Striatal spatial heterogeneity, clustering, and white matter association of GFAP+ astrocytes in a mouse model of Huntington’s disease

Introduction Huntington’s disease (HD) is a neurodegenerative disease that primarily affects the striatum, a brain region that controls movement and some forms of cognition. Neuronal dysfunction and loss in HD is accompanied by increased astrocyte density and astrocyte pathology. Astrocytes are a heterogeneous population classified into multiple subtypes depending on the expression of different gene markers. Studying whether mutant Huntingtin (HTT) alters specific subtypes of astrocytes is necessary to understand their relative contribution to HD. Methods Here, we studied whether astrocytes expressing two different markers; glial fibrillary acidic protein (GFAP), associated with astrocyte activation, and S100 calcium-binding protein B (S100B), a marker of matured astrocytes and inflammation, were differentially altered in HD. Results First, we found three distinct populations in the striatum of WT and symptomatic zQ175 mice: GFAP+, S100B+, and dual GFAP+S100B+. The number of GFAP+ and S100B+ astrocytes throughout the striatum was increased in HD mice compared to WT, coinciding with an increase in HTT aggregation. Overlap between GFAP and S100B staining was expected, but dual GFAP+S100B+ astrocytes only accounted for less than 10% of all tested astrocytes and the number of GFAP+S100B+ astrocytes did not differ between WT and HD, suggesting that GFAP+ astrocytes and S100B+ astrocytes are distinct types of astrocytes. Interestingly, a spatial characterization of these astrocyte subtypes in HD mice showed that while S100B+ were homogeneously distributed throughout the striatum, GFAP+ preferentially accumulated in “patches” in the dorsomedial (dm) striatum, a region associated with goal-directed behaviors. In addition, GFAP+ astrocytes in the dm striatum of zQ175 mice showed increased clustering and association with white matter fascicles and were preferentially located in areas with low HTT aggregate load. Discussion In summary, we showed that GFAP+ and S100B+ astrocyte subtypes are distinctly affected in HD and exist in distinct spatial arrangements that may offer new insights to the function of these specific astrocytes subtypes and their potential implications in HD pathology.

[1]  B. Khakh,et al.  Neuronal and astrocytic contributions to Huntington's disease dissected with zinc finger protein transcriptional repressors. , 2023, Cell reports.

[2]  D. Hu,et al.  The neurons in mouse V1 show different degrees of spatial clustering , 2022, Brain Research Bulletin.

[3]  J. Lucas,et al.  CK2 alpha prime and alpha-synuclein pathogenic functional interaction mediates synaptic dysregulation in Huntington’s disease , 2022, Acta Neuropathologica Communications.

[4]  T. V. Johnson,et al.  Analyses of transplanted human retinal ganglion cell morphology and localization in murine organotypic retinal explant culture , 2022, STAR protocols.

[5]  Luciano da F. Costa,et al.  Maternal high-fat diet in mice induces cerebrovascular, microglial and long-term behavioural alterations in offspring , 2022, Communications Biology.

[6]  OUP accepted manuscript , 2022, Brain.

[7]  M. Qiu,et al.  S100B is selectively expressed by gray matter protoplasmic astrocytes and myelinating oligodendrocytes in the developing CNS , 2021, Molecular Brain.

[8]  M. Götz,et al.  Molecular diversity of diencephalic astrocytes reveals adult astrogenesis regulated by Smad4 , 2021, The EMBO journal.

[9]  J. Wityak,et al.  Pharmacological characterization of mutant huntingtin aggregate-directed PET imaging tracer candidates , 2021, Scientific Reports.

[10]  G. Halliday,et al.  Early white matter pathology in the fornix of the limbic system in Huntington disease , 2021, Acta Neuropathologica.

[11]  S. Liddelow,et al.  Neuroinflammatory astrocyte subtypes in the mouse brain , 2021, Nature Neuroscience.

[12]  A. Benraiss,et al.  Cell-intrinsic glial pathology is conserved across human and murine models of Huntington's disease. , 2021, Cell reports.

[13]  Kira E. Poskanzer,et al.  Reactive astrocyte nomenclature, definitions, and future directions , 2021, Nature Neuroscience.

[14]  B. Balleine,et al.  The dorsomedial striatum: an optimal cellular environment for encoding and updating goal-directed learning , 2021, Current Opinion in Behavioral Sciences.

[15]  B. Eggen,et al.  Transcriptional heterogeneity between primary adult grey and white matter astrocytes underlie differences in modulation of in vitro myelination , 2020, Journal of Neuroinflammation.

[16]  B. Khakh,et al.  Context-Specific Striatal Astrocyte Molecular Responses Are Phenotypically Exploitable , 2020, Neuron.

[17]  A. Reiner,et al.  Progression of basal ganglia pathology in heterozygous Q175 knock‐in Huntington's disease mice , 2020, The Journal of comparative neurology.

[18]  Derek K. Jones,et al.  A Critical Review of White Matter Changes in Huntington’s Disease , 2020, Movement disorders : official journal of the Movement Disorder Society.

[19]  C. Lohr,et al.  Age-Dependent Heterogeneity of Murine Olfactory Bulb Astrocytes , 2020, Frontiers in Aging Neuroscience.

[20]  J. Grutzendler,et al.  Astrocytes and microglia play orchestrated roles and respect phagocytic territories during neuronal corpse removal in vivo , 2020, Science Advances.

[21]  R. Voituriez,et al.  Cell clusters adopt a collective amoeboid mode of migration in confined nonadhesive environments , 2020, bioRxiv.

[22]  F. Artigas,et al.  M2 cortex-dorsolateral striatum stimulation reverses motor symptoms and synaptic deficits in Huntington’s disease , 2020, bioRxiv.

[23]  Chris P. Ponting,et al.  Identification of region-specific astrocyte subtypes at single cell resolution , 2020, Nature Communications.

[24]  Gurrein K. Madan,et al.  Genome-wide In Vivo CNS Screening Identifies Genes that Modify CNS Neuronal Survival and mHTT Toxicity , 2020, Neuron.

[25]  H. Yin,et al.  Striatal Projection Neurons Require Huntingtin for Synaptic Connectivity and Survival. , 2020, Cell reports.

[26]  Stephen J. Eglen,et al.  From random to regular: Variation in the patterning of retinal mosaics* , 2019, The Journal of comparative neurology.

[27]  M. Tremblay,et al.  Immunofluorescence Staining Using IBA1 and TMEM119 for Microglial Density, Morphology and Peripheral Myeloid Cell Infiltration Analysis in Mouse Brain. , 2019, Journal of visualized experiments : JoVE.

[28]  B. Khakh,et al.  Astrocyte molecular signatures in Huntington’s disease , 2019, Science Translational Medicine.

[29]  V. Menon,et al.  Single-nucleus RNA-seq identifies Huntington disease astrocyte states , 2019, Acta Neuropathologica Communications.

[30]  A. Citri,et al.  Dorsal Striatal Circuits for Habits, Compulsions and Addictions , 2019, Front. Syst. Neurosci..

[31]  B. Khakh,et al.  The Emerging Nature of Astrocyte Diversity. , 2019, Annual review of neuroscience.

[32]  M. Zoli,et al.  Distribution and Relative Abundance of S100 Proteins in the Brain of the APP23 Alzheimer’s Disease Model Mice , 2019, Front. Neurosci..

[33]  F. Michetti,et al.  The S100B story: from biomarker to active factor in neural injury , 2018, Journal of neurochemistry.

[34]  C. Cepeda,et al.  Mutant huntingtin reduction in astrocytes slows disease progression in the BACHD conditional Huntington's disease mouse model , 2018, Human molecular genetics.

[35]  A. Verkhratsky,et al.  Neuroglia in Neurodegenerative Diseases , 2019, Advances in Experimental Medicine and Biology.

[36]  H. Künzle,et al.  Radial glial elements in the cerebral cortex of the lesser hedgehog tenrec , 2018, Brain Structure and Function.

[37]  F. Cicchetti,et al.  Mutant huntingtin protein expression and blood–spinal cord barrier dysfunction in huntington disease , 2017, Annals of neurology.

[38]  B. Khakh,et al.  Unravelling and Exploiting Astrocyte Dysfunction in Huntington’s Disease , 2017, Trends in Neurosciences.

[39]  S. Giannetti,et al.  The Astrocytic S100B Protein with Its Receptor RAGE Is Aberrantly Expressed in SOD1G93A Models, and Its Inhibition Decreases the Expression of Proinflammatory Genes , 2017, Mediators of inflammation.

[40]  S. Goto,et al.  Striatal Vulnerability in Huntington’s Disease: Neuroprotection Versus Neurotoxicity , 2017, Brain sciences.

[41]  Sarah R. Heilbronner,et al.  Organization of the Anterior Limb of the Internal Capsule in the Rat , 2017, The Journal of Neuroscience.

[42]  D. Thiele,et al.  Abnormal degradation of the neuronal stress-protective transcription factor HSF1 in Huntington's disease , 2017, Nature Communications.

[43]  Manoj Kumar,et al.  INGE GRUNDKE-IQBAL AWARD FOR ALZHEIMER’S RESEARCH: NEUROTOXIC REACTIVE ASTROCYTES ARE INDUCED BY ACTIVATED MICROGLIA , 2019, Alzheimer's & Dementia.

[44]  K. Nave,et al.  Inducible targeting of CNS astrocytes in Aldh1l1-CreERT2 BAC transgenic mice , 2016, F1000Research.

[45]  Tianyi Mao,et al.  A comprehensive excitatory input map of the striatum reveals novel functional organization , 2016, eLife.

[46]  Nicholas N. Foster,et al.  The mouse cortico-striatal projectome , 2016, Nature Neuroscience.

[47]  Jihwan Song,et al.  Human-to-mouse prion-like propagation of mutant huntingtin protein , 2016, Acta Neuropathologica.

[48]  R. Roos,et al.  Detection of Motor Changes in Huntington's Disease Using Dynamic Causal Modeling , 2015, Front. Hum. Neurosci..

[49]  E. Hol,et al.  Glial fibrillary acidic protein (GFAP) and the astrocyte intermediate filament system in diseases of the central nervous system. , 2015, Current opinion in cell biology.

[50]  M. Hayden,et al.  Huntington disease , 2015, Nature Reviews Disease Primers.

[51]  P. Barker,et al.  S100B protein activates a RAGE‐dependent autocrine loop in astrocytes: implications for its role in the propagation of reactive gliosis , 2014, Journal of neurochemistry.

[52]  M. Nedergaard,et al.  White matter astrocytes in health and disease , 2014, Neuroscience.

[53]  Guy M McKhann,et al.  Phenotypic Heterogeneity and Plasticity of Isocortical and Hippocampal Astrocytes in the Human Brain , 2014, The Journal of Neuroscience.

[54]  Amin Derouiche,et al.  Glutamine Synthetase as an Astrocytic Marker: Its Cell Type and Vesicle Localization , 2013, Front. Endocrinol..

[55]  Manoj K. Gottipati,et al.  Enhanced Ca2+-dependent glutamate release from astrocytes of the BACHD Huntington's disease mouse model , 2013, Neurobiology of Disease.

[56]  Rui Costa,et al.  Premotor cortex is critical for goal-directed actions , 2013, Front. Comput. Neurosci..

[57]  Milos Pekny,et al.  Versatile and Simple Approach to Determine Astrocyte Territories in Mouse Neocortex and Hippocampus , 2013, PloS one.

[58]  George V. Rebec,et al.  Role of cerebral cortex in the neuropathology of Huntington's disease , 2013, Front. Neural Circuits.

[59]  Mei Kwan,et al.  Comprehensive Behavioral and Molecular Characterization of a New Knock-In Mouse Model of Huntington’s Disease: zQ175 , 2012, PloS one.

[60]  Jack R Glaser,et al.  Characterization of Neurophysiological and Behavioral Changes, MRI Brain Volumetry and 1H MRS in zQ175 Knock-In Mouse Model of Huntington's Disease , 2012, PloS one.

[61]  S. Itohara,et al.  S100B is increased in Parkinson’s disease and ablation protects against MPTP-induced toxicity through the RAGE and TNF-α pathway , 2012, Brain : a journal of neurology.

[62]  E. Hol,et al.  GFAP Isoforms in Adult Mouse Brain with a Focus on Neurogenic Astrocytes and Reactive Astrogliosis in Mouse Models of Alzheimer Disease , 2012, PloS one.

[63]  J. Rothstein,et al.  Molecular comparison of GLT1+ and ALDH1L1+ astrocytes in vivo in astroglial reporter mice , 2011, Glia.

[64]  R. Roos,et al.  Huntington's disease: a clinical review , 2010, Orphanet journal of rare diseases.

[65]  Ben A. Barres,et al.  Regulation of synaptic connectivity by glia , 2010, Nature.

[66]  M. Guillermier,et al.  In vivo expression of polyglutamine-expanded huntingtin by mouse striatal astrocytes impairs glutamate transport: a correlation with Huntington's disease subjects , 2010, Human molecular genetics.

[67]  D. Oorschot,et al.  Cell loss in the motor and cingulate cortex correlates with symptomatology in Huntington's disease. , 2010, Brain : a journal of neurology.

[68]  Peng Yu,et al.  Altered white matter microstructure in the corpus callosum in Huntington's disease: Implications for cortical “disconnection” , 2010, NeuroImage.

[69]  Stefan Klöppel,et al.  Functional compensation of motor function in pre-symptomatic Huntington's disease , 2009, Brain : a journal of neurology.

[70]  Magdalena Götz,et al.  Origin and progeny of reactive gliosis: A source of multipotent cells in the injured brain , 2008, Proceedings of the National Academy of Sciences.

[71]  David S Tuch,et al.  Diffusion tensor imaging in presymptomatic and early Huntington's disease: Selective white matter pathology and its relationship to clinical measures , 2006, Movement disorders : official journal of the Movement Disorder Society.

[72]  B. Harper Huntington Disease , 2005, Journal of the Royal Society of Medicine.

[73]  L. Eng,et al.  Glial Fibrillary Acidic Protein: GFAP-Thirty-One Years (1969–2000) , 2000, Neurochemical Research.

[74]  K. Tohyama,et al.  Nearest-neighbor distance of intermediate filaments in axons and schwann cells , 2004, Acta Neuropathologica.

[75]  K. Fenger,et al.  Molecular and behavioral analysis of the r6/1 huntington′s disease transgenic mouse , 2003, Neuroscience.

[76]  N. D. Mar,et al.  Differential Changes in Striatal Projection Neurons in R6/2 Transgenic Mice for Huntington's Disease , 2002, Neurobiology of Disease.

[77]  G. Halliday,et al.  Pyramidal Cell Loss in Motor Cortices in Huntington's Disease , 2002, Neurobiology of Disease.

[78]  K. Lindenberg,et al.  Impaired glutamate transport and glutamate-glutamine cycling: downstream effects of the Huntington mutation. , 2002, Brain : a journal of neurology.

[79]  D. Neary,et al.  Behavior in Huntington's disease: dissociating cognition-based and mood-based changes. , 2002, The Journal of neuropsychiatry and clinical neurosciences.

[80]  Mark Ellisman,et al.  Protoplasmic Astrocytes in CA1 Stratum Radiatum Occupy Separate Anatomical Domains , 2002, The Journal of Neuroscience.

[81]  R. Mrak,et al.  The role of activated astrocytes and of the neurotrophic cytokine S100B in the pathogenesis of Alzheimer’s disease , 2001, Neurobiology of Aging.

[82]  A. Mahal,et al.  Impaired Glutamate Uptake in the R6 Huntington's Disease Transgenic Mice , 2001, Neurobiology of Disease.

[83]  R. Reep,et al.  Topographic organization of the striatal and thalamic connections of rat medial agranular cortex , 1999, Brain Research.

[84]  Claire-Anne Gutekunst,et al.  Nuclear and Neuropil Aggregates in Huntington’s Disease: Relationship to Neuropathology , 1999, The Journal of Neuroscience.

[85]  A. Araque,et al.  Astrocyte-induced modulation of synaptic transmission. , 1999, Canadian journal of physiology and pharmacology.

[86]  S. Thompson Advances in experimental medicine and biology , 1996 .

[87]  W. Griffin,et al.  S100β protein expression in Alzheimer disease: Potential role in the pathogenesis of neuritic plaques , 1994 .

[88]  Manish S. Shah,et al.  A novel gene containing a trinucleotide repeat that is expanded and unstable on Huntington's disease chromosomes , 1993, Cell.

[89]  J. Vonsattel,et al.  Morphometric Demonstration of Atrophic Changes in the Cerebral Cortex, White Matter, and Neostriatum in Huntington's Disease , 1988, Journal of neuropathology and experimental neurology.

[90]  R. Ferrante,et al.  Neuropathological Classification of Huntington's Disease , 1985, Journal of neuropathology and experimental neurology.

[91]  D. Selkoe,et al.  Huntington's disease: Changes in striatal proteins reflect astrocytic gliosis , 1982, Brain Research.

[92]  J. Vanderhaeghen,et al.  An acidic protein isolated from fibrous astrocytes. , 1971, Brain research.