Generic placeholder image

Current Alzheimer Research

Editor-in-Chief

ISSN (Print): 1567-2050
ISSN (Online): 1875-5828

Review Article

Brain Arrhythmias Induced by Amyloid Beta and Inflammation: Involvement in Alzheimer’s Disease and Other Inflammation-related Pathologies

Author(s): Fernando Peña-Ortega*

Volume 16, Issue 12, 2019

Page: [1108 - 1131] Pages: 24

DOI: 10.2174/1567205017666191213162233

Price: $65

Abstract

A variety of neurological diseases, including Alzheimer’s disease (AD), involve amyloid beta (Aβ) accumulation and/or neuroinflammation, which can alter synaptic and neural circuit functions. Consequently, these pathological conditions induce changes in neural network rhythmic activity (brain arrhythmias), which affects many brain functions. Neural network rhythms are involved in information processing, storage and retrieval, which are essential for memory consolidation, executive functioning and sensory processing. Therefore, brain arrhythmias could have catastrophic effects on circuit function, underlying the symptoms of various neurological diseases. Moreover, brain arrhythmias can serve as biomarkers for a variety of brain diseases. The aim of this review is to provide evidence linking Aβ and inflammation to neural network dysfunction, focusing on alterations in brain rhythms and their impact on cognition and sensory processing. I reviewed the most common brain arrhythmias characterized in AD, in AD transgenic models and those induced by Aβ. In addition, I reviewed the modulations of brain rhythms in neuroinflammatory diseases and those induced by immunogens, interleukins and microglia. This review reveals that Aβ and inflammation produce a complex set of effects on neural network function, which are related to the induction of brain arrhythmias and hyperexcitability, both closely related to behavioral alterations. Understanding these brain arrhythmias can help to develop therapeutic strategies to halt or prevent these neural network alterations and treat not only the arrhythmias but also the symptoms of AD and other inflammation-related pathologies.

Keywords: Amyloid, inflammation, oscillations, theta, gamma, epileptiform.

[1]
Qiu C, Kivipelto M, von Strauss E. Epidemiology of Alzheimer’s disease: occurrence, determinants, and strategies toward intervention. Dialogues Clin Neurosci 11(2): 111-28. (2009).
[2]
Ittner LM, Götz J. Amyloid-β and tau--a toxic pas de deux in Alzheimer’s disease. Nat Rev Neurosci 12(2): 65-72. (2011).
[3]
Heneka MT, Kummer MP, Latz E. Innate immune activation in neurodegenerative disease. Nat Rev Immunol 14(7): 463-77. (2014).
[4]
Heneka MT, Golenbock DT, Latz E. Innate immunity in Alzheimer’s disease. Nat Immunol 16(3): 229-36. (2015).
[5]
Mondragón-Rodríguez S, Perry G, Peña-Ortega F, Williams S. Tau, amyloid beta and deep brain stimulation: aiming to restore cognitive deficit in Alzheimer’s disease. Curr Alzheimer Res 14(1): 40-6. (2017).
[6]
Lue LF, Kuo YM, Roher AE, Brachova L, Shen Y, Sue L, et al. Soluble amyloid beta peptide concentration as a predictor of synaptic change in Alzheimer’s disease. Am J Pathol 155(3): 853-62. (1999).
[7]
Näslund J, Haroutunian V, Mohs R, Davis KL, Davies P, Greengard P, et al. Correlation between elevated levels of amyloid beta-peptide in the brain and cognitive decline. JAMA 283(12): 1571-7. (2000).
[8]
Chung WS, Welsh CA, Barres BA, Stevens B. Do glia drive synaptic and cognitive impairment in disease? Nat Neurosci 18(11): 1539-45. (2015).
[9]
Peña F, Gutiérrez-Lerma A, Quiroz-Baez R, Arias C. The role of beta-amyloid protein in synaptic function: implications for Alzheimer’s disease therapy. Curr Neuropharmacol 4(2): 149-63. (2006).
[10]
Márquez M, Blancas-Mejía LM, Campos A, Rojas L, Castañeda-Hernández G, Quintanar L. A bifunctional non-natural tetrapeptide modulates amyloid-beta peptide aggregation in the presence of Cu(ii). Metallomics 6(12): 2189-92. (2014).
[11]
Bazzari FH, Abdallah DM, El-Abhar HS. Pharmacological interventions to attenuate Alzheimer’s disease progression: the story so far. Curr Alzheimer Res 16(3): 261-77. (2019).
[12]
Forloni G, Balducci C. Alzheimer’s disease, oligomers, and inflammation. J Alzheimers Dis 62(3): 1261-76. (2018).
[13]
Wang J, Tan L, Wang HF, Tan CC, Meng XF, Wang C, et al. Anti-inflammatory drugs and risk of Alzheimer’s disease: an updated systematic review and meta-analysis. J Alzheimers Dis 44(2): 385-96. (2015).
[14]
Xu W, Tan L, Wang HF, Jiang T, Tan MS, Tan L, et al. Meta-analysis of modifiable risk factors for Alzheimer’s disease. J Neurol Neurosurg Psychiatry 86(12): 1299-306. (2015).
[15]
Maheshwari P, Eslick GD. Bacterial infection and Alzheimer’s disease: a meta-analysis. J Alzheimers Dis 43(3): 957-66. (2015).
[16]
Selkoe DJ. Alzheimer’s disease is a synaptic failure. Science 298(5594): 789-91. (2002).
[17]
Jyoti A, Plano A, Riedel G, Platt B. EEG, activity, and sleep architecture in a transgenic AβPPswe/PSEN1A246E Alzheimer’s disease mouse. J Alzheimers Dis 22(3): 873-87. (2010).
[18]
Platt B, Drever B, Koss D, Stoppelkamp S, Jyoti A, Plano A, et al. Abnormal cognition, sleep, EEG and brain metabolism in a novel knock-in Alzheimer mouse, PLB1. PLoS One 6(11)e27068 (2011).
[19]
Wams EJ, Wilcock GK, Foster RG, Wulff K. Sleep-wake patterns and cognition of older adults with Amnestic Mild Cognitive Impairment (aMCI): a comparison with cognitively healthy adults and moderate Alzheimer’s disease patients. Curr Alzheimer Res 14(10): 1030-41. (2017).
[20]
Janelsins MC, Mastrangelo MA, Oddo S, LaFerla FM, Federoff HJ, Bowers WJ. Early correlation of microglial activation with enhanced tumor necrosis factor-alpha and monocyte chemoattractant protein-1 expression specifically within the entorhinal cortex of triple transgenic Alzheimer’s disease mice. J Neuroinflammation 2: 23. (2005).
[21]
Boelen E, Stassen FR, van der Ven AJ, Lemmens MA, Steinbusch HP, Bruggeman CA, et al. Detection of amyloid beta aggregates in the brain of BALB/c mice after Chlamydia pneumoniae infection. Acta Neuropathol 114(3): 255-61. (2007).
[22]
Stewart WF, Kawas C, Corrada M, Metter EJ. Risk of Alzheimer’s disease and duration of NSAID use. Neurology 48(3): 626-32. (1997).
[23]
Mackenzie IR, Munoz DG. Nonsteroidal anti-inflammatory drug use and Alzheimer-type pathology in aging. Neurology 50(4): 986-90. (1998).
[24]
Hu J, Akama KT, Krafft GA, Chromy BA, Van Eldik LJ. Amyloid-beta peptide activates cultured astrocytes: morphological alterations, cytokine induction and nitric oxide release. Brain Res 785(2): 195-206. (1998).
[25]
Szczepanik AM, Funes S, Petko W, Ringheim GE. IL-4, IL-10 and IL-13 modulate A beta(1--42)-induced cytokine and chemokine production in primary murine microglia and a human monocyte cell line. J Neuroimmunol 113(1): 49-62. (2001).
[26]
Ayasolla K, Khan M, Singh AK, Singh I. Inflammatory mediator and beta-amyloid (25-35)-induced ceramide generation and iNOS expression are inhibited by vitamin E. Free Radic Biol Med 37(3): 325-38. (2004).
[27]
Garção P, Oliveira CR, Agostinho P. Comparative study of microglia activation induced by amyloid-beta and prion peptides: role in neurodegeneration. J Neurosci Res 84(1): 182-93. (2006).
[28]
Parachikova A, Vasilevko V, Cribbs DH, LaFerla FM, Green KN. Reductions in amyloid-beta-derived neuroinflammation, with minocycline, restore cognition but do not significantly affect tau hyperphosphorylation. J Alzheimers Dis 21(2): 527-42. (2010).
[29]
Akama KT, Albanese C, Pestell RG, Van Eldik LJ. Amyloid beta-peptide stimulates nitric oxide production in astrocytes through an NFkappaB-dependent mechanism. Proc Natl Acad Sci USA 95(10): 5795-800. (1998).
[30]
Samuelsson M, Fisher L, Iverfeldt K. beta-Amyloid and interleukin-1beta induce persistent NF-kappaB activation in rat primary glial cells. Int J Mol Med 16(3): 449-53. (2005).
[31]
Ryu JK, Franciosi S, Sattayaprasert P, Kim SU, McLarnon JG. Minocycline inhibits neuronal death and glial activation induced by beta-amyloid peptide in rat hippocampus. Glia 48(1): 85-90. (2004).
[32]
Jekabsone A, Mander PK, Tickler A, Sharpe M, Brown GC. Fibrillar beta-amyloid peptide Abeta1-40 activates microglial proliferation via stimulating TNF-alpha release and H2O2 derived from NADPH oxidase: a cell culture study. J Neuroinflammation 3: 24. (2006).
[33]
Szaingurten-Solodkin I, Hadad N, Levy R. Regulatory role of cytosolic phospholipase A2alpha in NADPH oxidase activity and in inducible nitric oxide synthase induction by aggregated Abeta1-42 in microglia. Glia 57(16): 1727-40. (2009).
[34]
Floden AM, Li S, Combs CK. Beta-amyloid-stimulated microglia induce neuron death via synergistic stimulation of tumor necrosis factor alpha and NMDA receptors. J Neurosci 25(10): 2566-75. (2005).
[35]
Seabrook TJ, Jiang L, Maier M, Lemere CA. Minocycline affects microglia activation, Abeta deposition, and behavior in APP-tg mice. Glia 53(7): 776-82. (2006).
[36]
Milton RH, Abeti R, Averaimo S, DeBiasi S, Vitellaro L, Jiang L, et al. CLIC1 function is required for beta-amyloid-induced generation of reactive oxygen species by microglia. J Neurosci 28(45): 11488-99. (2008).
[37]
Parvathy S, Rajadas J, Ryan H, Vaziri S, Anderson L, Murphy GM Jr. Abeta peptide conformation determines uptake and interleukin-1alpha expression by primary microglial cells. Neurobiol Aging 30(11): 1792-804. (2009).
[38]
Yang SG, Wang WY, Ling TJ, Feng Y, Du XT, Zhang X, et al. α-Tocopherol quinone inhibits β-amyloid aggregation and cytotoxicity, disaggregates preformed fibrils and decreases the production of reactive oxygen species, NO and inflammatory cytokines. Neurochem Int 57(8): 914-22. (2010).
[39]
Reed-Geaghan EG, Savage JC, Hise AG, Landreth GE. CD14 and toll-like receptors 2 and 4 are required for fibrillar Aβ-stimulated microglial activation. J Neurosci 29(38): 11982-92. (2009).
[40]
Balleza-Tapia H, Peña F. Pharmacology of the intracellular pathways activated by amyloid beta protein. Mini Rev Med Chem 9(6): 724-40. (2009).
[41]
El-Shimy IA, Heikal OA, Hamdi N. Minocycline attenuates Aβ oligomers-induced pro-inflammatory phenotype in primary microglia while enhancing Aβ fibrils phagocytosis. Neurosci Lett 609: 36-41. (2015).
[42]
Peña-Ortega F. Pharmacological tools to activate microglia and their possible use to study neural network patho-physiology. Curr Neuropharmacol 15(4): 595-619. (2017).
[43]
Ferretti MT, Allard S, Partridge V, Ducatenzeiler A, Cuello AC. Minocycline corrects early, pre-plaque neuroinflammation and inhibits BACE-1 in a transgenic model of Alzheimer’s disease-like amyloid pathology. J Neuroinflammation 9: 62. (2012).
[44]
Garcez ML, Mina F, Bellettini-Santos T, Carneiro FG, Luz AP, Schiavo GL, et al. Minocycline reduces inflammatory parameters in the brain structures and serum and reverses memory impairment caused by the administration of amyloid β (1-42) in mice. Prog Neuropsychopharmacol Biol Psychiatry 77: 23-31. (2017).
[45]
Driver JE, Racca C, Cunningham MO, Towers SK, Davies CH, Whittington MA, et al. Impairment of hippocampal gamma-frequency oscillations in vitro in mice overexpressing human amyloid precursor protein (APP). Eur J Neurosci 26(5): 1280-8. (2007).
[46]
Balleza-Tapia H, Huanosta-Gutiérrez A, Márquez-Ramos A, Arias N, Peña F. Amyloid β oligomers decrease hippocampal spontaneous network activity in an age-dependent manner. Curr Alzheimer Res 7(5): 453-62. (2010).
[47]
Balleza-Tapia H, Crux S, Andrade-Talavera Y, Dolz-Gaiton P, Papadia D, Chen G, et al. TrpV1 receptor activation rescues neuronal function and network gamma oscillations from Aβ-induced impairment in mouse hippocampus in vitro. eLife 7pii e37703 (2019).
[48]
Villette V, Poindessous-Jazat F, Simon A, Léna C, Roullot E, Bellessort B, et al. Decreased rhythmic GABAergic septal activity and memory-associated theta oscillations after hippocampal amyloid-beta pathology in the rat. J Neurosci 30(33): 10991-1003. (2010).
[49]
Gutiérrez-Lerma AI, Ordaz B, Peña-Ortega F. Amyloid Beta peptides differentially affect hippocampal theta rhythms in vitro. Int J Pept 2013328140 (2013).
[50]
Kurudenkandy FR, Zilberter M, Biverstål H, Presto J, Honcharenko D, Strömberg R, et al. Amyloid-β-induced action potential desynchronization and degradation of hippocampal gamma oscillations is prevented by interference with peptide conformation change and aggregation. J Neurosci 34(34): 11416-25. (2014).
[51]
Peña-Ortega F. Amyloid beta-protein and neural network dysfunction. J Neurodegener Dis 2013657470 (2013).
[52]
Peña-Ortega F. Neural Network reconfigurations: changes of the respiratory network by hypoxia as an example. Adv Exp Med Biol 1015: 217-37. (2017).
[53]
Salgado-Puga K, Peña-Ortega F. Cellular and network mechanisms underlying memory impairment induced by amyloid β protein. Protein Pept Lett 22(4): 303-21. (2015).
[54]
Vanderwolf CH. Hippocampal electrical activity and voluntary movement in the rat. Electroencephalogr Clin Neurophysiol 26(4): 407-18. (1969).
[55]
Kramis R, Vanderwolf CH, Bland BH. Two types of hippocampal rhythmical slow activity in both the rabbit and the rat: relations to behavior and effects of atropine, diethyl ether, urethane, and pentobarbital. Exp Neurol 49(1 Pt 1): 58-85. (1975).
[56]
Bland BH. The physiology and pharmacology of hippocampal formation theta rhythms. Prog Neurobiol 26(1): 1-54. (1986).
[57]
Dragoi G, Buzsáki G. Temporal encoding of place sequences by hippocampal cell assemblies. Neuron 50(1): 145-57. (2006).
[58]
Kemp A, Manahan-Vaughan D. Hippocampal long-term depression: master or minion in declarative memory processes? Trends Neurosci 30(3): 111-8. (2007).
[59]
Hasselmo ME, Stern CE. Theta rhythm and the encoding and retrieval of space and time. Neuroimage 85(Pt 2): 656-66. (2014).
[60]
Berry SD, Thompson RF. Prediction of learning rate from the hippocampal electroencephalogram. Science 200(4347): 1298-300. (1978).
[61]
Winson J. Loss of hippocampal theta rhythm results in spatial memory deficit in the rat. Science 201(4351): 160-3. (1978).
[62]
Buzsáki G. Theta oscillations in the hippocampus. Neuron 33(3): 325-40. (2002).
[63]
Vertes RP. Hippocampal theta rhythm: a tag for short-term memory. Hippocampus 15(7): 923-35. (2005).
[64]
Buzsáki G, Draguhn A. Neuronal oscillations in cortical networks. Science 304(5679): 1926-9. (2004).
[65]
Sederberg PB, Kahana MJ, Howard MW, Donner EJ, Madsen JR. Theta and gamma oscillations during encoding predict subsequent recall. J Neurosci 23(34): 10809-14. (2003).
[66]
Sederberg PB, Schulze-Bonhage A, Madsen JR, Bromfield EB, Litt B, Brandt A, et al. Gamma oscillations distinguish true from false memories. Psychol Sci 18(11): 927-32. (2007).
[67]
Lisman J, Buzsáki G. A neural coding scheme formed by the combined function of gamma and theta oscillations. Schizophr Bull 34(5): 974-80. (2008).
[68]
Colgin LL, Moser EI. Gamma oscillations in the hippocampus. Physiology (Bethesda) 25(5): 319-29. (2010).
[69]
Tapia R, Medina-Ceja L, Peña F. On the relationship between extracellular glutamate, hyperexcitation and neurodegeneration, in vivo. Neurochem Int 34(1): 23-31. (1999).
[70]
Hajós M, Hoffmann WE, Kocsis B. Activation of cannabinoid-1 receptors disrupts sensory gating and neuronal oscillation: relevance to schizophrenia. Biol Psychiatry 63(11): 1075-83. (2008).
[71]
Kinney GG, Patino P, Mermet-Bouvier Y, Starrett JE Jr, Gribkoff VK. Cognition-enhancing drugs increase stimulated hippocampal theta rhythm amplitude in the urethane-anesthetized rat. J Pharmacol Exp Ther 291(1): 99-106. (1999).
[72]
McNaughton N, Ruan M, Woodnorth MA. Restoring theta-like rhythmicity in rats restores initial learning in the Morris water maze. Hippocampus 16(12): 1102-10. (2006).
[73]
McNaughton N, Kocsis B, Hajós M. Elicited hippocampal theta rhythm: a screen for anxiolytic and procognitive drugs through changes in hippocampal function? Behav Pharmacol 18(5-6): 329-46. (2007).
[74]
Siok CJ, Rogers JA, Kocsis B, Hajós M. Activation of alpha7 acetylcholine receptors augments stimulation-induced hippocampal theta oscillation. Eur J Neurosci 23(2): 570-4. (2006).
[75]
Robbe D, Montgomery SM, Thome A, Rueda-Orozco PE, McNaughton BL, Buzsaki G. Cannabinoids reveal importance of spike timing coordination in hippocampal function. Nat Neurosci 9(12): 1526-33. (2006).
[76]
Cornwell BR, Johnson LL, Holroyd T, Carver FW, Grillon C. Human hippocampal and parahippocampal theta during goal-directed spatial navigation predicts performance on a virtual Morris water maze. J Neurosci 28(23): 5983-90. (2008).
[77]
Kirov R, Weiss C, Siebner HR, Born J, Marshall L. Slow oscillation electrical brain stimulation during waking promotes EEG theta activity and memory encoding. Proc Natl Acad Sci USA 106(36): 15460-5. (2009).
[78]
Tort AB, Komorowski RW, Manns JR, Kopell NJ, Eichenbaum H. Theta-gamma coupling increases during the learning of item-context associations. Proc Natl Acad Sci USA 106(49): 20942-7. (2009).
[79]
Ramirez JM, Tryba AK, Peña F. Pacemaker neurons and neuronal networks: an integrative view. Curr Opin Neurobiol 14(6): 665-74. (2004).
[80]
Buzsáki G, Leung LW, Vanderwolf CH. Cellular bases of hippocampal EEG in the behaving rat. Brain Res 287(2): 139-71. (1983).
[81]
Freund TF, Buzsáki G. Interneurons of the hippocampus. Hippocampus 6(4): 347-470. (1996).
[82]
Tóth K, Freund TF, Miles R. Disinhibition of rat hippocampal pyramidal cells by GABAergic afferents from the septum. J Physiol 500(Pt 2): 463-74. (1997).
[83]
Lubenov EV, Siapas AG. Hippocampal theta oscillations are travelling waves. Nature 459(7246): 534-9. (2009).
[84]
Peña F, Ordaz B, Balleza-Tapia H, Bernal-Pedraza R, Márquez-Ramos A, Carmona-Aparicio L, et al. Beta-amyloid protein (25-35) disrupts hippocampal network activity: role of Fyn-kinase. Hippocampus 20(1): 78-96. (2010).
[85]
Goutagny R, Gu N, Cavanagh C, Jackson J, Chabot JG, Quirion R, et al. Alterations in hippocampal network oscillations and theta-gamma coupling arise before Aβ overproduction in a mouse model of Alzheimer’s disease. Eur J Neurosci 37(12): 1896-902. (2013).
[86]
Papazoglou A, Soos J, Lundt A, Wormuth C, Ginde VR, Müller R, et al. Gender-specific hippocampal dysrhythmia and aberrant hippocampal and cortical excitability in the APPswePS1dE9 model of Alzheimer’s disease. Neural Plast 20167167358 (2016).
[87]
Mondragón-Rodríguez S, Salas-Gallardo A, González-Pereyra P, Macías M, Ordaz B, Peña-Ortega F, et al. Phosphorylation of Tau protein correlates with changes in hippocampal theta oscillations and reduces hippocampal excitability in Alzheimer’s model. J Biol Chem 293(22): 8462-72. (2018).
[88]
Mondragón-Rodríguez S, Gu N, Manseau F, Williams S. Alzheimer’s transgenic model is characterized by very early brain network alterations and β-CTF fragment accumulation: reversal by β-secretase inhibition. Front Cell Neurosci 12: 121. (2018).
[89]
Stoiljkovic M, Kelley C, Horvath TL, Hajós M. Neurophysiological signals as predictive translational biomarkers for Alzheimer’s disease treatment: effects of donepezil on neuronal network oscillations in TgF344-AD rats. Alzheimers Res Ther 10(1): 105. (2018).
[90]
Cornejo-Montes-de-Oca JM, Hernández-Soto R, Isla AG, Morado-Urbina CE, Peña-Ortega F. Tolfenamic acid prevents amyloid β-induced olfactory bulb dysfunction in vivo. Curr Alzheimer Res 15(8): 731-42. (2018).
[91]
Mondragón-Rodríguez S, Gu N, Fasano C, Peña-Ortega F, Williams S. Functional connectivity between hippocampus and lateral septum is affected in very young Alzheimer’s transgenic mouse model. Neuroscience 401: 96-105. (2019).
[92]
Hernández-Soto R, Rojas-García KD, Peña-Ortega F. Sudden intrabulbar amyloid increase simultaneously disrupts olfactory bulb oscillations and odor detection. Neural Plast 20193424906 (2019).
[93]
Rutishauser U, Ross IB, Mamelak AN, Schuman EM. Human memory strength is predicted by theta-frequency phase-locking of single neurons. Nature 464(7290): 903-7. (2010).
[94]
Lisman J, Redish AD. Prediction, sequences and the hippocampus. Philos Trans R Soc Lond B Biol Sci 364(1521): 1193-201. (2009).
[95]
Lega BC, Jacobs J, Kahana M. Human hippocampal theta oscillations and the formation of episodic memories. Hippocampus 22(4): 748-61. (2012).
[96]
Berry SD, Rinaldi PC, Thompson RF, Verzeano M. Analysis of temporal relations among units and slow waves in rabbit hippocampus. Brain Res Bull 3(5): 509-18. (1978).
[97]
Klimesch W, Doppelmayr M, Russegger H, Pachinger T. Theta band power in the human scalp EEG and the encoding of new information. Neuroreport 7(7): 1235-40. (1996).
[98]
van der Hiele K, Vein AA, Kramer CG, Reijntjes RH, van Buchem MA, Westendorp RG, et al. Memory activation enhances EEG abnormality in mild cognitive impairment. Neurobiol Aging 28(1): 85-90. (2007).
[99]
Czigler B, Csikós D, Hidasi Z, Anna Gaál Z, Csibri E, Kiss E, et al. Quantitative EEG in early Alzheimer’s disease patients - power spectrum and complexity features. Int J Psychophysiol 68(1): 75-80. (2008).
[100]
Moretti DV, Pievani M, Geroldi C, Binetti G, Zanetti O, Rossini PM, et al. EEG markers discriminate among different subgroup of patients with mild cognitive impairment. Am J Alzheimers Dis Other Demen 25(1): 58-73. (2010).
[101]
Caravaglios G, Castro G, Costanzo E, Di Maria G, Mancuso D, Muscoso EG. Power responses in mild Alzheimer’s disease during an auditory oddball paradigm: lack of theta enhancement during stimulus processing. J Neural Transm (Vienna) 117(10): 1195-208. (2010).
[102]
Başar E, Güntekin B, Tülay E, Yener GG. Evoked and event related coherence of Alzheimer patients manifest differentiation of sensory-cognitive networks. Brain Res 1357: 79-90. (2010).
[103]
Colgin LL, Denninger T, Fyhn M, Hafting T, Bonnevie T, Jensen O, et al. Frequency of gamma oscillations routes flow of information in the hippocampus. Nature 462(7271): 353-7. (2009).
[104]
Li W, Li S, Shen L, Wang J, Wu X, Li J, et al. Impairment of dendrodendritic inhibition in the olfactory bulb of APP/PS1 mice. Front Aging Neurosci 11: 2. (2019).
[105]
Gray CM, König P, Engel AK, Singer W. Oscillatory responses in cat visual cortex exhibit inter-columnar synchronization which reflects global stimulus properties. Nature 338(6213): 334-7. (1989).
[106]
Buzsáki G, Chrobak JJ. Temporal structure in spatially organized neuronal ensembles: a role for interneuronal networks. Curr Opin Neurobiol 5(4): 504-10. (1995).
[107]
Lisman JE, Idiart MA. Storage of 7 +/- 2 short-term memories in oscillatory subcycles. Science 267(5203): 1512-5. (1995).
[108]
Lisman J. The theta/gamma discrete phase code occuring during the hippocampal phase precession may be a more general brain coding scheme. Hippocampus 15(7): 913-22. (2005).
[109]
Herrmann CS, Demiralp T. Human EEG gamma oscillations in neuropsychiatric disorders. Clin Neurophysiol 116(12): 2719-33. (2005).
[110]
Koenig T, Prichep L, Dierks T, Hubl D, Wahlund LO, John ER, et al. Decreased EEG synchronization in Alzheimer’s disease and mild cognitive impairment. Neurobiol Aging 26(2): 165-71. (2005).
[111]
de Haan W, van der Flier WM, Koene T, Smits LL, Scheltens P, Stam CJ. Disrupted modular brain dynamics reflect cognitive dysfunction in Alzheimer’s disease. Neuroimage 59(4): 3085-93. (2012).
[112]
Gouw AA, Alsema AM, Tijms BM, Borta A, Scheltens P, Stam CJ, et al. EEG spectral analysis as a putative early prognostic biomarker in nondemented, amyloid positive subjects. Neurobiol Aging 57: 133-42. (2017).
[113]
Canolty RT, Edwards E, Dalal SS, Soltani M, Nagarajan SS, Kirsch HE, et al. High gamma power is phase-locked to theta oscillations in human neocortex. Science 313(5793): 1626-8. (2006).
[114]
Canolty RT, Ganguly K, Kennerley SW, Cadieu CF, Koepsell K, Wallis JD, et al. Oscillatory phase coupling coordinates anatomically dispersed functional cell assemblies. Proc Natl Acad Sci USA 107(40): 17356-61. (2010).
[115]
Belluscio MA, Mizuseki K, Schmidt R, Kempter R, Buzsáki G. Cross-frequency phase-phase coupling between θ and γ oscillations in the hippocampus. J Neurosci 32(2): 423-35. (2012).
[116]
Scheffer-Teixeira R, Belchior H, Leão RN, Ribeiro S, Tort AB. On high-frequency field oscillations (>100 Hz) and the spectral leakage of spiking activity. J Neurosci 33(4): 1535-9. (2013).
[117]
Matsumoto JY, Stead M, Kucewicz MT, Matsumoto AJ, Peters PA, Brinkmann BH, et al. Network oscillations modulate interictal epileptiform spike rate during human memory. Brain 136(Pt 8): 2444-56. (2013).
[118]
Yamamoto J, Suh J, Takeuchi D, Tonegawa S. Successful execution of working memory linked to synchronized high-frequency gamma oscillations. Cell 157(4): 845-57. (2014).
[119]
Vida I, Bartos M, Jonas P. Shunting inhibition improves robustness of gamma oscillations in hippocampal interneuron networks by homogenizing firing rates. Neuron 49(1): 107-17. (2006).
[120]
Fuchs EC, Zivkovic AR, Cunningham MO, Middleton S, Lebeau FE, Bannerman DM, et al. Recruitment of parvalbumin-positive interneurons determines hippocampal function and associated behavior. Neuron 53(4): 591-604. (2007).
[121]
Händel B, Haarmeier T. Cross-frequency coupling of brain oscillations indicates the success in visual motion discrimination. Neuroimage 45(3): 1040-6. (2009).
[122]
Axmacher N, Henseler MM, Jensen O, Weinreich I, Elger CE, Fell J. Cross-frequency coupling supports multi-item working memory in the human hippocampus. Proc Natl Acad Sci USA 107(7): 3228-33. (2010).
[123]
Wulff P, Ponomarenko AA, Bartos M, Korotkova TM, Fuchs EC, Bähner F, et al. Hippocampal theta rhythm and its coupling with gamma oscillations require fast inhibition onto parvalbumin-positive interneurons. Proc Natl Acad Sci USA 106(9): 3561-6. (2009).
[124]
Igarashi KM, Lu L, Colgin LL, Moser MB, Moser EI. Coordination of entorhinal-hippocampal ensemble activity during associative learning. Nature 510(7503): 143-7. (2014).
[125]
Girardeau G, Zugaro M. Hippocampal ripples and memory consolidation. Curr Opin Neurobiol (3): 452-9. (2011).
[126]
Buzsáki G. Hippocampal sharp wave-ripple: a cognitive biomarker for episodic memory and planning. Hippocampus 25(10): 1073-188. (2015).
[127]
Ego-Stengel V, Wilson MA. Disruption of ripple-associated hippocampal activity during rest impairs spatial learning in the rat. Hippocampus 20(1): 1-10. (2010).
[128]
Axmacher N, Elger CE, Fell J. Ripples in the medial temporal lobe are relevant for human memory consolidation. Brain 131(Pt 7): 1806-17. (2008).
[129]
Coben LA, Danziger WL, Berg L. Frequency analysis of the resting awake EEG in mild senile dementia of Alzheimer type. Electroencephalogr Clin Neurophysiol 55(4): 372-80. (1983).
[130]
Schreiter-Gasser U, Gasser T, Ziegler P. Quantitative EEG analysis in early onset Alzheimer’s disease: correlations with severity, clinical characteristics, visual EEG and CCT. Electroencephalogr Clin Neurophysiol 90(4): 267-72. (1994).
[131]
Kowalski JW, Gawel M, Pfeffer A, Barcikowska M. The diagnostic value of EEG in Alzheimer disease: correlation with the severity of mental impairment. J Clin Neurophysiol 18(6): 570-5. (2001).
[132]
Vecchio F, Babiloni C, Lizio R, Fallani Fde V, Blinowska K, Verrienti G, et al. Resting state cortical EEG rhythms in Alzheimer’s disease: toward EEG markers for clinical applications: a review. Suppl Clin Neurophysiol 62: 223-36. (2013).
[133]
Garn H, Waser M, Deistler M, Schmidt R, Dal-Bianco P, Ransmayr G, et al. Quantitative EEG in Alzheimer’s disease: cognitive state, resting state and association with disease severity. Int J Psychophysiol 93(3): 390-7. (2014).
[134]
Bero AW, Yan P, Roh JH, Cirrito JR, Stewart FR, Raichle ME, et al. Neuronal activity regulates the regional vulnerability to amyloid-β deposition. Nat Neurosci 14(6): 750-6. (2011).
[135]
Dolev I, Fogel H, Milshtein H, Berdichevsky Y, Lipstein N, Brose N, et al. Spike bursts increase amyloid-β 40/42 ratio by inducing a presenilin-1 conformational change. Nat Neurosci 16(5): 587-95. (2013).
[136]
Tampellini D, Capetillo-Zarate E, Dumont M, Huang Z, Yu F, Lin MT, et al. Effects of synaptic modulation on beta-amyloid, synaptophysin, and memory performance in Alzheimer’s disease transgenic mice. J Neurosci 30(43): 14299-304. (2010).
[137]
Iaccarino HF, Singer AC, Martorell AJ, Rudenko A, Gao F, Gillingham TZ, et al. Gamma frequency entrainment attenuates amyloid load and modifies microglia. Nature 540(7632): 230-5. (2016).
[138]
Isla AG, Vázquez-Cuevas FG, Peña-Ortega F. Exercise prevents amyloid-β-induced hippocampal network disruption by inhibiting GSK3β activation. J Alzheimers Dis 52(1): 333-43. (2016).
[139]
Ziegler-Waldkirch S, d’Errico P, Sauer JF, Erny D, Savanthrapadian S, Loreth D, et al. Seed-induced Aβ deposition is modulated by microglia under environmental enrichment in a mouse model of Alzheimer’s disease. EMBO J 37(2): 167-82. (2018).
[140]
Martorell AJ, Paulson AL, Suk HJ, Abdurrob F, Drummond GT, Guan W, et al. Multi-sensory gamma stimulation ameliorates Alzheimer's-associated pathology and improves cognition Cell pii: S0092-8674(19): 30163-1 (2019).
[141]
Lazarov O, Robinson J, Tang YP, Hairston IS, Korade-Mirnics Z, Lee VM, et al. Environmental enrichment reduces Abeta levels and amyloid deposition in transgenic mice. Cell 120(5): 701-13. (2005).
[142]
Mainardi M, Di Garbo A, Caleo M, Berardi N, Sale A, Maffei L. Environmental enrichment strengthens corticocortical interactions and reduces amyloid-β oligomers in aged mice. Front Aging Neurosci 6: 1. (2014).
[143]
Koo JH, Kang EB, Oh YS, Yang DS, Cho JY. Treadmill exercise decreases amyloid-β burden possibly via activation of SIRT-1 signaling in a mouse model of Alzheimer’s disease. Exp Neurol 288: 142-52. (2017).
[144]
Zhang J, Guo Y, Wang Y, Song L, Zhang R, Du Y. Long-term treadmill exercise attenuates Aβ burdens and astrocyte activation in APP/PS1 mouse model of Alzheimer’s disease. Neurosci Lett 666: 70-7. (2018).
[145]
de Medeiros CB, Fleming AS, Johnston CC, Walker CD. Artificial rearing of rat pups reveals the beneficial effects of mother care on neonatal inflammation and adult sensitivity to pain. Pediatr Res 66(3): 272-7. (2009).
[146]
Xu H, Gelyana E, Rajsombath M, Yang T, Li S, Selkoe D. Environmental enrichment potently prevents microglia-mediated neuroinflammation by human amyloid β-protein oligomers. J Neurosci 36(35): 9041-56. (2016).
[147]
Grier BD, Belluscio L, Cheetham CE. Olfactory sensory activity modulates microglial-neuronal interactions during dopaminergic cell loss in the olfactory bulb. Front Cell Neurosci 10: 178. (2016).
[148]
Denizet M, Cotter L, Lledo PM, Lazarini F. Sensory deprivation increases phagocytosis of adult-born neurons by activated microglia in the olfactory bulb. Brain Behav Immun 60: 38-43. (2017).
[149]
Delbeuck X, Van der Linden M, Collette F. Alzheimer’s disease as a disconnection syndrome? Neuropsychol Rev 13(2): 79-92. (2003).
[150]
Agosta F, Pievani M, Geroldi C, Copetti M, Frisoni GB, Filippi M. Resting state fMRI in Alzheimer’s disease: beyond the default mode network. Neurobiol Aging 33(8): 1564-78. (2012).
[151]
Binnewijzend MA, Schoonheim MM, Sanz-Arigita E, Wink AM, van der Flier WM, Tolboom N, et al. Resting-state fMRI changes in Alzheimer’s disease and mild cognitive impairment. Neurobiol Aging 33(9): 2018-28. (2012).
[152]
Greicius MD, Srivastava G, Reiss AL, Menon V. Default-mode network activity distinguishes Alzheimer’s disease from healthy aging: evidence from functional MRI. Proc Natl Acad Sci USA 101(13): 4637-42. (2004).
[153]
Wang K, Liang M, Wang L, et al. Altered functional connectivity in early Alzheimer’s disease: a resting-state fMRI study. Hum Brain Mapp 28(10): 967-78. (2007).
[154]
Sperling RA, Dickerson BC, Pihlajamaki M, Vannini P, LaViolette PS, Vitolo OV, et al. Functional alterations in memory networks in early Alzheimer’s disease. Neuromolecular Med 12(1): 27-43. (2010).
[155]
Shah D, Jonckers E, Praet J, et al. Resting state FMRI reveals diminished functional connectivity in a mouse model of amyloidosis. PLoS One 8(12)e84241 (2013).
[156]
Vasavada MM, Martinez B, Wang J, Eslinger PJ, Gill DJ, Sun X, et al. Central olfactory dysfunction in Alzheimer’s disease and mild cognitive impairment: a functional MRI study. J Alzheimers Dis 59(1): 359-68. (2017).
[157]
Wesson DW, Borkowski AH, Landreth GE, Nixon RA, Levy E, Wilson DA. Sensory network dysfunction, behavioral impairments, and their reversibility in an Alzheimer’s β-amyloidosis mouse model. J Neurosci 31(44): 15962-71. (2011).
[158]
Liu Q, Li A, Gong L, Zhang L, Wu N, Xu F. Decreased coherence between the two olfactory bulbs in Alzheimer’s disease model mice. Neurosci Lett 545: 81-5. (2013).
[159]
Babiloni C, Vecchio F, Del Percio C, Montagnese S, Schiff S, Lizio R, et al. Resting state cortical electroencephalographic rhythms in covert hepatic encephalopathy and Alzheimer’s disease. J Alzheimers Dis 34(3): 707-25. (2013).
[160]
Kramberger MG, Giske K, Cavallin L, et al. Subclinical white matter lesions and medial temporal lobe atrophy are associated with EEG slowing in a memory clinic cohort. Clin Neurophysiol 128(9): 1575-82. (2017).
[161]
Faigle R, Sutter R, Kaplan PW. Electroencephalography of encephalopathy in patients with endocrine and metabolic disorders. J Clin Neurophysiol 30(5): 505-16. (2013).
[162]
Young GB, Bolton CF, Austin TW, Archibald YM, Gonder J, Wells GA. The encephalopathy associated with septic illness. Clin Invest Med 13(6): 297-304. (1990).
[163]
Jelic V, Shigeta M, Julin P, Almkvist O, Winblad B, Wahlund LO. Quantitative electroencephalography power and coherence in Alzheimer’s disease and mild cognitive impairment. Dementia 7(6): 314-23. (1996).
[164]
Wada Y, Nanbu Y, Jiang ZY, Koshino Y, Yamaguchi N, Hashimoto T. Electroencephalographic abnormalities in patients with presenile dementia of the Alzheimer type: quantitative analysis at rest and during photic stimulation. Biol Psychiatry 41(2): 217-25. (1997).
[165]
Huang C, Wahlund L, Dierks T, Julin P, Winblad B, Jelic V. Discrimination of Alzheimer’s disease and mild cognitive impairment by equivalent EEG sources: a cross-sectional and longitudinal study. Clin Neurophysiol 111(11): 1961-7. (2000).
[166]
Amatniek JC, Hauser WA, DelCastillo-Castaneda C, Jacobs DM, Marder K, Bell K, et al. Incidence and predictors of seizures in patients with Alzheimer’s disease. Epilepsia 47(5): 867-72. (2006).
[167]
Mendez MF, Catanzaro P, Doss RC. ARguello R, Frey WH 2nd. Seizures in Alzheimer’s disease: clinicopathologic study. J Geriatr Psychiatry Neurol 7(4): 230-3. (1994).
[168]
Scarmeas N, Honig LS, Choi H, Cantero J, Brandt J, Blacker D, et al. Seizures in Alzheimer disease: who, when, and how common? Arch Neurol 66(8): 992-7. (2009).
[169]
Vossel KA, Ranasinghe KG, Beagle AJ, Mizuiri D, Honma SM, Dowling AF, et al. Incidence and impact of subclinical epileptiform activity in Alzheimer’s disease. Ann Neurol 80(6): 858-70. (2016).
[170]
Vossel KA, Beagle AJ, Rabinovici GD, Shu H, Lee SE, Naasan G, et al. Seizures and epileptiform activity in the early stages of Alzheimer disease. JAMA Neurol 70(9): 1158-66. (2013).
[171]
Vossel KA, Tartaglia MC, Nygaard HB, Zeman AZ, Miller BL. Epileptic activity in Alzheimer’s disease: causes and clinical relevance. Lancet Neurol 16(4): 311-22. (2017).
[172]
Wang J, Ikonen S, Gurevicius K, van Groen T, Tanila H. Alteration of cortical EEG in mice carrying mutated human APP transgene. Brain Res 943(2): 181-90. (2002).
[173]
Palop JJ, Chin J, Roberson ED, Wang J, Thwin MT, Bien-Ly N, et al. Aberrant excitatory neuronal activity and compensatory remodeling of inhibitory hippocampal circuits in mouse models of Alzheimer’s disease. Neuron 55(5): 697-711. (2007).
[174]
Akay M, Wang K, Akay YM, Dragomir A, Wu J. Nonlinear dynamical analysis of carbachol induced hippocampal oscillations in mice. Acta Pharmacol Sin 30(6): 859-67. (2009).
[175]
Scott L, Feng J, Kiss T, Needle E, Atchison K, Kawabe TT, et al. Age-dependent disruption in hippocampal θ oscillation in amyloid-β overproducing transgenic mice. Neurobiol Aging 33(7): 1481.e13-. (2012).
[176]
Rubio SE, Vega-Flores G, Martínez A, Bosch C, Pérez-Mediavilla A, del Río J, et al. Accelerated aging of the GABAergic septohippocampal pathway and decreased hippocampal rhythms in a mouse model of Alzheimer’s disease. FASEB J 26(11): 4458-67. (2012).
[177]
Verret L, Mann EO, Hang GB, Barth AM, Cobos I, Ho K, et al. Inhibitory interneuron deficit links altered network activity and cognitive dysfunction in Alzheimer model. Cell 149(3): 708-21. (2012).
[178]
Cramer PE, Cirrito JR, Wesson DW, Lee CY, Karlo JC, Zinn AE, et al. ApoE-directed therapeutics rapidly clear β-amyloid and reverse deficits in AD mouse models. Science 335(6075): 1503-6. (2012).
[179]
Schneider F, Baldauf K, Wetzel W, Reymann KG. Behavioral and EEG changes in male 5xFAD mice. Physiol Behav 135: 25-33. (2014).
[180]
Lison H, Happel MF, Schneider F, Baldauf K, Kerbstat S, Seelbinder B, et al. Disrupted cross-laminar cortical processing in β amyloid pathology precedes cell death. Neurobiol Dis 63: 62-73. (2014).
[181]
Busche MA, Kekuš M, Adelsberger H, Noda T, Förstl H, Nelken I, et al. Rescue of long-range circuit dysfunction in Alzheimer’s disease models. Nat Neurosci 18(11): 1623-30. (2015).
[182]
Kastanenka KV, Hou SS, Shakerdge N, Logan R, Feng D, Wegmann S, et al. Optogenetic restoration of disrupted slow oscillations halts amyloid deposition and restores calcium homeostasis in an animal model of Alzheimer’s disease. PLoS One 12(1)e0170275 (2017).
[183]
Zhurakovskaya E, Ishchenko I, Gureviciene I, Aliev R, Gröhn O, Tanila H. Impaired hippocampal-cortical coupling but preserved local synchrony during sleep in APP/PS1 mice modeling Alzheimer’s disease. Sci Rep 9(1): 5380. (2019).
[184]
Gurevicius K, Lipponen A, Tanila H. Increased cortical and thalamic excitability in freely moving APPswe/PS1dE9 mice modeling epileptic activity associated with Alzheimer’s disease. Cereb Cortex 23(5): 1148-58. (2013).
[185]
Maatuf Y, Stern EA, Slovin H. Abnormal population responses in the somatosensory cortex of Alzheimer’s disease model mice. Sci Rep 6: 24560. (2016).
[186]
Stoiljkovic M, Kelley C, Stutz B, Horvath TL, Hajós M. Altered cortical and hippocampal excitability in TGF344-AD rats modeling Alzheimer’s disease pathology. Cereb Cortex 29(6): 2716-27. (2018).
[187]
Cayzac S, Mons N, Ginguay A, Allinquant B, Jeantet Y, Cho YH. Altered hippocampal information coding and network synchrony in APP-PS1 mice. Neurobiol Aging 36(12): 3200-13. (2015).
[188]
Miki Stein A, Munive V, Fernandez AM, Nuñez A, Torres Aleman I. Acute exercise does not modify brain activity and memory performance in APP/PS1 mice. PLoS One 12(5)e0178247 (2017).
[189]
Siwek ME, Müller R, Henseler C, Trog A, Lundt A, Wormuth C, et al. Altered theta oscillations and aberrant cortical excitatory activity in the 5XFAD model of Alzheimer’s disease. Neural Plast 2015781731 (2015).
[190]
Stoiljkovic M, Kelley C, Hajós GP, Nagy D, Koenig G, Leventhal L, et al. Hippocampal network dynamics in response to α7 nACh receptors activation in amyloid-β overproducing transgenic mice. Neurobiol Aging 45: 161-8. (2016).
[191]
Bazzigaluppi P, Beckett TL, Koletar MM, Lai AY, Joo IL, Brown ME, et al. Early-stage attenuation of phase-amplitude coupling in the hippocampus and medial prefrontal cortex in a transgenic rat model of Alzheimer’s disease. J Neurochem 144(5): 669-79. (2018).
[192]
Cacucci F, Yi M, Wills TJ, Chapman P, O’Keefe J. Place cell firing correlates with memory deficits and amyloid plaque burden in Tg2576 Alzheimer mouse model. Proc Natl Acad Sci USA 105(22): 7863-8. (2008).
[193]
Cheng J, Ji D. Rigid firing sequences undermine spatial memory codes in a neurodegenerative mouse model. eLife 2e00647 (2013).
[194]
Zhao R, Fowler SW, Chiang AC, Ji D, Jankowsky JL. Impairments in experience-dependent scaling and stability of hippocampal place fields limit spatial learning in a mouse model of Alzheimer’s disease. Hippocampus 24(8): 963-78. (2014).
[195]
Mably AJ, Gereke BJ, Jones DT, Colgin LL. Impairments in spatial representations and rhythmic coordination of place cells in the 3xTg mouse model of Alzheimer’s disease. Hippocampus 27(4): 378-92. (2017).
[196]
Zhen J, Qian Y, Weng X, Su W, Zhang J, Cai L, et al. Gamma rhythm low field magnetic stimulation alleviates neuropathologic changes and rescues memory and cognitive impairments in a mouse model of Alzheimer’s disease. Alzheimers Dement (N Y) 3(4): 487-97. (2017).
[197]
Corbett BF, Leiser SC, Ling HP, Nagy R, Breysse N, Zhang X, et al. Sodium channel cleavage is associated with aberrant neuronal activity and cognitive deficits in a mouse model of Alzheimer’s disease. J Neurosci 33(16): 7020-6. (2013).
[198]
Ittner AA, Gladbach A, Bertz J, Suh LS, Ittner LM. p38 MAP kinase-mediated NMDA receptor-dependent suppression of hippocampal hypersynchronicity in a mouse model of Alzheimer’s disease. Acta Neuropathol Commun 2: 149. (2014).
[199]
Fontana R, Agostini M, Murana E, Mahmud M, Scremin E, Rubega M, et al. Early hippocampal hyperexcitability in PS2APP mice: role of mutant PS2 and APP. Neurobiol Aging 50: 64-76. (2017).
[200]
Jin N, Lipponen A, Koivisto H, Gurevicius K, Tanila H. Increased cortical beta power and spike-wave discharges in middle-aged APP/PS1 mice. Neurobiol Aging 71: 127-41. (2018).
[201]
Witton J, Staniaszek LE, Bartsch U, Randall AD, Jones MW, Brown JT. Disrupted hippocampal sharp-wave ripple-associated spike dynamics in a transgenic mouse model of dementia. J Physiol 594(16): 4615-30. (2016).
[202]
Nakazono T, Lam TN, Patel AY, et al. Impaired In Vivo gamma oscillations in the medial entorhinal cortex of knock-in Alzheimer model. Front Syst Neurosci 11: 48. (2017).
[203]
Bomben V, Holth J, Reed J, Cramer P, Landreth G, Noebels J. Bexarotene reduces network excitability in models of Alzheimer’s disease and epilepsy. Neurobiol Aging 35(9): 2091-5. (2014).
[204]
Born HA, Kim JY, Savjani RR, Das P, Dabaghian YA, Guo Q, et al. Genetic suppression of transgenic APP rescues Hypersynchronous network activity in a mouse model of Alzeimer’s disease. J Neurosci 34(11): 3826-40. (2014).
[205]
Minkeviciene R, Rheims S, Dobszay MB, Zilberter M, Hartikainen J, Fülöp L, et al. Amyloid beta-induced neuronal hyperexcitability triggers progressive epilepsy. J Neurosci 29(11): 3453-62. (2009).
[206]
Westmark CJ, Westmark PR, Beard AM, Hildebrandt SM, Malter JS. Seizure susceptibility and mortality in mice that over-express amyloid precursor protein. Int J Clin Exp Pathol 1(2): 157-68. (2008).
[207]
Westmark CJ, Westmark PR, Malter JS. Alzheimer’s disease and Down syndrome rodent models exhibit audiogenic seizures. J Alzheimers Dis 20(4): 1009-13. (2010).
[208]
Ziyatdinova S, Gurevicius K, Kutchiashvili N, Bolkvadze T, Nissinen J, Tanila H, et al. Spontaneous epileptiform discharges in a mouse model of Alzheimer’s disease are suppressed by antiepileptic drugs that block sodium channels. Epilepsy Res 94(1-2): 75-85. (2011).
[209]
Sanchez PE, Zhu L, Verret L, Vossel KA, Orr AG, Cirrito JR, et al. Levetiracetam suppresses neuronal network dysfunction and reverses synaptic and cognitive deficits in an Alzheimer’s disease model. Proc Natl Acad Sci USA 109(42): E2895-903. (2012).
[210]
Roberson ED, Scearce-Levie K, Palop JJ, Yan F, Cheng IH, Wu T, et al. Reducing endogenous tau ameliorates amyloid beta-induced deficits in an Alzheimer’s disease mouse model. Science 316(5825): 750-4. (2007).
[211]
Del Vecchio RA, Gold LH, Novick SJ, Wong G, Hyde LA. Increased seizure threshold and severity in young transgenic CRND8 mice. Neurosci Lett 367(2): 164-7. (2004).
[212]
Nygaard HB, Kaufman AC, Sekine-Konno T, Huh LL, Going H, Feldman SJ, et al. Brivaracetam, but not ethosuximide, reverses memory impairments in an Alzheimer’s disease mouse model. Alzheimers Res Ther 7(1): 25. (2015).
[213]
Baglietto-Vargas D, Moreno-Gonzalez I, Sanchez-Varo R, Jimenez S, Trujillo-Estrada L, Sanchez-Mejias E, et al. Calretinin interneurons are early targets of extracellular amyloid-beta pathology in PS1/AbetaPP Alzheimer mice hippocampus. J Alzheimers Dis 21(1): 119-32. (2010).
[214]
Mahar I, Albuquerque MS, Mondragon-Rodriguez S, Cavanagh C, Davoli MA, Chabot JG, et al. Phenotypic alterations in hippocampal NPY- and PV-expressing interneurons in a presymptomatic transgenic mouse model of Alzheimer’s disease. Front Aging Neurosci 8: 327. (2017).
[215]
Busche MA, Eichhoff G, Adelsberger H, Abramowski D, Wiederhold KH, Haass C, et al. Clusters of hyperactive neurons near amyloid plaques in a mouse model of Alzheimer’s disease. Science 321(5896): 1686-9. (2008).
[216]
Colby-Milley J, Cavanagh C, Jego S, Breitner JC, Quirion R, Adamantidis A. Sleep-wake cycle dysfunction in the TgCRND8 mouse model of alzheimer’s disease: from early to advanced pathological stages. PLoS One 10(6)e0130177 (2015).
[217]
Hamm V, Héraud C, Bott JB, Herbeaux K, Strittmatter C, Mathis C, et al. Differential contribution of APP metabolites to early cognitive deficits in a TgCRND8 mouse model of Alzheimer’s disease. Sci Adv 3(2)e1601068 (2017).
[218]
Sánchez-Alavez M, Chan SL, Mattson MP, Criado JR. Electrophysiological and cerebrovascular effects of the alpha-secretase-derived form of amyloid precursor protein in young and middle-aged rats. Brain Res 1131(1): 112-7. (2007).
[219]
Jiang Y, Mullaney KA, Peterhoff CM, Che S, Schmidt SD, Boyer-Boiteau A, et al. Alzheimer’s-related endosome dysfunction in Down syndrome is Abeta-independent but requires APP and is reversed by BACE-1 inhibition. Proc Natl Acad Sci USA 107(4): 1630-5. (2010).
[220]
Vogt DL, Thomas D, Galvan V, Bredesen DE, Lamb BT, Pimplikar SW. Abnormal neuronal networks and seizure susceptibility in mice overexpressing the APP intracellular domain. Neurobiol Aging 32(9): 1725-9. (2011).
[221]
Baker JE, Lim YY, Pietrzak RH, Hassenstab J, Snyder PJ, Masters CL, et al. Cognitive impairment and decline in cognitively normal older adults with high amyloid-β: a meta-analysis. Alzheimers Dement (Amst) 6: 108-21. (2016).
[222]
Yankner BA, Duffy LK, Kirschner DA. Neurotrophic and neurotoxic effects of amyloid beta protein: reversal by tachykinin neuropeptides. Science 250(4978): 279-82. (1990).
[223]
Pike CJ, Burdick D, Walencewicz AJ, Glabe CG, Cotman CW. Neurodegeneration induced by beta-amyloid peptides in vitro: the role of peptide assembly state. J Neurosci 13(4): 1676-87. (1993).
[224]
Walsh DM, Klyubin I, Fadeeva JV, Cullen WK, Anwyl R, Wolfe MS, et al. Naturally secreted oligomers of amyloid beta protein potently inhibit hippocampal long-term potentiation in vivo. Nature 416(6880): 535-9. (2002).
[225]
Adaya-Villanueva A, Ordaz B, Balleza-Tapia H, Márquez-Ramos A, Peña-Ortega F. Beta-like hippocampal network activity is differentially affected by amyloid beta peptides. Peptides 31(9): 1761-6. (2010).
[226]
Sun MK, Alkon DL. Impairment of hippocampal CA1 heterosynaptic transformation and spatial memory by beta-amyloid(25-35). J Neurophysiol 87(5): 2441-9. (2002).
[227]
Peña-Ortega F, Solis-Cisneros A, Ordaz B, Balleza-Tapia H, Javier Lopez-Guerrero J. Amyloid beta 1-42 inhibits entorhinal cortex activity in the beta-gamma range: role of GSK-3. Curr Alzheimer Res 9(7): 857-63. (2012).
[228]
Salgado-Puga K, Rodríguez-Colorado J, Prado-Alcalá RA, Peña-Ortega F. Subclinical doses of ATP-sensitive potassium channel modulators prevent alterations in memory and synaptic plasticity induced by amyloid-β. J Alzheimers Dis 57(1): 205-26. (2017).
[229]
Peña-Ortega F, Bernal-Pedraza R. Amyloid beta peptide slows down sensory-induced hippocampal oscillations. Int J Pept 2012236289 (2012).
[230]
Alvarado-Martínez R, Salgado-Puga K, Peña-Ortega F. Amyloid beta inhibits olfactory bulb activity and the ability to smell. PLoS One 8(9)e75745 (2013).
[231]
Flores-Martínez E, Peña-Ortega F. Amyloid β peptide-induced changes in prefrontal cortex activity and its response to hippocampal input. Int J Pept 20177386809 (2017).
[232]
Vorobyov V, Kaptsov V, Gordon R, Makarova E, Podolski I, Sengpiel F. Neuroprotective effects of hydrated fullerene C60: cortical and hippocampal EEG interplay in an amyloid-infused rat model of Alzheimer’s disease. J Alzheimers Dis 45(1): 217-33. (2015).
[233]
Kantar Gok D, Hidisoglu E, Ocak GA, Er H, Acun AD, Yargıcoglu P. Protective role of rosmarinic acid on amyloid beta 42-induced echoic memory decline: implication of oxidative stress and cholinergic impairment. Neurochem Int 118: 1-13. (2018).
[234]
Bai W, Xia M, Liu T, Tian X. Aβ1-42-induced dysfunction in synchronized gamma oscillation during working memory. Behav Brain Res 307: 112-9. (2016).
[235]
Hidisoglu E, Kantar-Gok D, Er H, Acun AD, Yargicoglu P. Alterations in spontaneous delta and gamma activity might provide clues to detect changes induced by amyloid-β administration. Eur J Neurosci 47(8): 1013-23. (2018).
[236]
Maleysson V, Page G, Janet T, Klein RL, Haida O, Maurin A, et al. Relevance of electroencephalogram assessment in amyloid and tau pathology in rat. Behav Brain Res 359: 127-34. (2019).
[237]
Salgado-Puga K, Prado-Alcalá RA, Peña-Ortega F. Amyloid β enhances typical rodent behavior while it impairs contextual memory consolidation. Behav Neurol 2015526912 (2015).
[238]
Nicole O, Hadzibegovic S, Gajda J, Bontempi B, Bem T, Meyrand P. Soluble amyloid beta oligomers block the learning-induced increase in hippocampal sharp wave-ripple rate and impair spatial memory formation. Sci Rep 6: 22728. (2016).
[239]
Kalweit AN, Yang H, Colitti-Klausnitzer J, Fülöp L, Bozsó Z, Penke B, et al. Acute intracerebral treatment with amyloid-beta (1-42) alters the profile of neuronal oscillations that accompany LTP induction and results in impaired LTP in freely behaving rats. Front Behav Neurosci 9: 103. (2015).
[240]
Skaper SD, Facci L, Zusso M, Giusti P. An inflammation-centric view of neurological disease: beyond the neuron. Front Cell Neurosci 12: 72. (2018).
[241]
Xanthos DN, Sandkühler J. Neurogenic neuroinflammation: inflammatory CNS reactions in response to neuronal activity. Nat Rev Neurosci 15(1): 43-53. (2014).
[242]
Perry VH, Nicoll JA, Holmes C. Microglia in neurodegenerative disease. Nat Rev Neurol 6(4): 193-201. (2010).
[243]
Eikelenboom P, Veerhuis R, Scheper W, Rozemuller AJ, van Gool WA, Hoozemans JJ. The significance of neuroinflammation in understanding Alzheimer’s disease. J Neural Transm (Vienna) 113(11): 1685-95. (2006).
[244]
Dursun E, Gezen-Ak D, Hanağası H, Bilgiç B, Lohmann E, Ertan S, et al. The interleukin 1 alpha, interleukin 1 beta, interleukin 6 and alpha-2-macroglobulin serum levels in patients with early or late onset Alzheimer’s disease, mild cognitive impairment or Parkinson’s disease. J Neuroimmunol 283: 50-7. (2015).
[245]
Solfrizzi V, D’Introno A, Colacicco AM, Capurso C, Todarello O, Pellicani V, et al. Circulating biomarkers of cognitive decline and dementia. Clin Chim Acta 364(1-2): 91-112. (2006).
[246]
Panza F, Frisardi V, Seripa D, Imbimbo BP, Sancarlo D, D’Onofrio G, et al. Metabolic syndrome, mild cognitive impairment, and dementia. Curr Alzheimer Res 8(5): 492-509. (2011).
[247]
Holmes C, Cunningham C, Zotova E, et al. Systemic inflammation and disease progression in Alzheimer disease. Neurology 73(10): 768-74. (2009).
[248]
Kamer AR, Craig RG, Dasanayake AP, Brys M, Glodzik-Sobanska L, de Leon MJ. Inflammation and Alzheimer’s disease: possible role of periodontal diseases. Alzheimers Dement 4(4): 242-50. (2008).
[249]
Kamer AR, Craig RG, Pirraglia E, Dasanayake AP, Norman RG, Boylan RJ, et al. TNF-alpha and antibodies to periodontal bacteria discriminate between Alzheimer’s disease patients and normal subjects. J Neuroimmunol 216(1-2): 92-7. (2009).
[250]
Guerreiro R, Wojtas A, Bras J, Carrasquillo M, Rogaeva E, Majounie E, et al. TREM2 variants in Alzheimer’s disease. N Engl J Med 368(2): 117-27. (2013).
[251]
Jonsson T, Stefansson H, Steinberg S, Jonsdottir I, Jonsson PV, Snaedal J, et al. Variant of TREM2 associated with the risk of Alzheimer’s disease. N Engl J Med 368(2): 107-16. (2013).
[252]
Rosenberg RN, Lambracht-Washington D, Yu G, Xia W. Genomics of Alzheimer disease: a review. JAMA Neurol 73(7): 867-74. (2016).
[253]
Efthymiou AG, Goate AM. Late onset Alzheimer’s disease genetics implicates microglial pathways in disease risk. Mol Neurodegener 12(1): 43. (2017).
[254]
Song W, Hooli B, Mullin K, Jin SC, Cella M, Ulland TK, et al. Alzheimer’s disease-associated TREM2 variants exhibit either decreased or increased ligand-dependent activation. Alzheimers Dement 13(4): 381-7. (2017).
[255]
Jay TR, Hirsch AM, Broihier ML, Miller CM, Neilson LE, Ransohoff RM, et al. Disease progression-dependent effects of TREM2 deficiency in a mouse model of Alzheimer’s disease. J Neurosci 37(3): 637-47. (2017).
[256]
Takahashi K, Rochford CD, Neumann H. Clearance of apoptotic neurons without inflammation by microglial triggering receptor expressed on myeloid cells-2. J Exp Med 201(4): 647-57. (2005).
[257]
Hsieh CL, Koike M, Spusta SC, Niemi EC, Yenari M, Nakamura MC, et al. A role for TREM2 ligands in the phagocytosis of apoptotic neuronal cells by microglia. J Neurochem 109(4): 1144-56. (2009).
[258]
Plescher M, Seifert G, Hansen JN, Bedner P, Steinhäuser C, Halle A. Plaque-dependent morphological and electrophysiological heterogeneity of microglia in an Alzheimer’s disease mouse model. Glia 66(7): 1464-80. (2018).
[259]
Krabbe G, Halle A, Matyash V, Rinnenthal JL, Eom GD, Bernhardt U, et al. Functional impairment of microglia coincides with Beta-amyloid deposition in mice with Alzheimer-like pathology. PLoS One 8(4)e60921 (2013).
[260]
Wendt S, Maricos M, Vana N, Meyer N, Guneykaya D, Semtner M, et al. Changes in phagocytosis and potassium channel activity in microglia of 5xFAD mice indicate alterations in purinergic signaling in a mouse model of Alzheimer’s disease. Neurobiol Aging 58: 41-53. (2017).
[261]
Mullington J, Korth C, Hermann DM, Orth A, Galanos C, Holsboer F, et al. Dose-dependent effects of endotoxin on human sleep. Am J Physiol Regul Integr Comp Physiol 278(4): R947-55. (2000).
[262]
Schmidt FM, Pschiebl A, Sander C, Kirkby KC, Thormann J, Minkwitz J, et al. Impact of serum cytokine levels on eeg-measured arousal regulation in patients with major depressive disorder and healthy controls. Neuropsychobiology 73(1): 1-9. (2016).
[263]
Darko DF, Miller JC, Gallen C, White J, Koziol J, Brown SJ, et al. Sleep electroencephalogram delta-frequency amplitude, night plasma levels of tumor necrosis factor alpha, and human immunodeficiency virus infection. Proc Natl Acad Sci USA 92(26): 12080-4. (1995).
[264]
Kapás L, Hansen MK, Chang HY, Krueger JM. Vagotomy attenuates but does not prevent the somnogenic and febrile effects of lipopolysaccharide in rats. Am J Physiol 274(2 Pt 2): R406-11. (1998).
[265]
Majde JA, Krueger JM. Links between the innate immune system and sleep. J Allergy Clin Immunol 116(6): 1188-98. (2005).
[266]
Dias de Sousa M, Bonatti RCFI, Rodrigues Jr V.I.I., Azevedo DSI, Santos MHAI, Pereira ROLI, et al. Cytokines in cerebrospinal fluid of children with West síndrome. J Epilepsy Clin Neurophysiol 18(2): 63-6. (2012).
[267]
Fukumoto Y, Okumura A, Hayakawa F, Suzuki M, Kato T, Watanabe K, et al. Serum levels of cytokines and EEG findings in children with influenza associated with mild neurological complications. Brain Dev 29(7): 425-30. (2007).
[268]
Kapás L, Bohnet SG, Traynor TR, Majde JA, Szentirmai E, Magrath P, et al. Spontaneous and influenza virus-induced sleep are altered in TNF-alpha double-receptor deficient mice. J Appl Physiol 105(4): 1187-98. (2008).
[269]
Cissé Y, Wang S, Inoue I, Kido H. Rat model of influenza-associated encephalopathy (IAE): studies of electroencephalogram (EEG) in vivo. Neuroscience 165(4): 1127-37. (2010).
[270]
Zhao J, Chen F, Ren M, Li L, Li A, Jing B, et al. Low-frequency fluctuation characteristics in rhesus macaques with SIV infection: a resting-state fMRI study. J Neurovirol 25(2): 141-9. (2018).
[271]
Riazi K, Galic MA, Kuzmiski JB, Ho W, Sharkey KA, Pittman QJ. Microglial activation and TNFalpha production mediate altered CNS excitability following peripheral inflammation. Proc Natl Acad Sci USA 105(44): 17151-6. (2008).
[272]
Nisticò R, Mango D, Mandolesi G, Piccinin S, Berretta N, Pignatelli M, et al. Inflammation subverts hippocampal synaptic plasticity in experimental multiple sclerosis. PLoS One 8(1)e54666 (2013).
[273]
Gao R, Ji MH, Gao DP, Yang RH, Zhang SG, Yang JJ, et al. Neuroinflammation-induced downregulation of hippocampacal neuregulin 1-erbb4 signaling in the parvalbumin interneurons might contribute to cognitive impairment in a mouse model of sepsis-associated encephalopathy. Inflammation 40(2): 387-400. (2017).
[274]
Harris-Bozer AL, Peng YB. Inflammatory pain by carrageenan recruits low-frequency local field potential changes in the anterior cingulate cortex. Neurosci Lett 632: 8-14. (2016).
[275]
Wang J, Wang J, Xing GG, Li X, Wan Y. Enhanced gamma oscillatory activity in rats with chronic inflammatory pain. Front Neurosci 10: 489. (2016).
[276]
Chen Z, Shen X. Huang, Wu H, Zhang M. Membrane potential synchrony of neurons in anterior cingulate cortex plays a pivotal role in generation of neuropathic pain. Sci Rep 8(1): 1691. (2018).
[277]
Pollmächer T, Schreiber W, Gudewill S, Vedder H, Fassbender K, Wiedemann K, et al. Influence of endotoxin on nocturnal sleep in humans. Am J Physiol 264(6 Pt 2): R1077-83. (1993).
[278]
Trachsel L, Schreiber W, Holsboer F, Pollmächer T. Endotoxin enhances EEG alpha and beta power in human sleep. Sleep 17(2): 132-9. (1994).
[279]
van den Boogaard M, Ramakers BP, van Alfen N, van der Werf SP, Fick WF, Hoedemaekers CW, et al. Endotoxemia-induced inflammation and the effect on the human brain. Crit Care 14(3): R81. (2010).
[280]
Krueger JM, Kubillus S, Shoham S, Davenne D. Enhancement of slow-wave sleep by endotoxin and lipid A. Am J Physiol 251(3 Pt 2): R591-7. (1986).
[281]
Lancel M, Crönlein J, Müller-Preuss P, Holsboer F. Lipopolysaccharide increases EEG delta activity within non-REM sleep and disrupts sleep continuity in rats. Am J Physiol 268(5 Pt 2): R1310-8. (1995).
[282]
Schiffelholz T, Lancel M. Sleep changes induced by lipopolysaccharide in the rat are influenced by age. Am J Physiol Regul Integr Comp Physiol 280(2): R398-403. (2001).
[283]
Ashley NT, Zhang N, Weil ZM, Magalang UJ, Nelson RJ. Photoperiod alters duration and intensity of non-rapid eye movement sleep following immune challenge in Siberian hamsters (Phodopus sungorus). Chronobiol Int 29(6): 683-92. (2012).
[284]
Albrecht MA, Vaughn CN, Erickson MA, Clark SM, Tonelli LH. Time and frequency dependent changes in resting state EEG functional connectivity following lipopolysaccharide challenge in rats. PLoS One 13(11) :e0206985 (2018).
[285]
Ingiosi AM, Opp MR. Sleep and immunomodulatory responses to systemic lipopolysaccharide in mice selectively expressing interleukin-1 receptor 1 on neurons or astrocytes. Glia 64(5): 780-91. (2016).
[286]
Mamad O, Islam MN, Cunningham C, Tsanov M. Differential response of hippocampal and prefrontal oscillations to systemic LPS application. Brain Res 1681: 64-74. (2018).
[287]
Chen Z, Jalabi W, Hu W, Park HJ, Gale JT, Kidd GJ, et al. Microglial displacement of inhibitory synapses provides neuroprotection in the adult brain. Nat Commun 5: 4486. (2014).
[288]
Papageorgiou IE, Lewen A, Galow LV, Cesetti T, Scheffel J, Regen T, et al. TLR4-activated microglia require IFN-γ to induce severe neuronal dysfunction and death in situ. Proc Natl Acad Sci USA 113(1): 212-7. (2016).
[289]
Ta TT, Dikmen HO, Schilling S, Chausse B, Lewen A, Hollnagel JO, et al. Priming of microglia with IFN-γ slows neuronal gamma oscillations in situ. Proc Natl Acad Sci USA 116(10): 4637-42. (2019).
[290]
Gullo F, Amadeo A, Donvito G, Lecchi M, Costa B, Constanti A, et al. Atypical “seizure-like” activity in cortical reverberating networks in vitro can be caused by LPS-induced inflammation: a multi-electrode array study from a hundred neurons. Front Cell Neurosci 8: 361. (2014).
[291]
Kovács Z, Czurkó A, Kékesi KA, Juhász G. Intracerebroventricularly administered lipopolysaccharide enhances spike-wave discharges in freely moving WAG/Rij rats. Brain Res Bull 85(6): 410-6. (2011).
[292]
Kovács Z, Dobolyi A, Juhász G, Kékesi KA. Lipopolysaccharide induced increase in seizure activity in two animal models of absence epilepsy WAG/Rij and GAERS rats and Long Evans rats. Brain Res Bull 104: 7-18. (2014).
[293]
Dean JM, van de Looij Y, Sizonenko SV, Lodygensky GA, Lazeyras F, Bolouri H, et al. Delayed cortical impairment following lipopolysaccharide exposure in preterm fetal sheep. Ann Neurol 70(5): 846-56. (2011).
[294]
Keogh MJ, Bennet L, Drury PP, Booth LC, Mathai S, Naylor AS, et al. Subclinical exposure to low-dose endotoxin impairs EEG maturation in preterm fetal sheep. Am J Physiol Regul Integr Comp Physiol 303(3): R270-8. (2012).
[295]
Gavilanes AW, Gantert M, Strackx E, Zimmermann LJ, Seeldrayers S, Vles JS, et al. Increased EEG delta frequency corresponds to chorioamnionitis-related brain injury. Front Biosci (Schol Ed) 2: 432-8. (2010).
[296]
Kim KM, Zamaleeva AI, Lee YW, Ahmed MR, Kim E, Lee HR, et al. Characterization of brain dysfunction induced by bacterial lipopeptides that alter neuronal activity and network in rodent brains. J Neurosci 38(50): 10672-91. (2018).
[297]
Costello DA, Lynch MA. Toll-like receptor 3 activation modulates hippocampal network excitability, via glial production of interferon-β. Hippocampus 23(8): 696-707. (2013).
[298]
Galic MA, Riazi K, Henderson AK, Tsutsui S, Pittman QJ. Viral-like brain inflammation during development causes increased seizure susceptibility in adult rats. Neurobiol Dis 36(2): 343-51. (2009).
[299]
Ducharme G, Lowe GC, Goutagny R, Williams S. Early alterations in hippocampal circuitry and theta rhythm generation in a mouse model of prenatal infection: implications for schizophrenia. PLoS Ono 7(1)e29754 (2012).
[300]
Späth-Schwalbe E, Hansen K, Schmidt F, Schrezenmeier H, Marshall L, Burger K, et al. Acute effects of recombinant human interleukin-6 on endocrine and central nervous sleep functions in healthy men. J Clin Endocrinol Metab 83(5): 1573-9. (1998).
[301]
Benedict C, Scheller J, Rose-John S, Born J, Marshall L. Enhancing influence of intranasal interleukin-6 on slow-wave activity and memory consolidation during sleep. FASEB J 23(10): 3629-36. (2009).
[302]
May U, Schiffelholz T, Baier PC, Krueger JM, Rose-John S, Scheller J. IL-6-trans-signalling increases rapid-eye-movement sleep in rats. Eur J Pharmacol 613(1-3): 141-5. (2009).
[303]
Oyanedel CN, Kelemen E, Scheller J, Born J, Rose-John S. Peripheral and central blockade of interleukin-6 trans-signaling differentially affects sleep architecture. Brain Behav Immun 50: 178-85. (2015).
[304]
Clarkson BDS, Kahoud RJ, McCarthy CB, Howe CL. Inflammatory cytokine-induced changes in neural network activity measured by waveform analysis of high-content calcium imaging in murine cortical neurons. Sci Rep 7(1): 9037. (2017).
[305]
Shoham S, Davenne D, Cady AB, Dinarello CA, Krueger JM. Recombinant tumor necrosis factor and interleukin 1 enhance slow-wave sleep. Am J Physiol 253(1 Pt 2): R142-9. (1987).
[306]
Kubota T, Fang J, Guan Z, Brown RA, Krueger JM. Vagotomy attenuates tumor necrosis factor-alpha-induced sleep and EEG delta-activity in rats. Am J Physiol Regul Integr Comp Physiol 280(4): R1213-20. (2001).
[307]
Kubota T, Li N, Guan Z, Brown RA, Krueger JM. Intrapreoptic microinjection of TNF-alpha enhances non-REM sleep in rats. Brain Res 932(1-2): 37-44. (2002).
[308]
Fang J, Wang Y, Krueger JM. Mice lacking the TNF 55 kDa receptor fail to sleep more after TNFalpha treatment. J Neurosci 17(15): 5949-55. (1997).
[309]
Deboer T, Fontana A, Tobler I. Tumor necrosis factor (TNF) ligand and TNF receptor deficiency affects sleep and the sleep EEG. J Neurophysiol 88(2): 839-46. (2002).
[310]
Terao A, Matsumura H, Yoneda H, Saito M. Enhancement of slow-wave sleep by tumor necrosis factor-alpha is mediated by cyclooxygenase-2 in rats. Neuroreport 9(17): 3791-6. (1998).
[311]
Zielinski MR, Dunbrasky DL, Taishi P, Souza G, Krueger JM. Vagotomy attenuates brain cytokines and sleep induced by peripherally administered tumor necrosis factor-α and lipopolysaccharide in mice Sleep 36(8): 1227-38, 1238A (2013).
[312]
Yoshida H, Peterfi Z, García-García F, Kirkpatrick R, Yasuda T, Krueger JM. State-specific asymmetries in EEG slow wave activity induced by local application of TNFalpha. Brain Res 1009(1-2): 129-36. (2004).
[313]
Churchill L, Yasuda K, Yasuda T, Blindheim KA, Falter M, Garcia-Garcia F, et al. Unilateral cortical application of tumor necrosis factor alpha induces asymmetry in Fos- and interleukin-1beta-immunoreactive cells within the corticothalamic projection. Brain Res 1055(1-2): 15-24. (2005).
[314]
Taishi P, Churchill L, Wang M, Kay D, Davis CJ, Guan X, et al. TNFalpha siRNA reduces brain TNF and EEG delta wave activity in rats. Brain Res 1156: 125-32. (2007).
[315]
Richter F, Lütz W, Eitner A, Leuchtweis J, Lehmenkühler A, Schaible HG. Tumor necrosis factor reduces the amplitude of rat cortical spreading depression in vivo. Ann Neurol 76(1): 43-53. (2014).
[316]
Tobler I, Borbély AA, Schwyzer M, Fontana A. Interleukin-1 derived from astrocytes enhances slow wave activity in sleep EEG of the rat. Eur J Pharmacol 104(1-2): 191-2. (1984).
[317]
Opp MR, Obal F Jr, Krueger JM. Interleukin 1 alters rat sleep: temporal and dose-related effects. Am J Physiol 260(1 Pt 2): R52-8. (1991).
[318]
Lancel M, Mathias S, Faulhaber J, Schiffelholz T. Effect of interleukin-1 beta on EEG power density during sleep depends on circadian phase. Am J Physiol 270(4 Pt 2): R830-7. (1996).
[319]
Hansen MK, Krueger JM. Subdiaphragmatic vagotomy blocks the sleep- and fever-promoting effects of interleukin-1beta. Am J Physiol 273(4 Pt 2): R1246-53. (1997).
[320]
Richter F, Eitner A, Leuchtweis J, Bauer R, Lehmenkühler A, Schaible HG. Effects of interleukin-1ß on cortical spreading depolarization and cerebral vasculature. J Cereb Blood Flow Metab 37(5): 1791-802. (2017).
[321]
Bray N, Burrows FE, Jones M, Berwick J, Allan SM, Schiessl I. Decreased haemodynamic response and decoupling of cortical gamma-band activity and tissue oxygen perfusion after striatal interleukin-1 injection. J Neuroinflammation 13(1): 195. (2016).
[322]
Nimmerjahn A, Kirchhoff F, Helmchen F. Resting microglial cells are highly dynamic surveillants of brain parenchyma in vivo. Science308( 5726): 1314-8 (2005).
[323]
Wake H, Moorhouse AJ, Jinno S, Kohsaka S, Nabekura J. Resting microglia directly monitor the functional state of synapses in vivo and determine the fate of ischemic terminals. J Neurosci 29(13): 3974-80. (2009).
[324]
Dissing-Olesen L, LeDue JM, Rungta RL, Hefendehl JK, Choi HB, MacVicar BA. Activation of neuronal NMDA receptors triggers transient ATP-mediated microglial process outgrowth. J Neurosci 34(32): 10511-27. (2014).
[325]
Tremblay MÈ, Lowery RL, Majewska AK. Microglial interactions with synapses are modulated by visual experience. PLoS Biol 8(11)e1000527 (2010).
[326]
Akiyoshi R, Wake H, Kato D, Horiuchi H, Ono R, Ikegami A, et al. Microglia enhance synapse activity to promote local network synchronization. eNeuro 5(5): pii: ENEURO.0088-18.2018 (2018).
[327]
Lorea-Hernández JJ, Morales T, Rivera-Angulo AJ, Alcantara-Gonzalez D, Peña-Ortega F. Microglia modulate respiratory rhythm generation and autoresuscitation. Glia 64(4): 603-19. (2016).
[328]
Camacho-Hernández NP, Lorea-Hernández JJ, Peña-Ortega F. Microglial modulators reduce respiratory rhythm long-term facilitation in vitro. Respir Physiol Neurobiol 265: 9-18. (2019).
[329]
Pardo-Peña K, Lorea-Hernández JJ, Camacho-Hernández NP, et al. Hydrogen peroxide extracellular concentration in the ventrolateral medulla and its increase in response to hypoxia in vitro: possible role of microglia. Brain Res 1692: 87-99. (2018).
[330]
Ames C, Boland E, Szentirmai É. Effects of macrophage depletion on sleep in mice. PLoS One 11(7)e0159812 (2016).
[331]
Wang X, Zhao L, Zhang J, Fariss RN, Ma W, Kretschmer F, et al. Requirement for microglia for the maintenance of synaptic function and integrity in the mature retina. J Neurosci 36(9): 2827-42. (2016).
[332]
Szalay G, Martinecz B, Lénárt N, Környei Z, Orsolits B, Judák L, et al. Microglia protect against brain injury and their selective elimination dysregulates neuronal network activity after stroke. Nat Commun 7: 11499. (2016).
[333]
Chini M, Lindemann C, Poepplau JA, Xu X, Ahlbeck J, Bitzenhofer SH, et al. Microglia inhibition rescues developmental hypofrontality in a mouse model of mental illness. bioRxiv. 254656v1.
[334]
Zhan Y, Paolicelli RC, Sforazzini F, Weinhard L, Bolasco G, Pagani F, et al. Deficient neuron-microglia signaling results in impaired functional brain connectivity and social behavior. Nat Neurosci 17(3): 400-6. (2014).
[335]
Zhan Y. Theta frequency prefrontal-hippocampal driving relationship during free exploration in mice. Neuroscience 300: 554-65. (2015).
[336]
Rodgers KM, Hutchinson MR, Northcutt A, Maier SF, Watkins LR, Barth DS. The cortical innate immune response increases local neuronal excitability leading to seizures. Brain 132(Pt 9): 2478-86. (2009).
[337]
Park KI, Dzhala V, Saponjian Y, Staley KJ. What elements of the inflammatory system are necessary for epileptogenesis in vitro? eNeuro 2(2): pii: ENEURO.0027-14.2015 (2015).
[338]
Zhao H, Zhu C, Huang D. Microglial activation: an important process in the onset of epilepsy. Am J Transl Res 10(9): 2877-89. (2018).
[339]
Zhang B, Zou J, Han L, Beeler B, Friedman JL, Griffin E, et al. The specificity and role of microglia in epileptogenesis in mouse models of tuberous sclerosis complex. Epilepsia 59(9): 1796-806. (2018).
[340]
Martinez-Losa M, Tracy TE, Ma K, Verret L, Clemente-Perez A, Khan AS, et al. Nav1.1-overexpressing interneuron transplants restore brain rhythms and cognition in a mouse model of Alzheimer’s disease. Neuron 98(1): 75-89.e5. (2018).
[341]
Zhu H, Yan H, Tang N, Li X, Pang P, Li H, et al. Impairments of spatial memory in an Alzheimer’s disease model via degeneration of hippocampal cholinergic synapses. Nat Commun 8(1): 1676. (2017).
[342]
Haffen E, Chopard G, Pretalli JB, Magnin E, Nicolier M, Monnin J, et al. A case report of daily left prefrontal repetitive transcranial magnetic stimulation (rTMS) as an adjunctive treatment for Alzheimer disease. Brain Stimul 5(3): 264-6. (2012).
[343]
Alcantara-Gonzalez D, Villasana-Salazar B, Peña-Ortega F. Single amyloid-beta injection exacerbates 4-aminopyridine-induced seizures and changes synaptic coupling in the hippocampus. Hippocampus 29(12): 1150-64. (2019).

Rights & Permissions Print Cite
© 2024 Bentham Science Publishers | Privacy Policy