)

    Article  Google Scholar 

  102. Rolls, A., Shechter, R., London, A., Segev, Y., Jacob-Hirsch, J., Amariglio, N., Rechavi, G., Schwartz, M.: Two faces of chondroitin sulfate proteoglycan in spinal cord repair: a role in microglia/macrophage activation. PLoS Med. 5, e171 (2008)

    Article  Google Scholar 

  103. Rong, Z., Cheema, U., Vadgama, P.: Needle enzyme electrode based glucose diffusive transport measurement in a collagen gel and validation of a simulation model. Analyst 131(7), 816–21 (2006)

    Article  Google Scholar 

  104. Roth, W., Wild-Bode, C., Platten, M., Grimmel, C., Melkonyan, H., Dichgans, J., Weller, M.: Secreted frizzled-related proteins inhibit motility and promote growth of human malignant glioma cells. Oncogene 19, 4210–4220 (2000)

    Article  Google Scholar 

  105. Saffarian, S., Collier, I., Marmer, B., Elson, E., Goldberg, G.: Interstitial collagenase is a Brownian ratchet driven by proteolysis of collagen. Science 306(5693), 108–111 (2004)

    Article  Google Scholar 

  106. Sander, L., Deisboeck, T.: Growth patterns of microscopic brain tumors. Phys. Rev. E 66, 051901 (2002)

    Article  Google Scholar 

  107. Shirasuna, K., Saka, M., Hayashido, Y., Yoshioka, H., Sugiura, T., Matsuya, T.: Extracellular matrix production and degradation by adenoid cystic carcinoma cells: participation of plasminogen activator and its inhibitor in matrix degradation. Cancer Res. 53(1), 147–152 (1993)

    Google Scholar 

  108. Sierra, A., Beccari, S., Diaz-Aparicio, I., Encinas, J.M., Comeau, S., ve Tremblay, M.: Surveillance, phagocytosis, and inflammation: how never-resting microglia influence adult hippocampal neurogenesis. Neural Plasticity 2014, 610343 (2014)

  109. Silver, D.J., Siebzehnrubl, F.A., Schildts, M.J., Yachnis, A.T., Smith, G.M., Smith, A.A., Scheffler, B., Reynolds, B.A., Silver, J., Steindler, D.A.: Chondroitin sulfate proteoglycans potently inhibit invasion and serve as a central organizer of the brain tumor microenvironment. J. Neurosci. 33(39), 15603–15617 (2013)

    Article  Google Scholar 

  110. Silver, J., Miller, J.H.: Regeneration beyond the glial scar. Nat. Rev. Neurosci. 5(2), 146–156 (2004)

    Article  Google Scholar 

  111. Snow, D.M., Steindler, D.A., Silver, J.: Molecular and cellular characterization of the glial roof plate of the spinal cord and optic tectum: a possible role for a proteoglycan in the development of an axon barrier. Dev. Biol. 138(2), 359–376 (1990)

    Article  Google Scholar 

  112. Stein, A., Demuth, T., Mobley, D., Berens, M., Sander, L.: A mathematical model of glioblastoma tumor spheroid invasion in a three-dimensional in vitro experiment. Biophys. J. 92(1), 356–365 (2007)

    Article  Google Scholar 

  113. Steindler, D.A., O’brien, T.F., Cooper, N.G.: Glycoconjugate boundaries during early postnatal development of the neostriatal mosaic. J. Comp. Neurol. 267(3), 357–369 (1988)

    Article  Google Scholar 

  114. Stolarska, M., Kim, Y., Othmer, H.: Multiscale models of cell and tissue dynamics. Phil. Trans. R. Soc. A 367, 3525–3553 (2009)

    Article  MATH  MathSciNet  Google Scholar 

  115. Stylli, S., Kaye, A., MacGregor, L., Howes, M., Rajendra, P.: Photodynamic therapy of high grade glioma—long term survival. J. Clin. Neurosci. 12(4), 389–398 (2005)

    Article  Google Scholar 

  116. Suarez, C., Maglietti, F., Colonna, M., Breitburd, K., Marshall, G.: Mathematical modeling of human glioma growth based on brain topological structures: Study of two clinical cases. PLoS One 7(6), e39616 (2012)

    Article  Google Scholar 

  117. Swanson, K., Alvord, E., Murray, J.: Virtual resection of gliomas: effect of extent of resection on recurrence. Math. Comput. Model. 37, 1177–1190 (2003)

    Article  MATH  Google Scholar 

  118. Thurber, G., Wittrup, K.: Quantitative spatiotemporal analysis of antibody fragment diffusion and endocytic consumption in tumor spheroids. Cancer Res. 68(9), 3334–3341 (2008)

    Article  Google Scholar 

  119. Thuret, S., Moon, L., Gage, F.: Therapeutic interventions after spinal cord injury. Nat. Rev. Neurosci. 7, 628–643 (2006)

    Article  Google Scholar 

  120. Tremblay, M.È., Stevens, B., Sierra, A., Wake, H., Bessis, A., Nimmerjahn, A.: The role of microglia in the healthy brain. J. Neurosci. 31(45), 16064–16069 (2011)

    Article  Google Scholar 

  121. Trottenberg, U., Oosterlee, C., Schuller: Multigrid. Academic Press, Dublin (2001)

  122. Vanka, S.: Block-implicit multigrid solution of Navier–Stokes equations in primitive variables. J. Comput. Phys. 65, 138–158 (1986)

    Article  MATH  MathSciNet  Google Scholar 

  123. Verheije, M., Lamfers, M., Wurdinger, T., Grinwis, G., Gerritsen, W., van Beusechem, V., Rottier, P.: Coronavirus genetically redirected to the epidermal growth factor receptor exhibits effective antitumor activity against a malignant glioblastoma. J. Virol. 83(15), 7507–7516 (2009)

    Article  Google Scholar 

  124. Viapiano, M., Lawler, S.: Glioma Invasion: Mechanisims and Therapeutic Challenges. Humana Press, New York (2009)

    Google Scholar 

  125. Viapiano, M., Matthews, R.: From barriers to bridges: chondroitin sulfate proteoglycans in neuropathology. Trends Mol. Med. 12, 488–496 (2006)

    Article  Google Scholar 

  126. Viapiano, M.S., Hockfield, S., Matthews, R.: Behab/brevican requires adamts-mediated proteolytic cleavage to promote glioma invasion. J. Neurooncol. 88, 261–272 (2008)

    Article  Google Scholar 

  127. Viapiano, M.S., Bi, W.L., Piepmeier, J., Hockfield, S., Matthews, R.T.: Novel tumor-specific isoforms of behab/brevican identified in human malignant gliomas. Cancer Res. 65(15), 6726–6733 (2005)

    Article  Google Scholar 

  128. Weller, M., Rieger, J., Grimmel, C., Meir, E.V., Tribolet, N.D., Krajewski, S., Reed, J., von Deimling, A., Dichgans, J.: Predicting chemoresistance in human malignant glioma cells: the role of molecular genetic analyses. Int. J. Cancer 79, 640–644 (1998)

    Article  Google Scholar 

  129. Windle, W.F., Chambers, W.W.: Regeneration in the spinal cord of the cat and dog. J. Comp. Neurol. 93(2), 241–257 (1950)

    Article  Google Scholar 

  130. Zhang, T., Fang, H.: Effective diffusion coefficients of glucose in artificial biofilms. Environ. Technol. 26(2), 155–160(6) (2005)

    Article  MathSciNet  Google Scholar 

  131. Zhang, W., Couldwell, W., Simard, M., Song, H., Lin, J., Nedergaard, M.: Direct gap junction communication between malignant glioma cells and astrocytes. Cancer Res. 59(8), 1994–2003 (1999)

    Google Scholar 

  132. Zhao, R., Muir, E., Alves, J., Rickman, H., Allan, A., Kwok, J., Roet, K., Verhaagen, J., Schneider, B., Bensadoun, J., Ahmed, S., Yanez-Munoz, R., Keynes, R., Fawcett, J., Rogers, J.: Lentiviral vectors express chondroitinase ABC in cortical projections and promote sprouting of injured corticospinal axons. J. Neurosci. Methods 201(1), 228–238 (2011)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yangjin Kim.

Additional information

Y. Kim work was supported by the Basic Science Research Program through the National Research Foundation of Korea by the Ministry of Education and Technology (2012R1A1A1043340) and 2014 Faculty research grant at Konkuk University, South Korea.

Appendices

Appendix 1: Parameter estimation

Random motility of invasive glioma cells (\(D_i\)) The experimental results in Hegedus et al. [52] and the experimental data of Demuth et al. [29] suggest a \(D_i\) between \(1.0 \times 10^{-5}\) and \(2.0 \times 10^{-4}\,\mathrm{cm}^2/\mathrm{day}\), which was also validated in the experimental study of motility of glioma cell lines U87\(\Delta \)EGFR and U87WT [112]. We take \(D_i = 1.1 \times 10^{-5}\,\mathrm{cm}^2/\mathrm{day}\) (=4.6\(\times 10^{-5}\,\mathrm{mm}^2/\mathrm{h}\)) within this experimental range for our work.

Diffusion coefficient of nutrient (glucose) (\(D_G\)) While the diffusion coefficient of G in the brain is estimated to be \(6.7 \times 10^{-7}\,\mathrm{cm}^2/\mathrm{s}\) [59], the experiments in collagen gel suggest the diffusion coefficient of G, \(1.3 \times 10^{-6}\,\mathrm{cm}^2/\mathrm{s}\) [103]. Effective diffusion coefficients in artificial biofilms were measured to be \(6.46 \times 10^{-6}\,\mathrm{cm}^2/\mathrm{s}\) [130]. We take \(D_G = 9.16 \times 10^{-7}\,\mathrm{cm}^2/\mathrm{s}\) (=3.3 \(\times 10^{-1}\,\mathrm{mm}^2/\mathrm{h}\)).

Diffusion coefficient of Chase-ABC (\(D_C\)) The diffusion coefficient of typical ECM proteinases such as MMPs is known to be small [71]. For example, the diffusion coefficient of MMP-1 was estimated to be \((8\pm 1.5)\times 10^{-9}\,\mathrm{cm}^2/\mathrm{s}\) for wide-type (activated) MMP-1 and \((6.7\pm 1.5)\times 10^{-9}\,\mathrm{cm}^2/\mathrm{s}\) for inactive MMP-1 [105]. From these experimental measurement, we take \(D_C = 9.17\times 10^{-10} ~cm^2/s\) (=3.3 \(\times 10^{-4}\,\mathrm{mm}^2/\mathrm{h}\)).

Proliferation rate of tumor cells (\(\lambda \)) and Removal rate of necrotic cells (\(\mu \)) A large range of doubling times of glioma cells were reported. For example, the volume doubling time (Vd) of the ethyl-nitrosourea-induced rat glioma by MBI imaging ranged from 3.3 to 29.2 days (\(11.3\pm 7.74\)) and potential doubling time (Tp) determined immunohistochemically ranged from 2.3 to 13.3 days (\(6.81\pm 3.33\)) [93]. In a secreted Frizzled-related proteins study on the motility and growth of human malignant glioma cells, the doubling time of glioma cell lines (U87MG, LN-229, and U373MG) was measured to be in the range of 29–49 h among Neo, sFRP-1, sFRP-2 cases [104]. However, the percentage of dead cells (Neo case) were 22 and 70 % at 96 and 144 h, respectively, in this study [104], indicating the smaller doubling time (larger growth rate). The doubling times of the glioma cell lines were measured to be in the range of 22–39 h (39 h for U87MG, 24 h for LN-229, 22 h for T98G, and 33 h for LN-18 cell lines) [99, 128]. In a morphological characterization study, the doubling time of NG97 glioma cell line was 24 h [85]. In an in vitro experimental study [112] of growing/invasive behaviors of glioma cell lines, U87\(\Delta \)EGFR and U87WT, the doubling time of growing cells in the tumor core was observed to be 20 h, leading to the proliferation rate of \(0.83/\mathrm{day}\) (=3.46 \(\times 10^{-2}/\mathrm{h}\)), but the proliferation rate of the invasive glioma cell lines were estimated to be in the range of 0–0.3/day (=0–1.25 \(\times 10^{-2}/\mathrm{h}\)) for mutant (U87\(\Delta \)EGFR) and 0.04–0.3/day (=(\(1.7\times 10^{-3}-1.25 \times 10^{-2})/\mathrm{h}\)) for the wild type (U87WT). On the other hand, the removal rate of necrotic cells was estimated to be \((1.2 - 2.08) \times 10^{-2}\,\mathrm{h}^{-1}\) in studies of OV therapy in glioblastoma [37, 72]. We take the removal of necrotic cells, \(\mu = 5.7\times 10^{-3}\,\mathrm{h}^{-1}\). The overall growth rate of tumor cells (\(=x+n+x_i\)) can be approximated from the sum of the equations of growing, dead, invasive glioma cells (Eqs. (8), (9), and (23)): \(d(x+n+x_i)/dt = \lambda x(1-x/x_0) - \mu n\). Based these experimental data of doubling time and taking into account the death of cells in the experiments, we take the proliferation rate of growing tumor cells, \(\lambda = 7.33 \times 10^{-2}\,\mathrm{h}^{-1}\).

Sensing radius of microglia for CSPGs (\(r_s\)) A cell can sense its microenvironmental conditions within a neighboring region, which was estimated to be a ball with a sending radius (\(r_s\)) in our work. Kim et al. [71] took the sensing radius of 25 \(\upmu \mathrm{m}\) based on experimental observation of 9 L gliosarcoma cells to C6 glioma cells [131]. In a study of accelerated microglial pathology in Alzheimer’s disease [7], the tree area of was estimated to be 1800 \(\upmu \mathrm{m}^2\), leading to the average radius of 24 \(\upmu \mathrm{m}\) cell trees. In general, microglia were shown to be exceptional sensors of their microenvironment, responding to even subtle changes in their milieu on a short time scale of minutes, by undergoing coordinated changes in gene expression and morphology [1, 64, 108]. By taking into these observations, we take 60 \(\upmu \mathrm{m}\) (=6.0 \(\times 10^{-2}\,\mathrm{mm}\)).

Table 1 Model parameters

ECM production rate from tumor cells (\(\lambda _{42}\)) The remodeling/production rate of ECM was estimated to be \(5.0\times 10^{-6}\,\mathrm{s}^{-1}=0.018\,\mathrm{h}^{-1}\) in a study of the pattern formation of glioma cells [71]. In a study of oncolytic viral therapy on glioma cells [72], the rate was taken to be \(6.0 \times 10^{-3}\,\mathrm{h}^{-1}\). In this work, we take a slightly smaller rate, \(\lambda _{42} = 2.9\times 10^{-3}\,\mathrm{h}^{-1}\), leading to the effective production rate by tumor cells \(\lambda _{42}\frac{x^{m_2}}{K_x^{m_2}+x^{m_2}}=(2.87\times 10^{-5} - 1.4\times 10^{-3})\,\mathrm{h}^{-1}\) by assuming cell density \(x=(10^4-10^6)\,\mathrm{cells}/\mathrm{mm}^3\) [37, 72, 95] and \(K_x=x^*, m_2=1\) in Table 1, and \(\lambda _{42}\) above.

CSPG degradation rate (\(\lambda _{41}\)) Kim et al. [71] used \(3.0\times 10^8\,\mathrm{cm}^3/(\mathrm{g}\cdot \mathrm{s})\) for the MMP-mediated degradation rate of the ECM (mainly collagen) by tumor cells. The degradation rate of ECM has a wild range. For instance, there were only 36.6, 24.5, 38.9 % reduction of ECM from degradation by trypsin, elastase, collagenase [107] over the long period of time, 5 days, leading to very small values of the degradation rate. In a study of the effect of Chase-ABC on CSPG content in the injured rate brain [83], Lin et al. found that there was an approximately 50 % decrease in GAG content (main constituent of CSPGs (E)) compared to normal tissue after 1 h of Chase-ABC treatment, indicating the approximated decay rate of 0.6931 \(\mathrm{h}^{-1}\). In a study of components of perineuronal net ECM in rat brain [26], Deepa et al. found that various CSPGs act differently with Chase-ABC treatment for 1 hour: neurocan and versican completely removed (large degradation rate \(>\) 10 h\(^{-1}\)) but aggrecan and phosphacan were not removed (very small degradation rate) while some others are partly removed. On the other hand, the rapid digestion of CSPGs followed by OV spread was observed in the experimental study of Chase-ABC in OV treatment of gliomas [30]. Based on the experimental data [30], the Chase-ABC-mediated degradation rate of CSPGs was estimated to be \(\lambda _{41} = 9.0\times 10^{1}\,\mathrm{h}^{-1}\) in a Chase-ABC-assisted OV spread study [72], which generated good agreement between the in vitro experiment and the simulation results. In order to take into account the CSPG-dominant compact microenvironment in the in vivo system in our model, we take a slightly smaller value \(\lambda _{41} =4.3 \times 10^{1}\,\mathrm{h}^{-1}\) in this work. By assuming \(C=(0.5-250)\,\mathrm{mU}/\mathrm{ml}\) [12, 49, 72, 83], \(K_C = C^*\), and \(m_1=1\) in Table 1, our choice of \(\lambda _{41}\) above leads to the effective CSPG degradation rate \(\lambda _{41}\frac{C^{m_1}}{K_C^{m_1}+C^{m_1}} =(0.42 - 35.8)\,\mathrm{h}^{-1}\).

Natural decay rate of Chase-ABC (\(\lambda _{62}\)) The natural decay rate of MMPs was estimated to be \(5.0\times 10^{-6}~s^{-1}=1.8 \times 10^{-2}\,\mathrm{h}^{-1}\) [71]. The half-life of Chase-ABC (at 37 \(^\circ \)C) varies significantly [132]; The relatively short half-life (6 days) was reported in the brain of the injected rat [83] but it may be 2–3 weeks with albumin or trehalose (stabilizers) [16, 82]. In the simulation, assuming the half-life of 230 h, we take \(\lambda _{62} = ln(2)/(230\,\mathrm{h}) = 6.0\times 10^{-3}\,\mathrm{h}^{-1}\).

Threshold values (\(G_{0}\)) In vitro experiments suggest that tumor cells enter quiescent/necrotic state in response to nutrient withdrawal [lower than 40 % of normal glucose levels (25  mM) [14, 54] and 25 % of normal oxygen levels (0.2 mM) [14, 36]] due to growing mass of the tumor [76]. For the transition of tumor cells to hypoxic/necrotic cells, we take \(G_{0} = 1.35 \times 10^{-3}\,\mathrm{g}/\mathrm{cm}^3\).

Chemotactic sensitivity of tumor cells (\(\chi _i\)) Glioma cells may travel a distance 0.4–0.5 cm in 150 h with EGF treatment [18]; the cells was shown to travel faster, up to a distance of 1.25 cm (0.75  cm) in 150 h in agar containing EGF (plain agar). The cell velocity was in the range of of 50–110 \(\upmu \mathrm{m}/24\,\mathrm{h}\) in U87MGmEGFR spheroids experiments [27]. In a study of miR-451-mediated cell migration and proliferation in response to high and low glucose levels in a glioma, Kim et al. [75] took the chemotactic sensitivity, \(1.86\times 10^{-7}\,\mathrm{cm}^5/(\mathrm{g}\cdot \mathrm{s})\), to glucose gradients in a slightly modified form with a upper bound. Assuming that gradient of the glucose concentration was in the range of \((10^{-4} - 3\times 10^{-2})\,\mathrm{g}/\mathrm{cm}^4\) and using drift velocity 25–110 \(\upmu \mathrm{m}/24\,\mathrm{h}\) of migratory glioma cells, we compute \(\chi _i=\frac{velocity}{gradient}\) to be \(\chi _n\)= (\(1.9\times 10^{-6} - 1.3\times 10^{-3} )\,\mathrm{cm}^5/(\mathrm{g}\cdot \mathrm{s})\). In this work, we take \(\chi _i= 2.9\times 10^{-6}\,\mathrm{cm}^5/(\mathrm{g}\cdot \mathrm{s})\) (=\(1.04\times 10^{3}\) \(\mathrm{mm}^5 \mathrm{g}^{-1}\mathrm{h}^{-1}\)).

Appendix 2: Nondimensionalization

Table 2 lists the reference values in the model. We define \(L=4\,\mathrm{mm}\) following the experimental setup in [109] and take the characteristic diffusion coefficient \(D=1.27\times 10^{-5}\,\mathrm{cm}^2/\mathrm{s}\) so that \(T = 3.5\,\mathrm{h}\). We determine the reference values for \(x,n,A,M_r, M_a, E, \rho , C\) as follows:

Cell density (\(x^*, n^*, A^*,M_r^* = M_a^*\)): various cell densities can be used in the experiments and modeling: \(5\times 10^5\,\mathrm{cells}/\mathrm{cm}^3\) [18], \(7.23 \times 10^8\,\mathrm{cells}/\mathrm{ml}\) [89, 118], \(10^9\,\mathrm{cells}/\mathrm{cm}^3\) [37, 72, 95]. In this study, we take \(x^* = n^* = 10^6\,\mathrm{cells}/\mathrm{mm}^3\) from [37, 72, 95]. Density of fibrous astrocytes was estimated to be in the range of \(2.0\times 10^5\,\mathrm{cells}/\mathrm{mm}^3\) in the brain [65]. We take \(A^* = 2.0\times 10^5\,\mathrm{cells}/\mathrm{mm}^3\). Density of microglia was measured to be in the range of (0.5–2.5) \(\times 10^{-5}\,\mathrm{cells}/(\upmu \mathrm{m}^3) =\) 0.5–2.5 \(\times 10^4\,\mathrm{cells}/(\mathrm{mm}^3)\) in response to Cx3cr1GFP/Cx3cr1KO reporter knockout in a study of synaptic pruning by microglia in normal brain development [96, 120]. We take \(M_r^* = M_a^* = 1.0 \times 10^4\,\mathrm{cells}/(\mathrm{mm}^3)\).

Table 2 Reference value

ECM density [CSPG (\(E^*\)), Tumor ECM (\(\rho ^*\))] In a study of cell infiltration in glioma, Silver et al. [109] investigated the role of various CSPG concentrations (0–500 \(\upmu \mathrm{g}/\mathrm{ml}\)) and laminin concentration (5 \(\upmu \mathrm{g}/\mathrm{ml}\)) in regulation of glioma infiltration using both invasive and non-invasive co-culture assays. They found that while high concentrations of CSPGs (250, 500 \(\upmu \mathrm{g}/\mathrm{ml}\)) induce non-invasive glioma with astrocytes on the periphery of the well-defined tumor core, low levels of CSPG (0–10 \(\upmu \mathrm{g}/\mathrm{ml}\)) can allow glioma cell infiltration in the presence of astrocytes in the tumor aggregates [109]. It was shown that Cat-301 CNS CSPG with high molecular weight from brain in the density of 1.4 g/ml [38] can have similar biological properties as aggrecan, CSPG with the high molecular weight from cartilage in the density of 1.35–1.4 g/ml [53]. Dutt et al. [31, 32] investigated effects of versican, one of major CSPG components isolated from human glioma cell line (U251MG) and others (0–100 \(\upmu \mathrm{g}/\mathrm{ml}\)), on neural crest cell migration. They found that low concentrations of version V0/V1 (\(\sim \)25 \(\upmu \mathrm{g}/\mathrm{ml}\)) can block active migration of neural crest stem cells [31, 32]. The expression level of noncleaved isoform of Brevican/BEHAB was \(>\)4-fold higher in human gliomas compared to normal brain tissue [127]. Isolated versican levels in brain tissue were estimated to be 3 mg/100 g wet tissue components [9]. Glial HA-binding protein (GHAP) usually localized in white matter is present in the concentration of 8.2 mg/100 g in human white matter [9]. We take \(E^*=250\,\upmu \mathrm{g}/\mathrm{cm}^3\) from [109]. In a OV spread study [72], the tumor ECM concentration was estimated to be \(1.0\,\mathrm{mg}/\mathrm{cm}^3\) based on the experimental observation in the brain [63, 71, 112]. We take \(\rho ^*=1.0\,\mathrm{mg}/\mathrm{cm}^3\) from [72].

Concentration of nutrients (glucose) (\(G^*\)): Sander and Deisboeck [106] used the characteristic glucose concentration \(2\times 10^{-4}\,\mathrm{g}/\mathrm{cm}^3\), and the boundary condition \(6\times 10^{-4}\,\mathrm{g}/\mathrm{cm}^3\) (cf. [27]). Kim et al. [71] took \(G^* = 6\times 10^{-4}\,\mathrm{g}/\mathrm{cm}^3\) as a reference value in a study of pattern formation of glioma cells in response to glucose withdrawal in vitro experiments. The miR-451 expression was up-regulated in response to a high glucose concentration (\(4.5\times 10^{-3}\,\mathrm{g}/\mathrm{cm}^3\)) while the expression level of its counterpart, AMPK complex, was up-regulated when the tumor was exposed to glucose withdrawal condition (\(1.0\times 10^{-4}\,\mathrm{g}/\mathrm{cm}^3\)) in experimental studies of miR-451 regulation in glioblastoma [43, 44]. We take \(G^* = 4.5\times 10^{-3}\,\mathrm{g}/\mathrm{cm}^3\) from a study of the miR-451-AMPK core control in regulation of cell migration and proliferation in GBM [75].

Concentration of Chase-ABC (\(C^*\)): High-dose infusions of Chase (\(\sim \)1000 U/ml) is necessary for delivery of Chase-ABC into deep regions of the spinal cord due to overflow and dilution beyond the intrathecal space [82]. A low (2 U/0.5 ml) dose of trehalose-assisted Chase was enough to digest CSPG decorin for functional recovery after injury [82]. A wide range of Chase ABC levels (0.5–50 mU/ml) were used in addition to 20 ng/ml EGF and 20 ng/ml bFGF in a study of biochemical effect of Chase-ABCs on the morphological changes in neural precursor cells [49]. Bruckner et al. found that a low dose of \(0.25\,\mathrm{U}/\upmu \mathrm{l}\) of Chase-ABCs was enough for significant breakdown of CSPGs in the rat brain, leading to long-lasting and acute changes in CSPGs [12]. The protease-free Chase-ABC treatment (50 U/ml) led to digestion of CSPGs around the injection regions (resulting in low GAG content (1200–2000 \(\upmu \mathrm{g}/\mathrm{mg}\)) and high content (\(4000\,\upmu \mathrm{g}/\mathrm{mg}\)) with penicillinase-treatment), promoting axon regeneration [83]. In a OV spread study in glioma [72], Kim et al. took 50 mU/ml for the reference value of Chase-ABC. We take \(C^* = 50\,\mathrm{mU}/\mathrm{ml}\).

We nondimensionalize the variables and parameters in the partial differential Eqs. (8)–(38) as follows:

$$\begin{aligned}&\bar{t} = \frac{t}{T}, ~\bar{{\mathbf x}} = \frac{{\mathbf x}}{L}, ~\bar{x} = \frac{x}{x^*}, ~\bar{n} = \frac{n}{n^*}, ~\bar{x}_i = \frac{x_i}{x_i^*}, ~\bar{E}=\frac{E}{E^*}, ~\bar{\rho }=\frac{\rho }{\rho ^*}, ~\bar{M}_r=\frac{M_r}{M_r^*}, \nonumber \\&~\bar{M}_a=\frac{M_a}{M_a^*}, ~\bar{A} = \frac{A}{A^*}, ~\bar{G}=\frac{C}{G^*}, ~\bar{C}=\frac{C}{C^*}, ~\bar{D}_{i} = \frac{D_{i}}{D}, ~\bar{D}_G = \frac{D_G}{D}, ~\bar{D}_C = \frac{D_C}{D}, \nonumber \\&~\bar{\lambda } = T \lambda , ~\bar{x}_0 = \frac{\bar{x}_0}{x^*}, ~\bar{\lambda }_{11}= T \lambda _{11}, ~\bar{\lambda }_{11}^\dagger = T \lambda _{11}\frac{x^*}{n^*}, ~\bar{\mu } = T \mu , ~\bar{\lambda }_{41} = T \lambda _{41}, \\&~\bar{K}_C = \frac{K_C}{C^*}, ~\bar{\lambda }_{42} = T \lambda _{42}, ~\bar{K}_x = \frac{K_x}{x^*}, \bar{\rho }_0 = \frac{\bar{\rho }_0}{\rho ^*}, \bar{k}_{1b} = k_{1b}\frac{E^*}{M_r^*}, \bar{\mu }_G = T \mu _G x^*, \nonumber \\&~\bar{\lambda }_{63} = T E^* \lambda _{63}, ~\bar{\lambda }_{62} = T \lambda _{62}, ~\bar{\chi }_i = \frac{T \chi _i G^*}{L^2}. \nonumber \end{aligned}$$
(39)

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lee, H.G., Kim, Y. The role of the microenvironment in regulation of CSPG-driven invasive and non-invasive tumor growth in glioblastoma. Japan J. Indust. Appl. Math. 32, 771–805 (2015). https://doi.org/10.1007/s13160-015-0188-2

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s13160-015-0188-2

Keywords

Mathematics Subject Classification

Navigation