Skip to main content
Log in

Chemotactic signaling, microglia, and Alzheimer’s disease senile plaques: Is there a connection?

  • Published:
Bulletin of Mathematical Biology Aims and scope Submit manuscript

Abstract

Chemotactic cells known as microglia are involved in the inflammation associated with pathology in Alzheimer’s disease (AD). We investigate conditions that lead to aggregation of microglia and formation of local accumulations of chemicals observed in AD senile plaques. We develop a model for chemotaxis in response to a combination of chemoattractant and chemorepellent signaling chemicals. Linear stability analysis and numerical simulations of the model predict that periodic patterns in cell and chemical distributions can evolve under local attraction, long-ranged repulsion, and other constraints on concentrations and diffusion coefficients of the chemotactic signals. Using biological parameters from the literature, we compare and discuss the applicability of this model to actual processes in AD.

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.

Similar content being viewed by others

References

  • Alt, W. (1980). Biased random walk models for chemotaxis and related diffusion approximations. J. Math. Biol. 9, 147–177.

    Article  MATH  MathSciNet  Google Scholar 

  • Bagnard, D., N. Thomasset, M. Lohrum, A. W. Püschel and J. Bolz (2000). Spatial distributions of guidance molecules regulate chemorepulsion and chemoattraction of growth cones. J. Neurosci. 20, 1030–1035.

    Google Scholar 

  • Banati, R. B. and K. Beyreuther (1995). Alzheimer’s disease, in Neuroglia, Chapter 68, H. Kettenmann and B. R. Ransom (Eds), New York: Oxford University Press, pp. 1027–1043.

    Google Scholar 

  • Benjamin, D., S. Wormsley and S. Dower (1990). Heterogeneity in interleukin (IL)-1 receptors expressed on human B Cell lines. Differences in the molecular properties of IL-1 alpha and IL-1 beta binding sites. J. Biol. Chem. 265, 9943–9951.

    Google Scholar 

  • Benveniste, E. N. (1995). Cytokine production, in Neuroglia, Chapter 5, H. Kettenmann and B. R. Ransom (Eds), New York: Oxford University Press, pp. 700–713.

    Google Scholar 

  • Chicoine, M. R., C. L. Madsen and D. L. Silbergeld (1995). Modification of human glioma locomotion in vitro by cytokines EGF, bFGF, PDGFbb, NGF and TNF-α. Neurosurgery 36, 1165–1171.

    Google Scholar 

  • Cowley, G. (2000). Alzheimer’s: unlocking the mystery. Newsweek January, 46–51.

  • Cross, M. and P. Hohenberg (1993). Pattern formation outside of equilibrium. Rev. Modern Phys. 65, 851–1112.

    Article  Google Scholar 

  • Davis, J. B., H. F. McMurray and D. Schubert (1992). The amyloid beta-protein of Alzheimer’s disease is chemotactic for mononuclear phagocytes. Biochem. Biophys. Res. Commun. 189, 1096–1100.

    Article  Google Scholar 

  • Davis, J. N. and J. C. Chisholm (1997). The ‘amyloid cascade hypothesis’ of AD: decoy or real McCoy? Trends Neurosci. 20, 558–559.

    Article  Google Scholar 

  • de Castro, F., L. Hu, H. Drabkin, C. Sotelo and A. Chédotal (1999). Chemoattraction and chemorepulsion of olfactory bulb axon by different secreted semaphorins. J. Neurosci. 19, 4428–4436.

    Google Scholar 

  • Dickson, D. W. (1997). The pathogenesis of senile plaques. J. Neuropathol. Exp. Neurol. 56, 321–339.

    Google Scholar 

  • Ding, A., E. Sanchez, S. Srimal and C. Nathan (1989). Macrophages rapidly internalize their tumor necrosis factor receptors in response to bacterial lipopolysaccharide. J. Biol. Chem. 264, 3924–3929.

    Google Scholar 

  • Edelstein-Keshet, L. and A. Spiros (2002). Exploring the formation of Alzheimer’s disease senile plaques in silico. J. Theor. Biol. 216, 301–326.

    Article  Google Scholar 

  • Farrell, B., R. Daniele and D. Lauffenburger (1990). Quantitative relationships between single-cell and cell-population model parameters for chemosensory migration responses of alveolar macrophages to C5a. Cell Motil. Cytoskeleton 16, 279–293.

    Article  Google Scholar 

  • Fiala, M. et al. (1998). Amyloid-β induces chemokine secretion and monocyte migration across a human blood-brain barrier model. Mol. Med. 4, 480–489.

    Google Scholar 

  • Goodhill, G. J. (1997). Diffusion in axon guidance. Eur. J. Neurosci. 9, 1414–1421.

    Article  Google Scholar 

  • Grindrod, P., J. Murray and S. Sinha (1989). Steady-state spatial patterns in a cell-chemotaxis model. IMA J. Math. Appl. Med. Biol. 6, 69–79.

    MathSciNet  MATH  Google Scholar 

  • Grunbaum, D. (1994). Translating stochastic density-dependent individual behavior to a continuum model of animal swarming. J. Math. Biol. 33, 139–161.

    Article  Google Scholar 

  • Grunbaum, D. (1999). Advection-diffusion equations for generalized tactic searching behaviors. J. Math. Biol. 38, 169–194.

    Article  MathSciNet  Google Scholar 

  • Hammacher, A., R. Simpson and E. Nice (1996). The interleukin-6 (IL-6) partial antagonist (Q159E, T162P)IL-6 interacts with the IL-6 receptor and gp130 but fails to induce a stable hexameric receptor complex. J. Biol. Chem. 271, 5464–5473.

    Article  Google Scholar 

  • Hardy, J. (1997). Amyloid, the presenilins and Alzheimer’s disease. Trends Neurosci. 20, 154–159.

    Article  Google Scholar 

  • Hillen, T. and A. Stevens (2000). Hyperbolic models for chemotaxis in 1D. Nonlinear Anal. Real World Appl. 1, 409–433.

    Article  MathSciNet  MATH  Google Scholar 

  • Huang, W., Y. Ren and R. Russell (1994a). Moving mesh methods based upon moving mesh partial differential equations. J. Comput. Phys. 113, 279–290.

    Article  MathSciNet  MATH  Google Scholar 

  • Huang, W., Y. Ren and R. Russell (1994b). Moving mesh partial differential equations (mmpdes) based on the equidistribution principle. SIAM J. Numer. Anal. 31, 709–730.

    Article  MathSciNet  MATH  Google Scholar 

  • Igoshin, O., A. Mogilner, R. Welsch, D. Kaiser and G. Oster (2001). Pattern formation and traveling waves in myxobacteria: theory and modeling. Proc. Natl Acad. Sci. USA 98, 14913–14918.

    Google Scholar 

  • Itagaki, S., P. L. McGeer, H. Akiyama, S. Zhu and D. Selkoe (1989). Relationship of microglia and astrocytes to amyloid deposits of Alzheimer’s disease. J. Neuroimmunol. 24, 173–182.

    Article  Google Scholar 

  • Jones, N. (2000). Soothing the inflamed brain. Sci. Am. June, 11–12.

  • Keller, E. F. and L. A. Segel (1970). Initiation of slime-mold aggregation viewed as an instability. J. Theor. Biol. 26, 399–415.

    Article  Google Scholar 

  • Kowall, N. W. (1994). Beta amyloid neurotoxicity and neuronal degeneration in Alzheimer’s disease. Neurobiol. Aging 15, 257–258.

    Article  Google Scholar 

  • Lauffenburger, D. A. and C. Kennedy (1983). Localized bacterial infection in a distributed model for tissue inflammation. J. Math. Biol. 16, 141–163.

    Article  MATH  Google Scholar 

  • Lee, C., M. Hoopes, J. Diehl, W. Gilliland, G. Huxel, E. v. Leaver, K. McCann, J. Umbanhowar and A. Mogilner (2001). Non-local concepts and models in biology. J. Theor. Biol. 210, 201–219.

    Article  Google Scholar 

  • Lee, S., W. Liu, D. Dickson, C. Brosnan and J. Berman (1993). Cytokine production by human fetal microglia and astrocytes. Differential induction by lipopolysaccharide and IL-1. J. Immunol. 150, 2659–2667.

    Google Scholar 

  • Mackenzie, I. R., C. Hao and D. G. Munoz (1995). Role of microglia in senile plaque formation. Neurobiol. Aging 16, 797–804.

    Article  Google Scholar 

  • Maini, P., M. Myerscough, K. H. Winters and J. Murray (1991). Bifurcating spatially heterogeneous solutions in a chemotaxis model for biological pattern generation. Bull. Math. Biol. 53, 701–719.

    Article  MATH  Google Scholar 

  • Mark, M. D., M. Lohrum and A. W. Püschel (1997). Patterning neuronal connections by chemorepulsion: the semaphorins. Cell Tissue Res. 290, 299–306.

    Article  Google Scholar 

  • Mazel, T., Z. Simonova and E. Sykova (1998). Diffusion heterogeneity and anisotropy in rat hippocampus. NeuroReport 9, 1299–1304.

    Google Scholar 

  • McLean, C., R. Cherny, F. Fraser, S. Fuller, M. Smith, K. Beyreuther, A. Bush and C. Masters (1999). Soluble pool of β-amyloid as a determinant of severity of neurodegeneration in Alzheimer’s disease. Ann. Neurobiol. 46, 860–866.

    Article  Google Scholar 

  • Michishita, M., Y. Yoshida, H. Uchino and K. Nagata (1990). Induction of tumor necrosis factor-α and its receptors during differentiation in myeloid leukemic cells along the monocytic pathway. a possible regulatory mechanism for TNF-α production. J. Biol. Chem. 265, 8751–8759.

    Google Scholar 

  • Moghe, P. V., R. D. Nelson and R. T. Tranquillo (1995). Cytokine-stimulated chemotaxis of human neutrophils in a 3-D conjoined fibrin gel assay. J. Immunol. Methods 180, 193–211.

    Article  Google Scholar 

  • Mogilner, A. and L. Edelstein-Keshet (1995). Selecting a common direction: I. how orientational order can arise from simple contact responses between interacting cells. J. Math. Biol. 33, 619–660.

    Article  MathSciNet  MATH  Google Scholar 

  • Mrak, R. E., J. G. Sheng and W. S. Griffin (2000). Glial cytokines in neurodegenerative conditions, in Neuro-Immune Interactions in Neurologic and Psychiatric Disorders, Patterson, Kordon and Christen (Eds), Berlin: Springer, pp. 9–17.

    Google Scholar 

  • Mrak, R. E., J. G. Sheng and W. S. T. Griffin (1995). Glial cytokines in Alzheimer’s disease: review and pathogenic implications. Human Pathology 26, 816–823.

    Article  Google Scholar 

  • Murray, J. D. (1993). Mathematical Biology, 2nd edn, Berlin, Heilderberg: Springer.

    MATH  Google Scholar 

  • Myerscough, M., P. Maini and K. Painter (1998). Pattern formation in a generalized chemotactic model. Bull. Math. Biol. 60, 1–26.

    Article  MATH  Google Scholar 

  • Nash, J. M. (2000). The new science of Alzheimer’s. Time July, 32–39.

  • Nicholson, C. and E. Sykova (1998). Extracellular space structure revealed by diffusion analysis. Trends Neurosci. 21, 207–215.

    Article  Google Scholar 

  • Nilsson, L., J. Rogers and H. Potter (1998). The essential role of inflammation and induced gene expression in the pathogenic pathway of Alzheimer’s disease. Front. Biosci. 3, 436–446.

    Google Scholar 

  • Nolte, C., T. Moller, T. Walter and H. Kettenmann (1996). Complement 5a controls motility of murine microglial cells in vitro via activation of an inhibitory G-protein and the rearrangement of the actin cytoskeleton. Neurosci. Sci. 73, 1109–1120.

    Article  Google Scholar 

  • Othmer, H. G., S. R. Dunbar and W. Alt (1988). Models of dispersal in biological systems. J. Math. Biol. 26, 263–298.

    Article  MathSciNet  MATH  Google Scholar 

  • Othmer, H. G. and A. Stevens (1997). Aggregation, blowup, and collapse: the ABC’s of taxis in reinforced random walks. SIAM J. Appl. Math. 57, 1044–1081.

    Article  MathSciNet  MATH  Google Scholar 

  • Painter, K., P. Maini and H. Othmer (1999). Stripe formation in juvenile pomacanthus explained by a generalized Turing mechanism with chemotaxis. Proc. Natl Acad. Sci. USA 96, 5549–5554.

    Article  Google Scholar 

  • Painter, K., P. Maini and H. Othmer (2000). Development and applications of a model for cellular response to multiple chemotactic cues. J. Math. Biol. 41, 285–314.

    Article  MathSciNet  MATH  Google Scholar 

  • Pennica, D., V. Lam, N. Mize, R. Weber, M. Lewis, B. Fendly, M. Lipari and D. Goeddel (1992). Biochemical properties of the 75-kDa tumor necrosis factor receptor. Characterization of ligand binding, internalization, and receptor phosphorylation. J. Biol. Chem. 267, 21172–21178.

    Google Scholar 

  • Rapoport, M., H. N. Dawes, L. I. Binder, M. P. Vitek and A. Ferriera (2002). Tau is essential to β-amyloid induced neurotoxicity. PNAS 99, 6364–6369.

    Article  Google Scholar 

  • Rivero, M. A., R. T. Tranquillo, H. M. Buettner and D. A. Lauffenburger (1989). Transport models for chemotactic cell populations based on individual cell behavior. Chem. Engineering Sci. 44, 2881–2897.

    Article  Google Scholar 

  • Sager, B. and D. Kaiser (1994). Intercellular c-signalling and the travelling waves of myxococcus. Genes Development 8, 2793–2804.

    Google Scholar 

  • Schaaf, R. (1985). Stationary solutions of chemotaxis systems. Trans. Am. Math. Soc. 292, 531–556.

    Article  MATH  MathSciNet  Google Scholar 

  • Selkoe, D. J. (1991). Amyloid protein and Alzheimer’s disease. Sci. Am. 265, 68–78.

    Article  Google Scholar 

  • Sheng, J. G., X. Q. Zhou, R. E. Mrak and W. S. T. Griffin (1998). Progressive neuronal injury associated with amyloid plaque formation in alzheimer disease. J. Neuropathol. Exp. Neurol. 57, 714–717.

    Google Scholar 

  • Sherratt, J. (1994). Chemotaxis and chemokinesis in eukaryotic cells: the Keller-Segel approximation to a detailed model. Bull. Math. Biol. 56, 129–146.

    Article  MATH  Google Scholar 

  • Sherratt, J., E. Sage and J. Murray (1992). Chemical control of eukaryotic cell movement: a new model. J. Theor. Biol. 162, 23–40.

    Article  Google Scholar 

  • Shi, W. and D. R. Zusman (1994). Sensory adaptation during negative chemotaxis in Myxococcus xanthus. J. Bacteriol. 176, 1517–1520.

    Google Scholar 

  • Smits, H., L. A. Boven, C. Pereira, J. Verhoef and H. S. L. M. Nottet (2000). Role of macrophage activation in the pathogenesis of Alzheimer’s disease and human immunodeficiency virus type I-associated dementia. Eur. J. Clin. Invest. 30, 526–535.

    Article  Google Scholar 

  • Stalder, M., A. Phinney, A. Probst, B. Sommer, M. Staufenbiel and M. Jucker (1999). Association of microglia with amyloid plaques in brains of APP23 transgenic mice. Am. J. Pathol. 154, 1673–1684.

    Google Scholar 

  • Streit, W. J. (1995). Microglia cells, in Neuroglia, Chapter 5, H. Kettenmann and B. R. Ransom (Eds), New York: Oxford University Press, pp. 85–96.

    Google Scholar 

  • Sykova, E. (1997). The extracellular space in the CNS: Its regulation, volume, and geometry in normal and pathological neuronal function. The Neuroscientist 3, 28–41.

    Google Scholar 

  • Sykova, E., T. Mazel and Z. Simonova (1998). Diffusion constraints and neuron-glia interaction during aging. Exp. Gerontol. 33, 837–851.

    Article  Google Scholar 

  • Tranquillo, R., S. Zigmond and D. Lauffenburger (1988). Measurement of the chemotaxis coefficient for human neutrophils in the under-agarose migration assay. Cell Motil. Cytoskeleton 11, 1–15.

    Article  Google Scholar 

  • Turing, A. M. (1952). The chemical basis of morphogenesis. Phil. Trans. R. Soc. B237, 37–72.

    Google Scholar 

  • Venters, H. D., R. Dantzer and K. W. Kelley (2000). A new concept in neurodegeneration: TNF-α is a silencer of survival signals. Trends Neurosci. 23, 175–180.

    Article  Google Scholar 

  • Yamaguchi, M., M. Michishita, K. Hirayoshi, K. Yasukawa, M. Okuma and K. Nagata (1992). Down-regulation of interleukin 6 receptors of mouse myelomonocytic leukemic cells by leukemia inhibitory factor. J. Biol. Chem. 267, 22035–22042.

    Google Scholar 

  • Yao, J., L. Harvath, D. L. Gilbert and C. A. Colton (1990). Chemotaxis by a CNS macrophage, the microglia. J. Neurosci. Res. 27, 36–42.

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Magdalena Luca.

Additional information

Reprint address.

Maternity leave.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Luca, M., Chavez-Ross, A., Edelstein-Keshet, L. et al. Chemotactic signaling, microglia, and Alzheimer’s disease senile plaques: Is there a connection?. Bull. Math. Biol. 65, 693–730 (2003). https://doi.org/10.1016/S0092-8240(03)00030-2

Download citation

  • Received:

  • Accepted:

  • Issue Date:

  • DOI: https://doi.org/10.1016/S0092-8240(03)00030-2

Keywords

Navigation