Next Article in Journal
Menthol Pretreatment Alleviates Campylobacter jejuni-Induced Enterocolitis in Human Gut Microbiota-Associated IL-10−/− Mice
Previous Article in Journal
Senescence Promotes the Recovery of Stemness among Cancer Cells via Reprograming
Previous Article in Special Issue
AMPKα1 Deficiency in Astrocytes from a Rat Model of ALS Is Associated with an Altered Metabolic Resilience
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Astrocytes: The Stars in Neurodegeneration?

by
Katarina Stoklund Dittlau
and
Kristine Freude
*
Department of Veterinary and Animal Sciences, Faculty of Health and Medical Sciences, University of Copenhagen, 1870 Frederiksberg, Denmark
*
Author to whom correspondence should be addressed.
Biomolecules 2024, 14(3), 289; https://doi.org/10.3390/biom14030289
Submission received: 29 January 2024 / Revised: 24 February 2024 / Accepted: 26 February 2024 / Published: 28 February 2024
(This article belongs to the Special Issue The Contribution of Astrocytes to Neuropathology)

Abstract

:
Today, neurodegenerative disorders like Alzheimer’s disease (AD), Parkinson’s disease (PD), frontotemporal dementia (FTD) and amyotrophic lateral sclerosis (ALS) affect millions of people worldwide, and as the average human lifespan increases, similarly grows the number of patients. For many decades, cognitive and motoric decline has been explained by the very apparent deterioration of neurons in various regions of the brain and spinal cord. However, more recent studies show that disease progression is greatly influenced by the vast population of glial cells. Astrocytes are traditionally considered star-shaped cells on which neurons rely heavily for their optimal homeostasis and survival. Increasing amounts of evidence depict how astrocytes lose their supportive functions while simultaneously gaining toxic properties during neurodegeneration. Many of these changes are similar across various neurodegenerative diseases, and in this review, we highlight these commonalities. We discuss how astrocyte dysfunction drives neuronal demise across a wide range of neurodegenerative diseases, but rather than categorizing based on disease, we aim to provide an overview based on currently known mechanisms. As such, this review delivers a different perspective on the disease causes of neurodegeneration in the hope to encourage further cross-disease studies into shared disease mechanisms, which might ultimately disclose potentially common therapeutic entry points across a wide panel of neurodegenerative diseases.

1. Introduction

Astrocytes received their name from Hungarian anatomist and histologist Michael von Lenhossék (Mihàly Lenhossék) in 1895 due to their mesmerizing star-like morphology [1]. Since then, multiple astrocytic subpopulations have been defined based on morphology, function and spatiotemporal distribution, and the portfolio is continuously growing from the emerging of single-cell and single-nuclei sequencing studies [2]. Now, it is evident that individual astrocytic subtypes might change during development and aging and in response to external stimuli in their immediate environment, which jointly changes the composition of the astrocyte population [2]. Similarly intriguing is the notion that individual astrocytes can exhibit both deleterious and protective properties in pathological scenarios and during disease progressions and that these functions might change in response to both intrinsic and external stimuli [3,4].
Much is already known about the astrocytes’ function during normal physiological conditions, which is an important foundation for the phenotypic understanding and extrapolation during diseases. Astrocytes are crucial for the integrity and function of the neuronal network, as they form tripartite synapses with neurons and thereby ensure activity modulation and neurotransmitter regulation through, e.g., glutamate uptake [5]. Additionally, astrocytes monitor and regulate the water flux and ion and pH homeostasis and provide structural integrity to the extracellular matrix in order to sustain an optimal environment [5]. As a key component and regulator of the selective permeable blood–brain barrier (BBB), astrocytes provide neurotrophic support, nutrients and waste removal, and together with microglia, they form the first line of defense against harmful agents [6]. In summary, astrocytes constantly cater to their surrounding neuronal population in order to maintain optimal function of the neural circuit.
Some subgroups of cortical astrocytes are shown to be located in individual three-dimensional domains, where they are believed to adapt their gene expression profile to neurons in their residing regions [2]. This notion is based on the observation that other astrocytes from different brain or spinal cord regions are unable to compensate in case of astrocyte domain loss in rodents [7,8]. Similarly, astrocytes are shown to display region-specific transcription factors, which contributes to their heterogeneity across brain regions [9,10,11]. Different anatomical regions are affected in various neurodegenerative diseases (AD: hippocampus and entorhinal (initially) and cortex (later), PD: substantia nigra, FTD: frontotemporal lobes, ALS: motor cortex, brain stem and spinal cord, Huntington disease (HD): striatum) and might therefore correlate with regional astrocytic failure to support the neurons locally, thereby progressing the diseases [12]. Additionally, some brain regions might also be more vulnerable to intrinsic pathological changes and external toxic insults in neurodegenerative diseases or naturally occurring changes such as during aging [13]. Aging is especially a common risk factor for developing neurodegenerative diseases and has, among other functions, been shown to affect the immune response in astrocytes [14,15,16]. Nonetheless, further research is required to clarify if astrocyte region-specific heterogeneity and dysregulation is a driving mechanism in neurodegeneration.
The functional complexity is mirrored in pathology. As a natural response to injury, infection or disease, astrocytes become reactive through morphological, transcriptional and functional transformations [17]. From primarily being tissue-embedded and non-motile, reactive astrocytes abandon certain neuronal supportive obligations in favor of an inflammatory activation through BBB remodeling, cytokine secretion and border formations [18,19]. These reactive astrocytes are believed to adapt based on the pathologic condition and can thus take on both neurosupportive and neurotoxic roles [3]. Such adaptations are highly complex and favor the concept of astrocytic “sub-states”, where specific pathologic conditions result in specific reactive astrocytes [20,21,22]. These astrocytic sub-states are likely modulated by the homeostatic need of their residing region and the nature of the pathological condition and might be dynamic during different disease states, thereby moving across a “reactive astrocyte spectrum” [21,22]. Importantly, astrocyte activation as an adaptive physiological response to insults to reestablish homeostasis and protect neuronal function should not be confused with disease-induced abnormal or chronic astrocyte reactivity observed in neurodegenerative diseases [20]. In the latter, the response often starts as being neuroprotective but eventually develops into toxicity, further contributing to cellular stress and neurodegeneration. Here, the underlying mechanisms are not well understood, and the widespread astrocyte heterogeneity only complicates the investigation [22,23]. However, despite the potential myriad of astrocyte subpopulations and sub-states, it is evident that multiple pathways are in fact common across a wider panel of neurodegeneration [24], which argues for a cross-disease investigative approach.

2. Neuroprotection versus Neurotoxicity

Neurodegenerative diseases affect more than 60 million people worldwide, and currently, no cure exists for any of the major groups such as AD, PD, FTD, ALS and HD. Despite being vastly different in genetic background, age of disease onset, clinical representation and phenotypic involvement, these neurodegenerative diseases have several hallmarks in common (Figure 1). The first and major one is their definition as proteinopathies. In each of these diseases, mutant proteins undergo pathological conformational changes, which results in accumulation and the formation of intracellular and/or extracellular aggregates (Figure 1) [12]. For AD, mutant amyloid-β (Aβ) aggregates into extracellular amyloid plaques and phosphorylated tau causes intracellular neurofibrillary tangles. For HD, mutated huntingtin forms intracellular aggregates, while α-synuclein accumulates in Lewi bodies in PD. In ALS, mutant TDP-43, FUS or SOD1 protein are shown to accumulate and aggregate, as for TDP-43 and/or tau in FTD, resulting in aggregation and neurofibrillary tangles, respectively. Many of these pathological intracellular protein buildups are found in the glia population as well [25,26,27,28,29,30,31]. This is likely an important contributing factor to the second common hallmark, which is neuroinflammation (Figure 1) [2,32].
Neuroinflammation can manifest in several ways. As mentioned previously, reactive astrocytes are shown to take on both neuroprotective and neurotoxic roles during the disease progression of neurodegenerative diseases [23,33]. Many aspects are still not well understood, but the general consensus is that the initial reactive response is considered neuroprotective but eventually and with continuous reactivity develops into a chronic and toxic inflammation, leading to degeneration of the neurons in the area. This development has detrimental consequences to the environment, including the neuronal network, which leads to the final common hallmark of neurodegenerative disease: neuronal toxicity and cell death (Figure 1). Often times, the astrocytic shift from neuroprotection to neurotoxicity happens gradually as the noxious stimuli accumulate, which ultimately escalates the disease progression [34]. As such, both predominantly neuroprotective and predominantly neurotoxic reactive astrocytes might reside in various regions of the brain at similar time points during the disease progression [13,23,33,35].
Examples in favor of a reactive astrocytic phenotype with neuroprotective characteristics can be observed in multiple neurodegenerative diseases. In AD, amyloid plaque deposition increased in an APPswe/PS1ΔE9 transgenic mouse model when astrocyte reactivity was reduced through glial fibrillary acid protein (GFAP)/Vimentin double knockout [36]. A similar acceleration of the disease progression was seen upon reactive astrocyte abolishment in prion disease [37]. In contrary, JAK2-STAT3 activation in reactive astrocytes in the Hdh140 HD transgenic mouse model was shown to induce proteostasis, thereby reducing the mutant huntingtin aggregation burden [38]. Additionally, increased clusterin (apolipoprotein J) secretion from astrocytes was able to rescue synaptic deficits and ameliorate Aβ neuropathology in the 5xFAD transgenic mouse model of AD [39]. Reactive astrocytes accumulate around amyloid plaques in human post-mortem tissue [13,40,41,42]; however, it is unknown if this is due to a reactive astrocytic attempt to constrain the toxicity or if amyloid plaques are driving reactivity in adjoining astrocytes [12]. In favor of neurotoxic reactivity, Liddelow and colleagues showed how some astrocytes encompassed deleterious properties, resulting in neurotoxicity after acute CNS injury in rodents [3]. This “A1” neurotoxic astrocyte sub-type was induced by microglial cytokine secretion and showed a specific transcriptional pattern while being unable to support synapse integrity. Interestingly, markers of A1 reactivity were also found in astrocytes in post-mortem samples from AD, PD, HD, ALS and multiple sclerosis (MS) patients, suggesting that neurotoxic astrocytes are present across a wide panel of neurodegenerative diseases [3,4].
Despite these cases of either predominant neuroprotection or predominant neurotoxicity, they very likely do not describe the full picture. For example, NFΚΒ pathway activation in astrocytes has been shown to mediate an initial neuroprotective astrocytic function with microglial anti-inflammatory activation, leucocyte infiltration and prolonging of disease onset during the pre-symptomatic phase of ALS [43]. However, the same continued astrocytic NFΚΒ activation later develops into a pro-inflammatory response, which accelerates the disease progression [43]. This argues for a dynamic transition between neuroprotection and neurotoxicity and, as a result, we must look at astrocytic contributions to neurodegenerative diseases with caution and avoid the temptation for simplistic categorization.
Many other cell types in the brain and spinal cord (neurons, microglia, endothelial cells, etc.) contribute to neuroinflammation and have shown to precede, initiate or drive astrocyte reactivity [3,44,45,46]. Similarly, astrocytes have been shown to govern microglial activation, and a large array of studies document intrinsic activation of monocultured astrocytes independent of their environment [43,47,48]. In the next part of this review, examples of shared astrocytic non-cell autonomous and cell autonomous mechanisms currently described across multiple neurodegenerative diseases will be provided.

3. Non-Cell Autonomous Mechanisms

Astrocytes are able to modulate their surroundings through secretion of many types of proteins. Through transsynaptic adhesion proteins or receptors modulations, astrocytic released molecules such as hevin (SPARCL1), thrombospondins (THBS1/2), glypicans (GPC4/6), transforming growth factor β1 (TGF-β1) and brain-derived neurotrophic factor (BDNF) are able to promote the formation and maturation of excitatory synapses [18,49]. In addition, astrocyte-secreted tumor necrosis factor-α (TNF-α) aids in increasing neuronal activity, as it promotes the presence of α-Amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptors on excitatory synapses and diminishes gamma-aminobutyric acid (GABA) receptors on inhibitory synapses [50]. Some secreted proteins like TGF-β1 have multiple functions, as it likewise promotes the formation of inhibitory synapses in the central nervous system [49].
One of the first known mechanisms of astrocyte non-cell autonomous effects in neurodegeneration is the apparent change in their secretome. Astrocytic releases of cytokines, chemokines and interleukins jointly increase the inflammatory response through astrocytic self-activation, microglial stimulation as well as periphery leukocyte involvement [6]. Simultaneously, chronic exposure to this secretory stimulation is causing several downstream pathologies, ultimately resulting in neurotoxicity and cell death.
In MS, astrocytes secrete chemokines such as CXCL10, CXCL12 and CCL20, which promote periphery immune cell infiltration [51,52,53], while cytokine secretion involving IL-6, TNF-α and IFN-γ is prominent in PD [29,54]. In AD, multiple complement effector proteins such as complement factor 3 (C3) are secreted directly or in exosomes in addition to IL-6, TNF-α and IL-1β, thereby driving a disease-stage-dependent inflammatory response involving microglial Aβ phagocytosis [55,56]. More specifically, Aβ is shown to drive an NFΚΒ-dependent and chronic astrocytic release of C3, which subsequently hampers microglia’s ability to phagocytose Aβ aggregates [57]. Additionally, astrocytes might also contribute to the amyloid burden in AD by increased secretion of Aβ [58], which could trigger additional astrocytic reactivity and self-activation [59]. Many of these inflammatory molecules are likely secreted by astrocytes in an attempt to evoke distinct downstream pathways thought to alleviate the pathology; however, chronic exposure evidently causes considerable dysfunction. As an example, we previously showed how FTD astrocyte-conditioned medium containing IL-6, IL-8 and IL-13 could inhibit axonal outgrowth of control neurons [60]. In ALS, large efforts have produced profound insight into the dysregulated astrocyte secretome. In both familial and sporadic ALS, multiple studies document a dysregulated astrocytic release of cytokines like TGF-β, TNF-α, IL-6 and IL-1β in addition to prostaglandins, inorganic polyphosphate, nitric oxide (NO), reactive oxygen species (ROS) and miRNA-containing extracellular vesicles, which cause abnormal microglial function and motor neuron toxicity [61,62,63,64,65,66,67,68,69,70,71,72,73,74,75]. This astrocyte secretome dysfunction in ALS has been shown to impair many key motor neuron functions involving autophagy [69,76,77], neurite outgrowth and length [62,69,73] and neuromuscular junction innervation [73,78,79], ultimately resulting in accelerated protein aggregation, excitotoxicity, cellular stress and motoric/respiratory degeneration and failure [48,79,80,81,82,83,84,85].

4. Cell Autonomous Mechanisms

A constant stream of evidence has firmly documented that astrocytes not only contribute to neurodegenerative disease pathology and progression through non-cell autonomous mechanisms but in fact also display multiple intrinsic pathological phenotypes. In AD, PD, ALS, FTD, MS as well as prion disease, astrocytes harbor increased expression of GFAP and C3, which are thought to be important (albeit not exclusive) markers of astrocyte reactivity encompassing a neurotoxic phenotype [3,37,60,78,86]. These protein expressions are part of a general observation of altered transcriptomic and proteomic levels, which largely result in a downregulation of astrocytic support functions such as ion and cholesterol homeostasis, glutamate uptake and synapse integrity and upregulation of pro-inflammatory pathways such as JAK-STAT3, NFAT and NFΚΒ with a resultant astrocytic toxic gain of functions [38,40,67,69,70,73,87,88,89,90,91,92,93,94]. In the following section, the cell autonomous mechanisms in astrocytes, which are common across multiple neurodegenerative diseases, are highlighted (see also Figure 2).

4.1. Communication

Unlike neurons, astrocytes are non-excitable and therefore have developed other means of communication. Besides cytokine, chemokine and interleukin releases mentioned above, astrocytes use calcium waves, transmitter release and gap junction couplings as a means of communication to maintain network homeostasis [49].

4.1.1. Calcium

Astrocytic calcium transients are highly dynamic. The “waves” are projected intra- and intercellularly through calcium-permeable ion channels and receptors, including various neurotransmitter receptors [5]. Upon neurotransmitter binding, a receptor-specific intracellular calcium signal is evoked, which allows astrocytes to distinguish between, e.g., glutamatergic and cholinergic synaptic activity [95,96]. Individual astrocytic subpopulations display unique calcium waves, which further contributes to their heterogeneity and complexity [97].
In multiple neurodegenerative diseases including AD [42,98,99,100,101,102], ALS [103,104], HD [105,106,107] and PD [29,108], calcium dyshomeostasis is observed. The dysregulated calcium response is often detected early in disease progression and might therefore contribute to the pathologically altered neuronal synapse activity [102,105,107,109,110,111]. For example, oligomeric forms of Aβ peptide were shown to drive astrocytic calcium hyperactivity early in AD disease progression, which consequently triggered glutamatergic hyperactivity in adjacent neurons [109,110]. Calcium is primarily stored in the endoplasmic reticulum (ER) within astrocytes but can also be imported into the cells through AMPA and N-methyl-D-aspartate (NMDA) receptors or through voltage-gated channels [49]. In ALS and PD, abnormal intracellular calcium dynamics with excess ER accumulation and storage release is observed, ultimately contributing to ER stress [29,103,104]. Both genetic PD mutations and drug-induced depletion of dopamine transmission mimicking PD have shown to increase astrocytic calcium excitability [29,108]. In the R6/2 transgenic mouse model of HD, spontaneous calcium signals and storage capacity are reduced, while evoked calcium responses are increased [105].

4.1.2. Intercellular Communication

In addition to receiving neurotransmitter information, calcium transients can release transmitters such as glutamate, GABA, D-serine and ATP from astrocytic processes, which contribute to the regulation of neuronal synapse excitability and plasticity [49,95]. In pathological conditions, the release of these transmitters from both neurons and astrocytes is dysregulated, and studies show that excess release causes a continuous self-activation, leading to cytotoxicity [42,99,112]. This phenomenon is observed in several models of PD and AD and is triggered by excessive astrocytic calcium-dependent release of glutamate and D-serine [110,111,113]. Similarly, astrocytic release of ATP can bind purinergic P2Y receptors on microglia, thereby modulating their phagocytic functions and release of cytokines further driving inflammation [114,115]. Astrocytic self-activation is also enhanced by the release of astrocytic TNF-α and prostaglandin, which, through calcium signaling, can promote additional “gliotransmitter” release [112].
The astrocyte–neuron communication can further be impaired through other means such as extracellular vesicle (EV) production. In LRKK2-mutant astrocytes, the EV biogenesis is altered, causing an accumulation of PD-related proteins within multivesicular bodies [116]. These astrocyte-secreted EVs are internalized by dopaminergic neurons and are linked to an astrocytic failure of providing neurotrophic support [116]. Furthermore, the previously mentioned C3-containing exosomes are released by AD astrocytes, consequently driving an inflammatory response and being linked to reduced neurite outgrowth [55,117].

4.1.3. Gap Junctions

Astrocytes not only communicate via neurotransmitters but also through gap junction couplings. Connexin-43 (Cx43) is the predominant connexin protein, which constitutes hemichannels and gap junctions in astrocytes [118,119,120]. Through these, diffusion of ions, metabolites, miRNAs and second messengers is facilitated, in addition to the important mitigation of calcium waves [121]. Therefore, gap junctions are key components in astrocyte networks, contributing to synapse activity modulation and homeostatic buffering. In ALS [69,122,123], AD [42,124] and PD [108], Cx43 is abnormally elevated, which causes increased gap junction coupling and hemichannel activity, resulting in calcium hyperactivity, neuronal excitability and cell death. Astrocytes also form collaborative glial networks with oligodendrocytes through their Cx43/Cx47 gap junction connections, which are important for their coordinated cross-talk [125]. In MS, oligodendrocytes and astrocytes lose their communication through decreased Cx43/Cx47 gap junction expression, consequently promoting demyelination and inflammation [126,127,128]. Lack of Cx43/Cx47 gap junction couplings is likewise found in AD [129].

4.2. Tripartite Synapse: Neurotransmitter Regulation and Synapse Function

Astrocytes are highly involved in neuronal synapse plasticity and function through their perisynaptic process ensheathment of neuronal synapses [130]. This tripartite synapse collaboration between pre- and postsynaptic neurons and astrocytes ensures optimal neuronal firing by continued astrocytic removal and recycle of excess neurotransmitters from the intersynaptic space [131,132]. In neurodegeneration, this function is impaired. Due to the downregulation of key astrocyte receptors, excitatory amino acid transporters 1 and 2 (EAAT1 and EAAT2), astrocytes fail to properly manage the uptake of the neurotransmitter glutamate. Glutamate is the main excitatory neurotransmitter in the brain, and a large part of its uptake from the synaptic cleft appears through EAAT1/2 receptors on astrocytes [133,134,135]. Persistent glutamate exposure is believed to cause excessive neuronal firing and abnormal neuronal calcium influx, which ultimately results in severe neuronal excitotoxicity [131,134,136,137]. Lack of EAAT1/2 receptors and consequent glutamate excitotoxicity is a common phenomenon in AD [138,139,140,141,142], PD [143,144], ALS [48,83,145,146] and HD [27,28,105]. Additionally, downregulation of astrocytic glutamate receptors drives abnormal microglial pruning and phagocytosis of hippocampal glutamatergic synapses in AD [147]. In HD, structural pathology, mediated by lack of astrocytic engagement of neuronal synapses, enables the hyper-excitability [148]. AD astrocytes likewise have perturbed glutamine metabolism and supply, which affects the GABA synthesis [149,150,151,152]. GABA is the main inhibitory neurotransmitter in the central nervous system, and lack thereof contributes to the excitatory imbalance [5]. In addition to neurotransmitter uptake, astrocytic release of various transmitters mentioned above is additionally important for synapse activity regulation. Abnormal gliotransmission has thus been shown to affect the synaptic transmission in PD and AD, consequently contributing to synapse loss and excitotoxicity [110,113,137,153,154,155]. Similarly, astrocytic release of C3 can bind to neuronal C3 receptors and hamper their synaptic density and dendritic morphology [156].
Astrocytes control the ion homeostasis through multiple ion channels, which is crucial for maintaining synapse functionality [157]. Potassium (K+) buffering is a key mechanism affected in neurodegeneration. Under physiological conditions, astrocytes regulate K+ levels in the extracellular space through clearance via K+ channels such as the main astrocytic Kir4.1 sub-type [49]. Through these mechanisms, astrocytes can modulate neuronal depolarization and thereby excitability [158]. In HD [26] and ALS [159], astrocytic Kir channels are downregulated, which hampers the K+ buffering and clearing, consequently causing increased extracellular K+ levels, overall leading to neuronal excitotoxicity. Besides through Kir.4.1 channels, K+ is released through EAAT2 and as mentioned previously, the astrocytic release and uptake of transmitters is modulated by calcium transients. This interconnected relationship between ion homeostasis, calcium dynamics and gliotransmission is therefore crucial for optimal neuronal function, and any dysregulation of one mechanism might consequently disrupt the others.

4.3. Mitochondrial Function

Mitochondrial dysfunction, ROS secretion and oxidative stress are common findings in neurodegeneration. In AD, mitochondrial dysfunctions are predicted as an early astrocytic phenotype consequently driving astrocyte reactivity [23], and in PD, maintenance of mitochondrial DNA is correlated with reduced mitochondrial respiration [29]. Astrocytic accumulations of α-synuclein in PD have been linked to damage within the mitochondrial structure, consequently lowering the total ATP level as well as causing disruption of the fission and fusion dynamics [160,161]. In ALS, mitochondria display decreased membrane potential and a compromised oxygen consumption, possibly correlated with a lower secretion of antioxidants [85,162]. Astrocytes express and release antioxidants as a part of their function in regulating the redox balance through removal of ROS in order to prevent oxidative damage of neurons [163]. Astrocytic nuclear factor erythroid 2-related factor 2 (Nrf2) transcription factor is a key regulator of antioxidant, detoxification and proteostasis pathways and might therefore be an important mediator in neurodegeneration [163]. In amyloid and tau pathology models of AD, pre-emptive activation of astrocytic Nrf2 was shown to be neuroprotective by attenuating the aggregation burden and slowing disease progression [33]. Similar observations were seen by overexpressing Nrf2 in astrocytes in SOD1-ALS mice [164] and in α-synuclein-mutant mice in PD [165], which advocates for a common therapeutic mechanism in neurodegeneration. Interestingly, Nrf2 target gene levels are in fact increased in AD but might appear too limited or too late in the disease progression to make a considerable neuroprotective difference [33,166].

4.4. Energy Metabolism

Neurons rely on astrocytes for nutritional support to meet their optimal energy consumption. Through key mechanisms such as the lactate shuttle, astrocytes convert glucose or glycogen to lactate via aerobic glycolysis and provide the lactate to adjacent neurons via monocarboxylate transporters [167]. In the neurons, the lactate is incorporated into the oxidative cycle for ATP production [167]. Lactate is also shuttled directly to neurons from the bloodstream via the astrocytes’ perisynaptic process engagement with the vasculature [168]. In ALS, astrocytic intra- and extracellular levels of lactate are decreased, possibly due to diminished lactate production and/or transport [169,170]. Similarly, AD astrocytes have decreased glycolysis and secretion of lactate to the environment [101]. Lack of lactate production causes metabolic dysfunction and failure of adequate neuronal energy supply. Additionally, lactate production is coupled to glutamate transporter activity, and their shared impairment might therefore further hamper the energy metabolism in neurodegenerative disorders [12].
A widespread dysregulation of metabolites is generally observed in neurodegenerative diseases. Astrocyte metabolic proteomics is increasingly enriched during the commencement and disease progression of AD, with markers expressing both neuroprotective and neurotoxic phenotypes [24]. More specifically, AD astrocytes display augmented glycolytic flux and reduced glycogen storage [142]. In PD, astrocytes have altered polyamine and phospholipid levels [29], and in ALS, astrocytic metabolic dysfunction with compromised adenosine, fructose and glycogen metabolism is observed [171,172].
Finally, astrocytes are the main synthesizers and suppliers of apolipoprotein E (ApoE), which ensures sufficient transport of cholesterol to neurons [49]. Cholesterol is an essential lipid in cell membranes and presynaptic vesicle formation and is therefore crucial for the integrity and function of synapses. One of the main risk factors of late-onset AD is the human isoform APOEε4 variant [173,174,175,176]. Carriers are shown to have pathologically altered glucose metabolism, lactate production, cholesterol homeostasis and calcium dynamics, resulting in AD [98,177,178,179,180]. Dysfunctional cholesterol and lipid metabolism are also found in HD [35,181,182] and PD [183,184] and appear to be a general mechanism of reactive astrocytes, potentially influencing many neurodegenerative diseases [185].

4.5. Clearance of Protein Aggregates

Astrocytes have been shown to contribute to the insufficient clearance of protein aggregates in neurodegeneration due to failure of the glymphatic system. Under physiological conditions, astrocytes regulate water flux as well as removal of metabolic waste products from the brain interstitium to and from the perivascular space through their Aquaporin-4 (AQP4) water channels [5]. Additionally, through the synergistic collaboration between AQP4 and Kir4.1 channels, osmohomeostasis is maintained in the brain [186,187,188]. During disease progressions, astrocytes remove Aβ peptides and mutant huntingtin through the glymphatic system [189,190]. However, due to the downregulations of Kir4.1 and AQP4, this clearing mechanism is impaired, consequently accelerating the aggregation burden [26,191,192]. Additionally, astrocytic ApoE is normally involved in Aβ clearance, but in APOEε4 carriers, this function appears to be compromised, which likewise contributes to the enhanced Aβ aggregation [179,193].
Secondly, astrocytes contribute to clearance of proteins through their intracellular lysosomal pathways. This mechanism is affected in PD, where both α-synuclein oligomers as well as mutations in the LRRK2 gene interfere with the astrocytic clearance of the aggregation burden [160,194]. Neuronal α-synuclein is taken up by astrocytes from the extracellular space via endocytosis as well as transmitted directly from neurons [93,195]. Similarly, aggregates are transferred between astrocytes via nanotubules [161]. This increased inclusion burden aided by overload/stress-induced insufficient lysosomal degradation further triggers cellular toxicity and reactivity [93,161]. More specifically, α-synuclein is shown to alter the lysosomal morphology, distribution and function by alkalinization and decrease activity of lysosomal proteases in neuronal cells and idiopathic PD brains [196,197], which correlates with similar observations of disrupted lysosomal proteolysis recently observed in astrocytes in an early-onset PD model [198]. Astrocytic engulfment of monomeric tau from the extracellular environment has been observed in tauopathies [199], and the endo-lysosomal dysfunction in astrocytes is likewise predicted as an early disease phenotype in AD [23,200].
Finally, microglial and astrocytic cross-communication is important for the optimal clearance of Aβ and α-synuclein aggregates [201]. Intercellular miscommunication or toxic reactivity might therefore hamper this essential function, thereby accelerating disease progressions.

4.6. Autophagy

Autophagy is a vital cellular process, which facilitates the degradation and recycling of intracellular components such as damaged organelles and protein aggregates [202]. Several conditions like starvation, stress and pharmacological treatment can modulate autophagy to facilitate a faster release of important nutrients in the reestablishment of cellular homeostasis [203]. Autophagy is a multi-step process, where formation of large double-membrane vesicles termed autophagosomes bind and enclose cargo destined for degradation through ubiquitin-P62-LC3 protein interactions [202]. The autophagosomes fuse with late endosomes and finally with lysosomes, and the lysosomal enzymes degrade the vesicular content for reuse [204].
Mutations in LRRK2 are a common cause of genetic PD [205]. LRRK2 contributes, among many functions, to the phosphorylation of various proteins from the Rab family as well as P62, thereby filling a prominent role in autophagy initiation and vesicle transport [206]. Consequently, LRRK2-mutant astrocytes display impaired autophagy, possibly linked to progressive accumulation of α-synuclein as mentioned above [207]. Additionally, α-synuclein oligomers have been shown to interfere with the autophagosome–lysosome fusion, consequently halting the autophagic flux in astrocytes [161]. This degradation impairment results in insufficient turnover of damaged mitochondria [161]. Moreover, in ALS/FTD-relevant C9ORF72-mutant models, accumulation of P62 is present in astrocytes [208], and we previously showed how human induced pluripotent stem cell-derived FTD astrocytes displayed insufficient autophagy, consequently perturbing the mitochondrial turnover [60]. This lack of mitophagy resulted in augmented mitochondrial fusion, impaired mitochondrial respiration and glycolysis, increased ROS and stress granules formation, which overall triggered astrocyte reactivity and cytokine secretion [60].

4.7. Blood–Brain Barrier

Through close interaction with capillary endothelial cells and pericytes, astrocytes function as efficient gatekeepers of the central nervous system through regulation of the integrity and permeability of the BBB [6]. Additionally, neurons rely on astrocytes for shuttling of nutrients and metabolic substrates from the blood stream through the BBB [6], and with calcium-dependent release of vasodilators or vasoconstrictors, astrocytes control the blood flow depending on the neuronal energy demand [209]. As mentioned previously, AQP4 levels are changed in various neurodegenerative diseases, which also contribute to the compromised BBB [13,210]. Therefore, cerebrovascular deterioration in AD could be correlated with astrocytic calcium hyperactivity [42].
In ALS, astrocytes contribute to BBB disruption and pericyte loss, which allow periphery leukocyte infiltration and a disturbance of the homeostatic environment [43,211,212,213,214]. In AD, APOEε4 carriers display pericyte degeneration due to insufficient astrocyte suppression of proinflammatory pathways, consequently damaging the BBB [215,216]. Additionally, the barrier function of BBB tight junctions is impaired when astrocytes carry the APOEε4 isoform [217], and astrocytes modulate leukocyte infiltration due to Aβ exposure [218]. In HD, astrocytes trigger endothelial cell proliferation and pericyte damage, which disrupts the vascular function [219]. Finally, in PD, reactive astrocytes fail to support vessel formation and barrier integrity [220].

4.8. Glial Border Formation

Astrocytes are crucial for the integrity of the stroma in the central nervous system and provide important structural support [5]. Through their secretion of molecules such as proteoglycans, astrocytes contribute to the extracellular matrix (ECM) structure surrounding synapses and within the synaptic cleft. This is important for capturing nutrients and growth factors as well as acting as a diffusion barrier for neurotransmitter concentration buffering [49]. During neurodegeneration, subpopulations of astrocytes initiate a border formation, which primarily functions to contain the injury. However, emerging evidence point towards a rising role of abnormal scarring in these disorders [221]. Here, ECM molecules promote a pro-inflammatory signal mediated by microglia and infiltrating macrophages, which further drives the astrocytic reactivity and matrix release [221]. Conversely, glial border formation is shown to be beneficial in its formation of “glial bridges”, along which axonal regrowth is permitted if appropriately guided by growth factors [222]. Lack of astrocytic neuro-supportive functions likely influences this mechanism in neurodegenerative diseases [34].
Chondroitin sulfate proteoglycans (CSPGs) are central components of glial border formation and mediated by reactive astrocytes in MS lesions [223]. Although this physical barrier aids in the containment of the lesion, it also prevents axonal outgrowth and remyelination [223]. As a consequence, excessive glial border formation by astrocytes might distort the tissue architecture, thereby prolonging the disease course of MS. Similarly, in ALS, CSPG accumulation is present at the site of motor neuron degeneration, and excessive and abnormal CSPG receptors are also found on the surface of reactive astrocytes, further contributing to their self-activation and the overall impairment of the homeostatic environment [224,225]. Additionally, chronic release of TGF-β promotes excessive fibrosis in ALS [226] and acts as an upstream regulator of CSPG secretion [227,228]. In AD, astrocytic proteoglycans interact with Aβ plaques, consequently promoting aggregation and inhibiting clearance [229].
Finally, the STAT3 pathway in astrocytes is an important modulator of glial border formation after spinal cord injury [230]. Abnormal expression of the STAT3 pathway in multiple neurodegenerative disease could likely contribute to border formation imbalances. Similarly, the damage to the BBB generally observed in neurodegeneration permits the infiltration of mesenchymal cells and blood proteins, which mediates excessive scarring [221].

5. Conclusions

Astrocyte physiology and pathology are complex. It is fascinating to speculate how many sub-types and sub-states exist and how astrocytic adaptability might drive neurodegenerative disease progressions. Despite their profound heterogeneity, astrocytes as a cell group share common pathological characteristics across a broad spectrum of neurodegeneration. Here, we have highlighted the disease commonalities in some of the most common mechanisms, but many more are likely present, and further research is undoubtedly warranted. What we learn from one disease could potentially be extrapolated to others, ultimately benefitting more patients. Current treatment regimens largely follow the dominant field of neurocentric studies, but this neuronal favoring has unfortunately resulted in many unsuccessful clinical trials. Given the paramount role of astrocytes in ensuring optimal neuronal function and survival, these glial cells constitute important therapeutic targets in drug development. With this review, we encourage future studies in astrocyte mechanisms with a cross-disease and cross-model investigative approach in an effort to lower the variability and potentially disclose new and shared therapeutic targets for multiple neurodegenerative diseases.

Author Contributions

K.S.D. planned and wrote the manuscript. K.F. edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Lundbeck Foundation, grant number R434-2023-242, Innovation Fund Denmark, grant number 2078-00002B (StemScreen), Novo Nordisk Foundation, grant number NNF21OC0071571 (RhoAD) and Alzheimer Foundation Denmark.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Lenhossék, M. Der Feinere Bau Des Nervensystems Im Lichte Neuester Forschung, 2nd ed.; Fischer’s Medicinische Buchhandlung H. Kornfield: Berlin, Germany, 1895. [Google Scholar]
  2. Patani, R.; Hardingham, G.E.; Liddelow, S.A. Functional Roles of Reactive Astrocytes in Neuroinflammation and Neurodegeneration. Nat. Rev. Neurol. 2023, 19, 395–409. [Google Scholar] [CrossRef]
  3. Liddelow, S.A.; Guttenplan, K.A.; Clarke, L.E.; Bennett, F.C.; Bohlen, C.J.; Schirmer, L.; Bennett, M.L.; Münch, A.E.; Chung, W.S.; Peterson, T.C.; et al. Neurotoxic Reactive Astrocytes Are Induced by Activated Microglia. Nature 2017, 541, 481–487. [Google Scholar] [CrossRef]
  4. Zamanian, J.L.; Xu, L.; Foo, L.C.; Nouri, N.; Zhou, L.; Giffard, R.G.; Barres, B.A. Genomic Analysis of Reactive Astrogliosis. J. Neurosci. 2012, 32, 6391–6410. [Google Scholar] [CrossRef] [PubMed]
  5. Verkhratsky, A.; Nedergaard, M. Physiology of astroglia. Physiol. Rev. 2018, 98, 239–389. [Google Scholar] [CrossRef] [PubMed]
  6. Ahmad, A.; Patel, V.; Xiao, J.; Moshahid Khan, M. The Role of Neurovascular System in Neurodegenerative Diseases. Mol. Neurobiol. 2020, 57, 4373–4393. [Google Scholar] [CrossRef]
  7. Tsai, H.-H.; Li, H.; Fuentealba, L.C.; Molofsky, A.V.; Taveira-Marques, R.; Zhuang, H.; Tenney, A.; Murnen, A.T.; Fancy, S.P.J.; Merkle, F.; et al. Regional Astrocyte Allocation Regulates CNS Synaptogenesis and Repair. Science 2012, 337, 358–362. [Google Scholar] [CrossRef]
  8. Wilhelmsson, U.; Bushong, E.A.; Price, D.L.; Smarr, B.L.; Phung, V.; Terada, M.; Ellisman, M.H.; Pekny, M. Redefining the Concept of Reactive Astrocytes as Cells That Remain within Their Unique Domains upon Reaction to Injury. Proc. Natl. Acad. Sci. USA 2006, 103, 17513–17518. [Google Scholar] [CrossRef]
  9. Bayraktar, O.A.; Bartels, T.; Holmqvist, S.; Kleshchevnikov, V.; Martirosyan, A.; Polioudakis, D.; Ben Haim, L.; Young, A.M.H.; Batiuk, M.Y.; Prakash, K.; et al. Astrocyte Layers in the Mammalian Cerebral Cortex Revealed by a Single-Cell in Situ Transcriptomic Map. Nat. Neurosci. 2020, 23, 500–509. [Google Scholar] [CrossRef] [PubMed]
  10. Batiuk, M.Y.; Martirosyan, A.; Wahis, J.; de Vin, F.; Marneffe, C.; Kusserow, C.; Koeppen, J.; Viana, J.F.; Oliveira, J.F.; Voet, T.; et al. Identification of Region-Specific Astrocyte Subtypes at Single Cell Resolution. Nat. Commun. 2020, 11, 1220. [Google Scholar] [CrossRef]
  11. Huang, A.Y.S.; Woo, J.; Sardar, D.; Lozzi, B.; Bosquez Huerta, N.A.; Lin, C.C.J.; Felice, D.; Jain, A.; Paulucci-Holthauzen, A.; Deneen, B. Region-Specific Transcriptional Control of Astrocyte Function Oversees Local Circuit Activities. Neuron 2020, 106, 992–1008.e9. [Google Scholar] [CrossRef]
  12. Brandebura, A.N.; Paumier, A.; Onur, T.S.; Allen, N.J. Astrocyte Contribution to Dysfunction, Risk and Progression in Neurodegenerative Disorders. Nat. Rev. Neurosci. 2023, 24, 23–39. [Google Scholar] [CrossRef] [PubMed]
  13. Habib, N.; McCabe, C.; Medina, S.; Varshavsky, M.; Kitsberg, D.; Dvir-Szternfeld, R.; Green, G.; Dionne, D.; Nguyen, L.; Marshall, J.L.; et al. Disease-Associated Astrocytes in Alzheimer’s Disease and Aging. Nat. Neurosci. 2020, 23, 701–706. [Google Scholar] [CrossRef] [PubMed]
  14. Boisvert, M.M.; Erikson, G.A.; Shokhirev, M.N.; Allen, N.J. The Aging Astrocyte Transcriptome from Multiple Regions of the Mouse Brain. Cell Rep. 2018, 22, 269–285. [Google Scholar] [CrossRef]
  15. Clarke, L.E.; Liddelow, S.A.; Chakraborty, C.; Münch, A.E.; Heiman, M.; Barres, B.A. Normal Aging Induces A1-like Astrocyte Reactivity. Proc. Natl. Acad. Sci. USA 2018, 115, E1896–E1905. [Google Scholar] [CrossRef] [PubMed]
  16. Orre, M.; Kamphuis, W.; Osborn, L.M.; Melief, J.; Kooijman, L.; Huitinga, I.; Klooster, J.; Bossers, K.; Hol, E.M. Acute Isolation and Transcriptome Characterization of Cortical Astrocytes and Microglia from Young and Aged Mice. Neurobiol. Aging 2014, 35, 1–14. [Google Scholar] [CrossRef]
  17. Escartin, C.; Galea, E.; Lakatos, A.; O’Callaghan, J.P.; Petzold, G.C.; Serrano-Pozo, A.; Steinhäuser, C.; Volterra, A.; Carmignoto, G.; Agarwal, A.; et al. Reactive Astrocyte Nomenclature, Definitions, and Future Directions. Nat. Neurosci. 2021, 24, 312–325. [Google Scholar] [CrossRef]
  18. Vainchtein, I.D.; Molofsky, A.V. Astrocytes and Microglia: In Sickness and in Health. Trends Neurosci. 2020, 43, 144–154. [Google Scholar] [CrossRef]
  19. Sica, R.E.; Caccuri, R.; Quarracino, C.; Capani, F. Are Astrocytes Executive Cells within the Central Nervous System? Arq. Neuropsiquiatr. 2016, 74, 671–678. [Google Scholar] [CrossRef]
  20. Sofroniew, M.V. Astrocyte Reactivity: Subtypes, States, and Functions in CNS Innate Immunity. Trends Immunol. 2020, 41, 758–770. [Google Scholar] [CrossRef]
  21. Sadick, J.S.; O’Dea, M.R.; Hasel, P.; Dykstra, T.; Faustin, A.; Liddelow, S.A. Astrocytes and Oligodendrocytes Undergo Subtype-Specific Transcriptional Changes in Alzheimer’s Disease. Neuron 2022, 110, 1788–1805. [Google Scholar] [CrossRef]
  22. Hasel, P.; Rose, I.V.L.; Sadick, J.S.; Kim, R.D.; Liddelow, S.A. Neuroinflammatory Astrocyte Subtypes in the Mouse Brain. Nat. Neurosci. 2021, 24, 1475–1487. [Google Scholar] [CrossRef]
  23. Galea, E.; Weinstock, L.D.; Larramona-Arcas, R.; Pybus, A.F.; Giménez-Llort, L.; Escartin, C.; Wood, L.B. Multi-Transcriptomic Analysis Points to Early Organelle Dysfunction in Human Astrocytes in Alzheimer’s Disease. Neurobiol. Dis. 2022, 166, 105655. [Google Scholar] [CrossRef] [PubMed]
  24. Johnson, E.C.B.; Dammer, E.B.; Duong, D.M.; Ping, L.; Zhou, M.; Yin, L.; Higginbotham, L.A.; Guajardo, A.; White, B.; Troncoso, J.C.; et al. Large-Scale Proteomic Analysis of Alzheimer’s Disease Brain and Cerebrospinal Fluid Reveals Early Changes in Energy Metabolism Associated with Microglia and Astrocyte Activation. Nat. Med. 2020, 26, 769–780. [Google Scholar] [CrossRef] [PubMed]
  25. Tziortzouda, P.; Van Den Bosch, L.; Hirth, F. Triad of TDP43 Control in Neurodegeneration: Autoregulation, Localization and Aggregation. Nat. Rev. Neurosci. 2021, 22, 197–208. [Google Scholar] [CrossRef] [PubMed]
  26. Tong, X.; Ao, Y.; Faas, G.C.; Nwaobi, S.E.; Xu, J.; Haustein, M.D.; Anderson, M.A.; Mody, I.; Olsen, M.L.; Sofroniew, M.V.; et al. Astrocyte Kir4.1 Ion Channel Deficits Contribute to Neuronal Dysfunction in Huntington’s Disease Model Mice. Nat. Neurosci. 2014, 17, 694–703. [Google Scholar] [CrossRef] [PubMed]
  27. Shin, J.Y.; Fang, Z.H.; Yu, Z.X.; Wang, C.E.; Li, S.H.; Li, X.J. Expression of Mutant Huntingtin in Glial Cells Contributes to Neuronal Excitotoxicity. J. Cell Biol. 2005, 171, 1001–1012. [Google Scholar] [CrossRef] [PubMed]
  28. Faideau, M.; Kim, J.; Cormier, K.; Gilmore, R.; Welch, M.; Auregan, G.; Dufour, N.; Guillermier, M.; Brouillet, E.; Hantraye, P.; et al. In Vivo Expression of Polyglutamine-Expanded Huntingtin by Mouse Striatal Astrocytes Impairs Glutamate Transport: A Correlation with Huntington’s Disease Subjects. Hum. Mol. Genet. 2010, 19, 3053–3067. [Google Scholar] [CrossRef] [PubMed]
  29. Sonninen, T.M.; Hämäläinen, R.H.; Koskuvi, M.; Oksanen, M.; Shakirzyanova, A.; Wojciechowski, S.; Puttonen, K.; Naumenko, N.; Goldsteins, G.; Laham-Karam, N.; et al. Metabolic Alterations in Parkinson’s Disease Astrocytes. Sci. Rep. 2020, 10, 14474. [Google Scholar] [CrossRef]
  30. Wakabayashi, K.; Hayashi, S.; Yoshimoto, M.; Kudo, H.; Takahashi, H. NACP/a-Synuclein-Positive Filamentous Inclusions in Astrocytes and Oligodendrocytes of Parkinson’s Disease Brains. Acta Neuropathol. 2000, 99, 14–20. [Google Scholar] [CrossRef]
  31. Braak, H.; Sastre, M.; Del Tredici, K. Development of α-Synuclein Immunoreactive Astrocytes in the Forebrain Parallels Stages of Intraneuronal Pathology in Sporadic Parkinson’s Disease. Acta Neuropathol. 2007, 114, 231–241. [Google Scholar] [CrossRef]
  32. Stoklund Dittlau, K.; Van Den Bosch, L. Why Should We Care about Astrocytes in a Motor Neuron Disease? Front. Mol. Med. 2023, 3, 1047540. [Google Scholar] [CrossRef]
  33. Jiwaji, Z.; Tiwari, S.S.; Avilés-Reyes, R.X.; Hooley, M.; Hampton, D.; Torvell, M.; Johnson, D.A.; McQueen, J.; Baxter, P.; Sabari-Sankar, K.; et al. Reactive Astrocytes Acquire Neuroprotective as Well as Deleterious Signatures in Response to Tau and Aß Pathology. Nat. Commun. 2022, 13, 135. [Google Scholar] [CrossRef]
  34. Verkhratsky, A.; Butt, A.; Li, B.; Illes, P.; Zorec, R.; Semyanov, A.; Tang, Y.; Sofroniew, M.V. Astrocytes in Human Central Nervous System Diseases: A Frontier for New Therapies. Signal Transduct. Target Ther. 2023, 8, 396. [Google Scholar] [CrossRef]
  35. Al-Dalahmah, O.; Sosunov, A.A.; Shaik, A.; Ofori, K.; Liu, Y.; Vonsattel, J.P.; Adorjan, I.; Menon, V.; Goldman, J.E. Single-Nucleus RNA-Seq Identifies Huntington Disease Astrocyte States. Acta Neuropathol. Commun. 2020, 8, 19. [Google Scholar] [CrossRef]
  36. Kraft, A.W.; Hu, X.; Yoon, H.; Yan, P.; Xiao, Q.; Wang, Y.; Gil, S.C.; Brown, J.; Wilhelmsson, U.; Restivo, J.L.; et al. Attenuating Astrocyte Activation Accelerates Plaque Pathogenesis in APP/PS1 Mice. FASEB J. 2013, 27, 187–198. [Google Scholar] [CrossRef]
  37. Hartmann, K.; Sepulveda-Falla, D.; Rose, I.V.L.; Madore, C.; Muth, C.; Matschke, J.; Butovsky, O.; Liddelow, S.; Glatzel, M.; Krasemann, S. Complement 3+-Astrocytes Are Highly Abundant in Prion Diseases, but Their Abolishment Led to an Accelerated Disease Course and Early Dysregulation of Microglia. Acta Neuropathol. Commun. 2020, 7, 83. [Google Scholar] [CrossRef]
  38. Abjean, L.; Haim, L.B.; Riquelme-Perez, M.; Gipchtein, P.; Derbois, C.; Palomares, M.A.; Petit, F.; Hérard, A.S.; Gaillard, M.C.; Guillermier, M.; et al. Reactive Astrocytes Promote Proteostasis in Huntington’s Disease through the JAK2-STAT3 Pathway. Brain 2023, 146, 149–166. [Google Scholar] [CrossRef] [PubMed]
  39. Chen, F.; Swartzlander, D.B.; Ghosh, A.; Fryer, J.D.; Wang, B.; Zheng, H. Clusterin Secreted from Astrocyte Promotes Excitatory Synaptic Transmission and Ameliorates Alzheimer’s Disease Neuropathology. Mol. Neurodegener. 2021, 16, 5. [Google Scholar] [CrossRef] [PubMed]
  40. Leng, K.; Rose, I.V.L.; Kim, H.; Xia, W.; Romero-Fernandez, W.; Rooney, B.; Koontz, M.; Li, E.; Ao, Y.; Wang, S.; et al. CRISPRi Screens in Human IPSC-Derived Astrocytes Elucidate Regulators of Distinct Inflammatory Reactive States. Nat. Neurosci. 2022, 25, 1528–1542. [Google Scholar] [CrossRef] [PubMed]
  41. Beach, T.G.; Mcgeer, E.G. Lamina-Specific Arrangement of Astrocytic Gliosis and Senile Plaques in Alzheimer’s Disease Visual Cortex. Brain Res. 1988, 463, 357–361. [Google Scholar] [CrossRef] [PubMed]
  42. Delekate, A.; Füchtemeier, M.; Schumacher, T.; Ulbrich, C.; Foddis, M.; Petzold, G.C. Metabotropic P2Y1 Receptor Signalling Mediates Astrocytic Hyperactivity in Vivo in an Alzheimer’s Disease Mouse Model. Nat. Commun. 2014, 5, 5422. [Google Scholar] [CrossRef]
  43. Ouali Alami, N.; Schurr, C.; Olde Heuvel, F.; Tang, L.; Li, Q.; Tasdogan, A.; Kimbara, A.; Nettekoven, M.; Ottaviani, G.; Raposo, C.; et al. NF-κB Activation in Astrocytes Drives a Stage-specific Beneficial Neuroimmunological Response in ALS. EMBO J. 2018, 37, e98697. [Google Scholar] [CrossRef]
  44. Gerber, Y.N.; Sabourin, J.C.; Rabano, M.; Vivanco, M.; Perrin, F.E. Early Functional Deficit and Microglial Disturbances in a Mouse Model of Amyotrophic Lateral Sclerosis. PLoS ONE 2012, 7, e360002012. [Google Scholar] [CrossRef]
  45. Joshi, A.U.; Minhas, P.S.; Liddelow, S.A.; Haileselassie, B.; Andreasson, K.I.; Dorn, G.W.; Mochly-Rosen, D. Fragmented Mitochondria Released from Microglia Trigger A1 Astrocytic Response and Propagate Inflammatory Neurodegeneration. Nat. Neurosci. 2019, 22, 1635–1648. [Google Scholar] [CrossRef]
  46. Taylor, X.; Cisternas, P.; Jury, N.; Martinez, P.; Huang, X.; You, Y.; Redding-Ochoa, J.; Vidal, R.; Zhang, J.; Troncoso, J.; et al. Activated Endothelial Cells Induce a Distinct Type of Astrocytic Reactivity. Commun. Biol. 2022, 5, 282. [Google Scholar] [CrossRef] [PubMed]
  47. Ajmone-Cat, M.A.; Onori, A.; Toselli, C.; Stronati, E.; Morlando, M.; Bozzoni, I.; Monni, E.; Kokaia, Z.; Lupo, G.; Minghetti, L.; et al. Increased FUS Levels in Astrocytes Leads to Astrocyte and Microglia Activation and Neuronal Death. Sci. Rep. 2019, 9, 4572. [Google Scholar] [CrossRef] [PubMed]
  48. Papadeas, S.T.; Kraig, S.E.; O’Banion, C.; Lepore, A.C.; Maragakis, N.J. Astrocytes Carrying the Superoxide Dismutase 1 (SOD1 G93A) Mutation Induce Wild-Type Motor Neuron Degeneration in Vivo. Proc. Natl. Acad. Sci. USA 2011, 108, 17803–17808. [Google Scholar] [CrossRef]
  49. Liu, X.; Ying, J.; Wang, X.; Zheng, Q.; Zhao, T.; Yoon, S.; Yu, W.; Yang, D.; Fang, Y.; Hua, F. Astrocytes in Neural Circuits: Key Factors in Synaptic Regulation and Potential Targets for Neurodevelopmental Disorders. Front. Mol. Neurosci. 2021, 14, 729273. [Google Scholar] [CrossRef]
  50. Allen, N.J.; Eroglu, C. Cell Biology of Astrocyte-Synapse Interactions. Neuron 2017, 96, 697–708. [Google Scholar] [CrossRef] [PubMed]
  51. Ambrosini, E.; Remoli, M.E.; Giacomini, E.; Rosicarelli, B.; Serafini, B.; Lande, R.; Aloisi, F.; Coccia, E.M. Astrocytes Produce Dendritic Cell-Attracting Chemokines In Vitro and in Multiple Sclerosis Lesions. J. Neuropathol. Exp. Neurol. 2005, 64, 706–715. [Google Scholar] [CrossRef]
  52. Mills Ko, E.; Ma, J.H.; Guo, F.; Miers, L.; Lee, E.; Bannerman, P.; Burns, T.; Ko, D.; Sohn, J.; Soulika, A.M.; et al. Deletion of Astroglial CXCL10 Delays Clinical Onset but Does Not Affect Progressive Axon Loss in a Murine Autoimmune Multiple Sclerosis Model. J. Neuroinflammation 2014, 11, 105. [Google Scholar] [CrossRef]
  53. Krumbholz, M.; Theil, D.; Cepok, S.; Hemmer, B.; Kivisäkk, P.; Ransohoff, R.M.; Hofbauer, M.; Farina, C.; Derfuss, T.; Hartle, C.; et al. Chemokines in Multiple Sclerosis: CXCL12 and CXCL13 up-Regulation Is Differentially Linked to CNS Immune Cell Recruitment. Brain 2006, 129, 200–211. [Google Scholar] [CrossRef]
  54. Barcia, C.; Ros, C.M.; Annese, V.; Gómez, A.; Ros-Bernal, F.; Aguado-Year, C.; Martínez-Paǵn, M.E.; De Pablos, V.; Fernandez-Villalba, E.; Herrero, M.T. IFN-γ Signaling, with the Synergistic Contribution of TNF-α, Mediates Cell Specific Microglial and Astroglial Activation in Experimental Models of Parkinson’s Disease. Cell Death Dis. 2011, 2, e1422011. [Google Scholar] [CrossRef] [PubMed]
  55. Goetzl, E.J.; Schwartz, J.B.; Abner, E.L.; Jicha, G.A.; Kapogiannis, D. High Complement Levels in Astrocyte-Derived Exosomes of Alzheimer Disease. Ann. Neurol. 2018, 83, 544–552. [Google Scholar] [CrossRef] [PubMed]
  56. Fu, H.; Liu, B.; Frost, J.L.; Hong, S.; Jin, M.; Ostaszewski, B.; Shankar, G.M.; Costantino, I.M.; Carroll, M.C.; Mayadas, T.N.; et al. Complement Component C3 and Complement Receptor Type 3 Contribute to the Phagocytosis and Clearance of Fibrillar Aβ by Microglia. Glia 2012, 60, 993–1003. [Google Scholar] [CrossRef]
  57. Lian, H.; Litvinchuk, A.; Chiang, A.C.-A.; Aithmitti, N.; Jankowsky, J.L.; Zheng, H. Astrocyte-Microglia Cross Talk through Complement Activation Modulates Amyloid Pathology in Mouse Models of Alzheimer’s Disease. J. Neurosci. 2016, 36, 577–589. [Google Scholar] [CrossRef] [PubMed]
  58. Liao, M.C.; Muratore, C.R.; Gierahn, T.M.; Sullivan, S.E.; Srikanth, P.; De Jager, P.L.; Love, J.C.; Young-Pearse, T.L. Single-Cell Detection of Secreted Aβ and SAPPα from Human IPSC-Derived Neurons and Astrocytes. J. Neurosci. 2016, 36, 1730–1746. [Google Scholar] [CrossRef]
  59. LaRocca, T.J.; Cavalier, A.N.; Roberts, C.M.; Lemieux, M.R.; Ramesh, P.; Garcia, M.A.; Link, C.D. Amyloid Beta Acts Synergistically as a Pro-Inflammatory Cytokine. Neurobiol. Dis. 2021, 159, 105493. [Google Scholar] [CrossRef] [PubMed]
  60. Chandrasekaran, A.; Stoklund Dittlau, K.; Corsi, G.I.; Haukedal, H.; Doncheva, N.T.; Ramakrishna, S.; Ambardar, S.; Salcedo, C.; Schmidt, S.I.; Zhang, Y.; et al. Astrocytic Reactivity Triggered by Defective Autophagy and Metabolic Failure Causes Neurotoxicity in Frontotemporal Dementia Type 3. Stem Cell Rep. 2021, 16, 1–16. [Google Scholar] [CrossRef] [PubMed]
  61. Rojas, F.; Cortes, N.; Abarzua, S.; Dyrda, A.; van Zundert, B. Astrocytes Expressing Mutant SOD1 and TDP43 Trigger Motoneuron Death That Is Mediated via Sodium Channels and Nitroxidative Stress. Front. Cell Neurosci. 2014, 8, 24. [Google Scholar] [CrossRef]
  62. Varcianna, A.; Myszczynska, M.A.; Castelli, L.M.; O’Neill, B.; Kim, Y.; Talbot, J.; Nyberg, S.; Nyamali, I.; Heath, P.R.; Stopford, M.J.; et al. Micro-RNAs Secreted through Astrocyte-Derived Extracellular Vesicles Cause Neuronal Network Degeneration in C9orf72 ALS. EBioMedicine 2019, 40, 626–635. [Google Scholar] [CrossRef]
  63. Arredondo, C.; Cefaliello, C.; Dyrda, A.; Jury, N.; Martinez, P.; Díaz, I.; Amaro, A.; Tran, H.; Morales, D.; Pertusa, M.; et al. Excessive Release of Inorganic Polyphosphate by ALS/FTD Astrocytes Causes Non-Cell-Autonomous Toxicity to Motoneurons. Neuron 2022, 110, 1656–1670. [Google Scholar] [CrossRef] [PubMed]
  64. Jensen, B.K.; McAvoy, K.J.; Heinsinger, N.M.; Lepore, A.C.; Ilieva, H.; Haeusler, A.R.; Trotti, D.; Pasinelli, P. Targeting TNFα Produced by Astrocytes Expressing Amyotrophic Lateral Sclerosis-Linked Mutant Fused in Sarcoma Prevents Neurodegeneration and Motor Dysfunction in Mice. Glia 2022, 70, 1426–1449. [Google Scholar] [CrossRef] [PubMed]
  65. Mishra, P.S.; Dhull, D.K.; Nalini, A.; Vijayalakshmi, K.; Sathyaprabha, T.N.; Alladi, P.A.; Raju, T.R. Astroglia Acquires a Toxic Neuroinflammatory Role in Response to the Cerebrospinal Fluid from Amyotrophic Lateral Sclerosis Patients. J. Neuroinflammation 2016, 13, 212. [Google Scholar] [CrossRef] [PubMed]
  66. Lee, S.; Kim, S.; Kang, H.Y.; Lim, H.R.; Kwon, Y.; Jo, M.; Jeon, Y.M.; Kim, S.R.; Kim, K.; Ha, C.M.; et al. The Overexpression of TDP-43 in Astrocytes Causes Neurodegeneration via a PTP1B-Mediated Inflammatory Response. J. Neuroinflammation 2020, 17, 299. [Google Scholar] [CrossRef] [PubMed]
  67. Marchetto, M.C.N.; Muotri, A.R.; Mu, Y.; Smith, A.M.; Cezar, G.G.; Gage, F.H. Non-Cell-Autonomous Effect of Human SOD1G37R Astrocytes on Motor Neurons Derived from Human Embryonic Stem Cells. Cell Stem Cell 2008, 3, 649–657. [Google Scholar] [CrossRef]
  68. Kia, A.; McAvoy, K.; Krishnamurthy, K.; Trotti, D.; Pasinelli, P. Astrocytes Expressing ALS-Linked Mutant FUS Induce Motor Neuron Death through Release of Tumor Necrosis Factor-Alpha. Glia 2018, 66, 1016–1033. [Google Scholar] [CrossRef]
  69. Gomes, C.; Sequeira, C.; Likhite, S.; Dennys, C.N.; Kolb, S.J.; Shaw, P.J.; Vaz, A.R.; Kaspar, B.K.; Meyer, K.; Brites, D. Neurotoxic Astrocytes Directly Converted from Sporadic and Familial ALS Patient Fibroblasts Reveal Signature Diversities and MiR-146a Theragnostic Potential in Specific Subtypes. Cells 2022, 11, 1186. [Google Scholar] [CrossRef]
  70. Di Giorgio, F.P.; Boulting, G.L.; Bobrowicz, S.; Eggan, K.C. Human Embryonic Stem Cell-Derived Motor Neurons Are Sensitive to the Toxic Effect of Glial Cells Carrying an ALS-Causing Mutation. Cell Stem Cell 2008, 3, 637–648. [Google Scholar] [CrossRef]
  71. Chen, Y.; Xia, K.; Chen, L.; Fan, D. Increased Interleukin-6 Levels in the Astrocyte-Derived Exosomes of Sporadic Amyotrophic Lateral Sclerosis Patients. Front. Neurosci. 2019, 13, 574. [Google Scholar] [CrossRef]
  72. Nagai, M.; Re, D.B.; Nagata, T.; Chalazonitis, A.; Jessell, T.M.; Wichterle, H.; Przedborski, S. Astrocytes Expressing ALS-Linked Mutated SOD1 Release Factors Selectively Toxic to Motor Neurons. Nat. Neurosci. 2007, 10, 615–622. [Google Scholar] [CrossRef] [PubMed]
  73. Stoklund Dittlau, K.; Terrie, L.; Baatsen, P.; Kerstens, A.; De Swert, L.; Janky, R.; Corthout, N.; Masrori, P.; Van Damme, P.; Hyttel, P.; et al. FUS-ALS HiPSC-Derived Astrocytes Impair Human Motor Units through Both Gain-of-Toxicity and Loss-of-Support Mechanisms. Mol. Neurodegener. 2023, 18, 5. [Google Scholar] [CrossRef]
  74. Barbosa, M.; Gomes, C.; Sequeira, C.; Gonçalves-Ribeiro, J.; Pina, C.C.; Carvalho, L.A.; Moreira, R.; Vaz, S.H.; Vaz, A.R.; Brites, D. Recovery of Depleted MiR-146a in ALS Cortical Astrocytes Reverts Cell Aberrancies and Prevents Paracrine Pathogenicity on Microglia and Motor Neurons. Front. Cell Dev. Biol. 2021, 9, 634355. [Google Scholar] [CrossRef] [PubMed]
  75. Endo, F.; Komine, O.; Fujimori-Tonou, N.; Katsuno, M.; Jin, S.; Watanabe, S.; Sobue, G.; Dezawa, M.; Wyss-Coray, T.; Yamanaka, K. Astrocyte-Derived TGF-Β1 Accelerates Disease Progression in ALS Mice by Interfering with the Neuroprotective Functions of Microglia and T Cells. Cell Rep. 2015, 11, 592–604. [Google Scholar] [CrossRef]
  76. Madill, M.; McDonagh, K.; Ma, J.; Vajda, A.; McLoughlin, P.; O’Brien, T.; Hardiman, O.; Shen, S. Amyotrophic Lateral Sclerosis Patient IPSC-Derived Astrocytes Impair Autophagy via Non-Cell Autonomous Mechanisms. Mol. Brain 2017, 10, 22. [Google Scholar] [CrossRef]
  77. Tripathi, P.; Rodriguez-Muela, N.; Klim, J.R.; de Boer, A.S.; Agrawal, S.; Sandoe, J.; Lopes, C.S.; Ogliari, K.S.; Williams, L.A.; Shear, M.; et al. Reactive Astrocytes Promote ALS-like Degeneration and Intracellular Protein Aggregation in Human Motor Neurons by Disrupting Autophagy through TGF-Β1. Stem Cell Rep. 2017, 9, 667–680. [Google Scholar] [CrossRef]
  78. Guttenplan, K.A.; Weigel, M.K.; Adler, D.I.; Couthouis, J.; Liddelow, S.A.; Gitler, A.D.; Barres, B.A. Knockout of Reactive Astrocyte Activating Factors Slows Disease Progression in an ALS Mouse Model. Nat. Commun. 2020, 11, 3753. [Google Scholar] [CrossRef] [PubMed]
  79. Qian, K.; Huang, H.; Peterson, A.; Hu, B.; Maragakis, N.J.; Ming, G.L.; Chen, H.; Zhang, S.C. Sporadic ALS Astrocytes Induce Neuronal Degeneration In Vivo. Stem Cell Rep. 2017, 8, 843–855. [Google Scholar] [CrossRef]
  80. Di Giorgio, F.P.; Carrasco, M.A.; Siao, M.C.; Maniatis, T.; Eggan, K. Non-Cell Autonomous Effect of Glia on Motor Neurons in an Embryonic Stem Cell-Based ALS Model. Nat. Neurosci. 2007, 10, 608–614. [Google Scholar] [CrossRef]
  81. Re, D.B.; Le Verche, V.; Yu, C.; Amoroso, M.W.; Politi, K.A.; Phani, S.; Ikiz, B.; Hoffmann, L.; Koolen, M.; Nagata, T.; et al. Necroptosis Drives Motor Neuron Death in Models of Both Sporadic and Familial ALS. Neuron 2014, 81, 1001–1008. [Google Scholar] [CrossRef]
  82. Meyer, K.; Ferraiuolo, L.; Miranda, C.J.; Likhite, S.; McElroy, S.; Renusch, S.; Ditsworth, D.; Lagier-Tourenne, C.; Smith, R.A.; Ravits, J.; et al. Direct Conversion of Patient Fibroblasts Demonstrates Non-Cell Autonomous Toxicity of Astrocytes to Motor Neurons in Familial and Sporadic ALS. Proc. Natl. Acad. Sci. USA 2014, 111, 829–832. [Google Scholar] [CrossRef]
  83. Tong, J.; Huang, C.; Bi, F.; Wu, Q.; Huang, B.; Liu, X.; Li, F.; Zhou, H.; Xia, X.G. Expression of ALS-Linked TDP-43 Mutant in Astrocytes Causes Non-Cell-Autonomous Motor Neuron Death in Rats. EMBO J. 2013, 32, 1917–1926. [Google Scholar] [CrossRef]
  84. Fritz, E.; Izaurieta, P.; Weiss, A.; Mir, F.R.; Rojas, P.; Gonzalez, D.; Rojas, F.; Brown, R.H.; Madrid, R.; van Zundert, B.; et al. Mutant SOD1-Expressing Astrocytes Release Toxic Factors That Trigger Motoneuron Death by Inducing Hyperexcitability. J. Neurophysiol. 2013, 109, 2803–2814. [Google Scholar] [CrossRef]
  85. Cassina, P.; Cassina, A.; Pehar, M.; Castellanos, R.; Gandelman, M.; De León, A.; Robinson, K.M.; Mason, R.P.; Beckman, J.S.; Barbeito, L.; et al. Mitochondrial Dysfunction in SOD1G93A-Bearing Astrocytes Promotes Motor Neuron Degeneration: Prevention by Mitochondrial-Targeted Antioxidants. J. Neurosci. 2008, 28, 4115–4122. [Google Scholar] [CrossRef]
  86. Hou, B.; Zhang, Y.; Liang, P.; He, Y.; Peng, B.; Liu, W.; Han, S.; Yin, J.; He, X. Inhibition of the NLRP3-Inflammasome Prevents Cognitive Deficits in Experimental Autoimmune Encephalomyelitis Mice via the Alteration of Astrocyte Phenotype. Cell Death Dis. 2020, 11, 377. [Google Scholar] [CrossRef]
  87. Haidet-Phillips, A.M.; Hester, M.E.; Miranda, C.J.; Meyer, K.; Braun, L.; Frakes, A.; Song, S.; Likhite, S.; Murtha, M.J.; Foust, K.D.; et al. Astrocytes from Familial and Sporadic ALS Patients Are Toxic to Motor Neurons. Nat. Biotechnol. 2011, 29, 824–828. [Google Scholar] [CrossRef]
  88. Taha, D.M.; Clarke, B.E.; Hall, C.E.; Tyzack, G.E.; Ziff, O.J.; Greensmith, L.; Kalmar, B.; Ahmed, M.; Alam, A.; Thelin, E.P.; et al. Astrocytes Display Cell Autonomous and Diverse Early Reactive States in Familial Amyotrophic Lateral Sclerosis. Brain 2022, 145, 481–489. [Google Scholar] [CrossRef] [PubMed]
  89. Miller, S.J.; Glatzer, J.C.; Hsieh, Y.C.; Rothstein, J.D. Cortical Astroglia Undergo Transcriptomic Dysregulation in the G93A SOD1 ALS Mouse Model. J. Neurogenet. 2018, 32, 322–335. [Google Scholar] [CrossRef] [PubMed]
  90. Ziff, O.J.; Clarke, B.E.; Taha, D.M.; Crerar, H.; Luscombe, N.M.; Patani, R. Meta-Analysis of Human and Mouse ALS Astrocytes Reveals Multi-Omic Signatures of Inflammatory Reactive States. Genome Res. 2022, 32, 71–84. [Google Scholar] [CrossRef] [PubMed]
  91. Ceyzériat, K.; Ben Haim, L.; Denizot, A.; Pommier, D.; Matos, M.; Guillemaud, O.; Palomares, M.A.; Abjean, L.; Petit, F.; Gipchtein, P.; et al. Modulation of Astrocyte Reactivity Improves Functional Deficits in Mouse Models of Alzheimer’s Disease. Acta Neuropathol. Commun. 2018, 6, 104. [Google Scholar] [CrossRef] [PubMed]
  92. Haim, L.B.; Ceyzériat, K.; de Sauvage, M.A.C.; Aubry, F.; Auregan, G.; Guillermier, M.; Ruiz, M.; Petit, F.; Houitte, D.; Faivre, E.; et al. The JAK/STAT3 Pathway Is a Common Inducer of Astrocyte Reactivity in Alzheimer’s and Huntington’s Diseases. J. Neurosci. 2015, 35, 2817–2829. [Google Scholar] [CrossRef]
  93. Lee, H.J.; Suk, J.E.; Patrick, C.; Bae, E.J.; Cho, J.H.; Rho, S.; Hwang, D.; Masliah, E.; Lee, S.J. Direct Transfer of α-Synuclein from Neuron to Astroglia Causes Inflammatory Responses in Synucleinopathies. J. Biol. Chem. 2010, 285, 9262–9272. [Google Scholar] [CrossRef] [PubMed]
  94. Wang, P.; Ye, Y. Filamentous Recombinant Human Tau Activates Primary Astrocytes via an Integrin Receptor Complex. Nat. Commun. 2021, 12, 95. [Google Scholar] [CrossRef] [PubMed]
  95. Panatier, A.; Vallée, J.; Haber, M.; Murai, K.K.; Lacaille, J.C.; Robitaille, R. Astrocytes Are Endogenous Regulators of Basal Transmission at Central Synapses. Cell 2011, 146, 785–798. [Google Scholar] [CrossRef] [PubMed]
  96. Perea, G.; Araque, A. Properties of Synaptically Evoked Astrocyte Calcium Signal Reveal Synaptic Information Processing by Astrocytes. J. Neurosci. 2005, 25, 2192–2203. [Google Scholar] [CrossRef]
  97. Clarke, B.E.; Taha, D.M.; Tyzack, G.E.; Patani, R. Regionally Encoded Functional Heterogeneity of Astrocytes in Health and Disease: A Perspective. Glia 2021, 69, 20–27. [Google Scholar] [CrossRef] [PubMed]
  98. Larramona-Arcas, R.; González-Arias, C.; Perea, G.; Gutiérrez, A.; Vitorica, J.; García-Barrera, T.; Gómez-Ariza, J.L.; Pascua-Maestro, R.; Ganfornina, M.D.; Kara, E.; et al. Sex-Dependent Calcium Hyperactivity Due to Lysosomal-Related Dysfunction in Astrocytes from APOE4 versus APOE3 Gene Targeted Replacement Mice. Mol. Neurodegener. 2020, 15, 35. [Google Scholar] [CrossRef]
  99. Reichenbach, N.; Delekate, A.; Breithausen, B.; Keppler, K.; Poll, S.; Schulte, T.; Peter, J.; Plescher, M.; Hansen, J.N.; Blank, N.; et al. P2Y1 Receptor Blockade Normalizes Network Dysfunction and Cognition in an Alzheimer’s Disease Model. J. Exp. Med. 2018, 215, 1649–1663. [Google Scholar] [CrossRef]
  100. Kuchibhotla, K.V.; Lattarulo, C.R.; Hyman, B.T.; Bacskai, B.J. Synchronous Hyperactivity and Intercellular Calcium Waves in Astrocytes in Alzheimer Mice. Science 2009, 323, 1211–1215. [Google Scholar] [CrossRef]
  101. Oksanen, M.; Petersen, A.J.; Naumenko, N.; Puttonen, K.; Lehtonen, Š.; Gubert Olivé, M.; Shakirzyanova, A.; Leskelä, S.; Sarajärvi, T.; Viitanen, M.; et al. PSEN1 Mutant IPSC-Derived Model Reveals Severe Astrocyte Pathology in Alzheimer’s Disease. Stem Cell Rep. 2017, 9, 1885–1897. [Google Scholar] [CrossRef]
  102. Shah, D.; Gsell, W.; Wahis, J.; Luckett, E.S.; Jamoulle, T.; Vermaercke, B.; Preman, P.; Moechars, D.; Hendrickx, V.; Jaspers, T.; et al. Astrocyte Calcium Dysfunction Causes Early Network Hyperactivity in Alzheimer’s Disease. Cell Rep. 2022, 40, 111280. [Google Scholar] [CrossRef] [PubMed]
  103. Kawamata, H.; Ng, S.K.; Diaz, N.; Burstein, S.; Morel, L.; Osgood, A.; Sider, B.; Higashimori, H.; Haydon, P.G.; Manfredi, G.; et al. Abnormal Intracellular Calcium Signaling and SNARE Dependent Exocytosis Contributes to SOD1G93A Astrocyte- Mediated Toxicity in Amyotrophic Lateral Sclerosis. J. Neurosci. 2014, 34, 2331–2348. [Google Scholar] [CrossRef] [PubMed]
  104. Martorana, F.; Brambilla, L.; Valori, C.F.; Bergamaschi, C.; Roncoroni, C.; Aronica, E.; Volterra, A.; Bezzi, P.; Rossi, D. The BH4 Domain of Bcl-X L Rescues Astrocyte Degeneration in Amyotrophic Lateral Sclerosis by Modulating Intracellular Calcium Signals. Hum. Mol. Genet. 2012, 21, 826–840. [Google Scholar] [CrossRef] [PubMed]
  105. Jiang, R.; Diaz-Castro, B.; Looger, L.L.; Khakh, B.S. Dysfunctional Calcium and Glutamate Signaling in Striatal Astrocytes from Huntington’s Disease Model Mice. J. Neurosci. 2016, 36, 3453–3470. [Google Scholar] [CrossRef] [PubMed]
  106. Diaz-Castro, B.; Gangwani, M.R.; Yu, X.; Coppola, G.; Khakh, B.S. Astrocyte Molecular Signatures in Huntington’s Disease. Sci. Transl. Med. 2019, 11, eaaw8546. [Google Scholar] [CrossRef] [PubMed]
  107. Yu, X.; Taylor, A.M.W.; Nagai, J.; Golshani, P.; Evans, C.J.; Coppola, G.; Khakh, B.S. Reducing Astrocyte Calcium Signaling In Vivo Alters Striatal Microcircuits and Causes Repetitive Behavior. Neuron 2018, 99, 1170–1187. [Google Scholar] [CrossRef]
  108. Bosson, A.; Boisseau, S.; Buisson, A.; Savasta, M.; Albrieux, M. Disruption of Dopaminergic Transmission Remodels Tripartite Synapse Morphology and Astrocytic Calcium Activity within Substantia Nigra Pars Reticulata. Glia 2015, 63, 673–683. [Google Scholar] [CrossRef]
  109. Bosson, A.; Paumier, A.; Boisseau, S.; Jacquier-Sarlin, M.; Buisson, A.; Albrieux, M. TRPA1 Channels Promote Astrocytic Ca2+ Hyperactivity and Synaptic Dysfunction Mediated by Oligomeric Forms of Amyloid-β Peptide. Mol. Neurodegener. 2017, 12, 53. [Google Scholar] [CrossRef]
  110. Paumier, A.; Boisseau, S.; Jacquier-Sarlin, M.; Pernet-Gallay, K.; Buisson, A.; Albrieux, M. Astrocyte-Neuron Interplay Is Critical for Alzheimer’s Disease Pathogenesis and Is Rescued by TRPA1 Channel Blockade. Brain 2022, 145, 388–405. [Google Scholar] [CrossRef]
  111. Pirttimaki, T.M.; Codadu, N.K.; Awni, A.; Pratik, P.; Nagel, D.A.; Hill, E.J.; Dineley, K.T.; Parri, H.R. A7 Nicotinic Receptor-Mediated Astrocytic Gliotransmitter Release: Aβ Effects in a Preclinical Alzheimer’s Mouse Model. PLoS ONE 2013, 8, e81828. [Google Scholar] [CrossRef]
  112. Harada, K.; Kamiya, T.; Tsuboi, T. Gliotransmitter Release from Astrocytes: Functional, Developmental, and Pathological Implications in the Brain. Front. Neurosci. 2016, 9, 499. [Google Scholar] [CrossRef]
  113. Trudler, D.; Sanz-Blasco, S.; Eisele, Y.S.; Ghatak, S.; Bodhinathan, K.; Akhtar, M.W.; Lynch, W.P.; Pinã-Crespo, J.C.; Talantova, M.; Kelly, J.W.; et al. A-Synuclein Oligomers Induce Glutamate Release from Astrocytes and Excessive Extrasynaptic NMDAR Activity in Neurons, Thus Contributing to Synapse Loss. J. Neurosci. 2021, 41, 2264–2273. [Google Scholar] [CrossRef] [PubMed]
  114. Liu, P.-W.; Yue, M.-X.; Zhou, R.; Niu, J.; Huang, D.-J.; Xu, T.; Luo, P.; Liu, X.-H.; Zeng, J.-W. P2Y12 and P2Y13 Receptors Involved in ADPbetas Induced the Release of IL-1beta, IL-6 and TNF-Alpha from Cultured Dorsal Horn Microglia. J. Pain Res. 2017, 10, 1755–1767. [Google Scholar] [CrossRef] [PubMed]
  115. Koizumi, S.; Shigemoto-Mogami, Y.; Nasu-Tada, K.; Shinozaki, Y.; Ohsawa, K.; Tsuda, M.; Joshi, B.V.; Jacobson, K.A.; Kohsaka, S.; Inoue, K. UDP Acting at P2Y6 Receptors Is a Mediator of Microglial Phagocytosis. Nature 2007, 446, 1091–1095. [Google Scholar] [CrossRef] [PubMed]
  116. de Rus Jacquet, A.; Tancredi, J.L.; Lemire, A.L.; Desantis, M.C.; Li, W.P.; O’shea, E.K. The LRRK2 G2019S Mutation Alters Astrocyte-to-Neuron Communication via Extracellular Vesicles and Induces Neuron Atrophy in a Human Ipsc-Derived Model of Parkinson’s Disease. Elife 2021, 10, e73062. [Google Scholar] [CrossRef]
  117. Peterson, S.L.; Nguyen, H.X.; Mendez, O.A.; Anderson, A.J. Complement Protein C3 Suppresses Axon Growth and Promotes Neuron Loss. Sci. Rep. 2017, 7, 12904. [Google Scholar] [CrossRef] [PubMed]
  118. Giaume, C.; Fromaget, C.; el Aoumari, A.; Cordier, J.; Glowinski, J.; Gros, D. Gap Junctions in Cultured Astrocytes: Single-Channel Currents and Characterization of Channel-Forming Protein. Neuron 1991, 6, 133–143. [Google Scholar] [CrossRef] [PubMed]
  119. Dermietrel, R.; Hertzberg, E.L.; Kessler, J.A.; Spray, D.C. Gap Junctions between Cultured Astrocytes: Immunocytochemical, Molecular, and Electrophysiological Analysis. J. Neurosci. 1991, 11, 1421–1432. [Google Scholar] [CrossRef] [PubMed]
  120. Batter, D.K.; Corpina, R.A.; Roy, C.; Spray, D.C.; Hertzberg, E.L.; Kessler, J.A. Heterogeneity in Gap Junction Expression in Astrocytes Cultured from Different Brain Regions. Glia 1992, 6, 213–221. [Google Scholar] [CrossRef]
  121. Giaume, C.; Koulakoff, A.; Roux, L.; Holcman, D.; Rouach, N. Astroglial Networks: A Step Further in Neuroglial and Gliovascular Interactions. Nat. Rev. Neurosci. 2010, 11, 87–99. [Google Scholar] [CrossRef]
  122. Almad, A.A.; Doreswamy, A.; Gross, S.K.; Richard, J.P.; Huo, Y.; Haughey, N.; Maragakis, N.J. Connexin 43 in Astrocytes Contributes to Motor Neuron Toxicity in Amyotrophic Lateral Sclerosis. Glia 2016, 64, 1154–1169. [Google Scholar] [CrossRef]
  123. Almad, A.A.; Taga, A.; Joseph, J.; Gross, S.K.; Welsh, C.; Patankar, A.; Richard, J.-P.; Rust, K.; Pokharel, A.; Plott, C.; et al. Cx43 Hemichannels Contribute to Astrocyte-Mediated Toxicity in Sporadic and Familial ALS. Proc. Natl. Acad. Sci. USA 2022, 119, e2107391119. [Google Scholar] [CrossRef]
  124. Mei, X.; Ezan, P.; Giaume, C.; Koulakoff, A. Astroglial Connexin Immunoreactivity Is Specifically Altered at β-Amyloid Plaques in β-Amyloid Precursor Protein/Presenilin1 Mice. Neuroscience 2010, 171, 92–105. [Google Scholar] [CrossRef] [PubMed]
  125. Basu, R.; Das Sarma, J. Connexin 43/47 Channels Are Important for Astrocyte/Oligodendrocyte Cross-Talk in Myelination and Demyelination. J. Biosci. 2018, 43, 1055–1068. [Google Scholar] [CrossRef] [PubMed]
  126. Zhao, Y.; Yamasaki, R.; Yamaguchi, H.; Nagata, S.; Une, H.; Cui, Y.; Masaki, K.; Nakamuta, Y.; Iinuma, K.; Watanabe, M.; et al. Oligodendroglial Connexin 47 Regulates Neuroinflammation upon Autoimmune Demyelination in a Novel Mouse Model of Multiple Sclerosis. Proc. Natl. Acad. Sci. USA 2020, 117, 2160–2169. [Google Scholar] [CrossRef] [PubMed]
  127. Masaki, K.; Suzuki, S.O.; Matsushita, T.; Matsuoka, T.; Imamura, S.; Yamasaki, R.; Suzuki, M.; Suenaga, T.; Iwaki, T.; Kira, J.I. Connexin 43 Astrocytopathy Linked to Rapidly Progressive Multiple Sclerosis and Neuromyelitis Optica. PLoS ONE 2013, 8, e72919. [Google Scholar] [CrossRef] [PubMed]
  128. Markoullis, K.; Sargiannidou, I.; Schiza, N.; Hadjisavvas, A.; Roncaroli, F.; Reynolds, R.; Kleopa, K.A. Gap Junction Pathology in Multiple Sclerosis Lesions and Normal-Appearing White Matter. Acta Neuropathol. 2012, 123, 873–886. [Google Scholar] [CrossRef] [PubMed]
  129. Angeli, S.; Kousiappa, I.; Stavrou, M.; Sargiannidou, I.; Georgiou, E.; Papacostas, S.S.; Kleopa, K.A. Altered Expression of Glial Gap Junction Proteins Cx43, Cx30, and Cx47 in the 5XFAD Model of Alzheimer’s Disease. Front. Neurosci. 2020, 14, 582934. [Google Scholar] [CrossRef] [PubMed]
  130. Heller, J.P.; Rusakov, D.A. Morphological Plasticity of Astroglia: Understanding Synaptic Microenvironment. Glia 2015, 63, 2133–2151. [Google Scholar] [CrossRef] [PubMed]
  131. Henneberger, C.; Bard, L.; Panatier, A.; Reynolds, J.P.; Kopach, O.; Medvedev, N.I.; Minge, D.; Herde, M.K.; Anders, S.; Kraev, I.; et al. LTP Induction Boosts Glutamate Spillover by Driving Withdrawal of Perisynaptic Astroglia. Neuron 2020, 108, 919–936. [Google Scholar] [CrossRef]
  132. Tani, H.; Dulla, C.G.; Farzampour, Z.; Taylor-Weiner, A.; Huguenard, J.R.; Reimer, R.J. A Local Glutamate-Glutamine Cycle Sustains Synaptic Excitatory Transmitter Release. Neuron 2014, 81, 888–900. [Google Scholar] [CrossRef]
  133. Furness, D.N.; Dehnes, Y.; Akhtar, A.Q.; Rossi, D.J.; Hamann, M.; Grutle, N.J.; Gundersen, V.; Holmseth, S.; Lehre, K.P.; Ullensvang, K.; et al. A Quantitative Assessment of Glutamate Uptake into Hippocampal Synaptic Terminals and Astrocytes: New Insights into a Neuronal Role for Excitatory Amino Acid Transporter 2 (EAAT2). Neuroscience 2008, 157, 80–94. [Google Scholar] [CrossRef] [PubMed]
  134. Danbolt, N.C.; Furness, D.N.; Zhou, Y. Neuronal vs Glial Glutamate Uptake: Resolving the Conundrum. Neurochem. Int. 2016, 98, 29–45. [Google Scholar] [CrossRef]
  135. Roberts, R.C.; Roche, J.K.; McCullumsmith, R.E. Localization of Excitatory Amino Acid Transporters EAAT1 and EAAT2 in Human Postmortem Cortex: A Light and Electron Microscopic Study. Neuroscience 2014, 277, 522–540. [Google Scholar] [CrossRef]
  136. Rothstein, J.D.; Dykes-Hoberg, M.; Pardo, C.A.; Bristol, L.A.; Jin, L.; Kuncl, R.W.; Kanai, Y.; Hediger, M.A.; Wang, Y.; Schielke, J.P.; et al. Knockout of Glutamate Transporters Reveals a Major Role for Astroglial Transport in Excitotoxicity and Clearance of Glutamate. Neuron 1996, 16, 675–686. [Google Scholar] [CrossRef]
  137. Talantova, M.; Sanz-Blasco, S.; Zhang, X.; Xia, P.; Akhtar, M.W.; Okamoto, S.; Dziewczapolski, G.; Nakamura, T.; Cao, G.; Pratt, A.E.; et al. Aβ Induces Astrocytic Glutamate Release, Extrasynaptic NMDA Receptor Activation, and Synaptic Loss. Proc. Natl. Acad. Sci. USA 2013, 110, E2518–E2527. [Google Scholar] [CrossRef] [PubMed]
  138. Scimemi, A.; Meabon, J.S.; Woltjer, R.L.; Sullivan, J.M.; Diamond, J.S.; Cook, D.G. Amyloid-Β1-42 Slows Clearance of Synaptically Released Glutamate by Mislocalizing Astrocytic GLT-1. J. Neurosci. 2013, 33, 5312–5318. [Google Scholar] [CrossRef]
  139. Jacob, C.P.; Koutsilieri, E.; Bartl, J.; Neuen-Jacob, E.; Arzberger, T.; Zander, N.; Ravid, R.; Roggendorf, W.; Riederer, P.; Grünblatt, E. Alterations in Expression of Glutamatergic Transporters and Receptors in Sporadic Alzheimer’s Disease. J. Alzheimer’s Dis. 2007, 11, 97–116. [Google Scholar] [CrossRef] [PubMed]
  140. Lauderback, C.M.; Hackett, J.M.; Huang, F.F.; Keller, J.N.; Szweda, L.I.; Markesbery, W.R.; Allan Butterfield, D. The Glial Glutamate Transporter, GLT-1, Is Oxidatively Modified by 4-Hydroxy-2-Nonenal in the Alzheimer’s Disease Brain: The Role of Aβ1-42. J. Neurochem. 2001, 78, 413–416. [Google Scholar] [CrossRef]
  141. Masliah, E.; Alford, M.; DeTeresa, R.; Mallory, M.; Hansen, L. Deficient Glutamate Transport Is Associated with Neurodegeneration in Alzheimer’s Disease. Ann. Neurol. 1996, 40, 759–766. [Google Scholar] [CrossRef]
  142. Salcedo, C.; Pozo Garcia, V.; García-Adán, B.; Ameen, A.O.; Gegelashvili, G.; Waagepetersen, H.S.; Freude, K.K.; Aldana, B.I. Increased Glucose Metabolism and Impaired Glutamate Transport in Human Astrocytes Are Potential Early Triggers of Abnormal Extracellular Glutamate Accumulation in HiPSC-Derived Models of Alzheimer’s Disease. J. Neurochem. 2023, 1–19. [Google Scholar] [CrossRef]
  143. Zhang, Y.; Meng, X.; Jiao, Z.; Liu, Y.; Zhang, X.; Qu, S. Generation of a Novel Mouse Model of Parkinson’s Disease via Targeted Knockdown of Glutamate Transporter GLT-1 in the Substantia Nigra. ACS Chem. Neurosci. 2020, 11, 406–417. [Google Scholar] [CrossRef]
  144. Iovino, L.; Giusti, V.; Pischedda, F.; Giusto, E.; Plotegher, N.; Marte, A.; Battisti, I.; Di Iacovo, A.; Marku, A.; Piccoli, G.; et al. Trafficking of the Glutamate Transporter Is Impaired in LRRK2-Related Parkinson’s Disease. Acta Neuropathol. 2022, 144, 81–106. [Google Scholar] [CrossRef] [PubMed]
  145. Bruijn, L.I.; Becher, M.W.; Lee, M.K.; Anderson, K.L.; Jenkins, N.A.; Copeland, N.G.; Sisodia, S.S.; Rothstein, J.D.; Borchelt, D.R.; Price, D.L.; et al. ALS-Linked SOD1 Mutant G85R Mediates Damage to Astrocytes and Promotes Rapidly Progressive Disease with SOD1-Containing Inclusions. Neuron 1997, 18, 327–338. [Google Scholar] [CrossRef]
  146. Van Den Bosch, L.; Van Damme, P.; Bogaert, E.; Robberecht, W. The Role of Excitotoxicity in the Pathogenesis of Amyotrophic Lateral Sclerosis. Biochim. Biophys. Acta (BBA)—Mol. Basis Dis. 2006, 1762, 1068–1082. [Google Scholar] [CrossRef] [PubMed]
  147. Wu, J.; Bie, B.; Foss, J.F.; Naguib, M. Amyloid Fibril–Induced Astrocytic Glutamate Transporter Disruption Contributes to Complement C1q-Mediated Microglial Pruning of Glutamatergic Synapses. Mol. Neurobiol. 2020, 57, 2290–2300. [Google Scholar] [CrossRef] [PubMed]
  148. Villanueva, C.B.; Stephensen, H.J.T.; Mokso, R.; Benraiss, A.; Sporring, J.; Goldman, S.A. Astrocytic Engagement of the Corticostriatal Synaptic Cleft Is Disrupted in a Mouse Model of Huntington s Disease. Proc. Natl. Acad. Sci. USA 2023, 120, e2210719129. [Google Scholar] [CrossRef] [PubMed]
  149. Andersen, J.V.; Christensen, S.K.; Westi, E.W.; Diaz-delCastillo, M.; Tanila, H.; Schousboe, A.; Aldana, B.I.; Waagepetersen, H.S. Deficient Astrocyte Metabolism Impairs Glutamine Synthesis and Neurotransmitter Homeostasis in a Mouse Model of Alzheimer’s Disease. Neurobiol. Dis. 2021, 148, 105198. [Google Scholar] [CrossRef] [PubMed]
  150. Andersen, J.V.; Skotte, N.H.; Christensen, S.K.; Polli, F.S.; Shabani, M.; Markussen, K.H.; Haukedal, H.; Westi, E.W.; Diaz-delCastillo, M.; Sun, R.C.; et al. Hippocampal Disruptions of Synaptic and Astrocyte Metabolism Are Primary Events of Early Amyloid Pathology in the 5xFAD Mouse Model of Alzheimer’s Disease. Cell Death Dis. 2021, 12, 954. [Google Scholar] [CrossRef] [PubMed]
  151. Salcedo, C.; Andersen, J.V.; Vinten, K.T.; Pinborg, L.H.; Waagepetersen, H.S.; Freude, K.K.; Aldana, B.I. Functional Metabolic Mapping Reveals Highly Active Branched-Chain Amino Acid Metabolism in Human Astrocytes, Which Is Impaired in IPSC-Derived Astrocytes in Alzheimer’s Disease. Front. Aging Neurosci. 2021, 13, 736580. [Google Scholar] [CrossRef]
  152. Salcedo, C.; Wagner, A.; Andersen, J.V.; Vinten, K.T.; Waagepetersen, H.S.; Schousboe, A.; Freude, K.K.; Aldana, B.I. Downregulation of GABA Transporter 3 (GAT3) Is Associated with Deficient Oxidative GABA Metabolism in Human Induced Pluripotent Stem Cell-Derived Astrocytes in Alzheimer’s Disease. Neurochem. Res. 2021, 46, 2676–2686. [Google Scholar] [CrossRef] [PubMed]
  153. Balu, D.T.; Pantazopoulos, H.; Huang, C.C.Y.; Muszynski, K.; Harvey, T.L.; Uno, Y.; Rorabaugh, J.M.; Galloway, C.R.; Botz-Zapp, C.; Berretta, S.; et al. Neurotoxic Astrocytes Express the D-Serine Synthesizing Enzyme, Serine Racemase, in Alzheimer’s Disease. Neurobiol. Dis. 2019, 130, 104511. [Google Scholar] [CrossRef] [PubMed]
  154. Jo, S.; Yarishkin, O.; Hwang, Y.J.; Chun, Y.E.; Park, M.; Woo, D.H.; Bae, J.Y.; Kim, T.; Lee, J.; Chun, H.; et al. GABA from Reactive Astrocytes Impairs Memory in Mouse Models of Alzheimer’s Disease. Nat. Med. 2014, 20, 886–896. [Google Scholar] [CrossRef] [PubMed]
  155. Shrivastava, A.N.; Kowalewski, J.M.; Renner, M.; Bousset, L.; Koulakoff, A.; Melki, R.; Giaume, C.; Triller, A. β-Amyloid and ATP-Induced Diffusional Trapping of Astrocyte and Neuronal Metabotropic Glutamate Type-5 Receptors. Glia 2013, 61, 1673–1686. [Google Scholar] [CrossRef] [PubMed]
  156. Lian, H.; Yang, L.; Cole, A.; Sun, L.; Chiang, A.C.A.; Fowler, S.W.; Shim, D.J.; Rodriguez-Rivera, J.; Taglialatela, G.; Jankowsky, J.L.; et al. NFκB-Activated Astroglial Release of Complement C3 Compromises Neuronal Morphology and Function Associated with Alzheimer’s Disease. Neuron 2015, 85, 101–115. [Google Scholar] [CrossRef]
  157. Sibille, J.; Pannasch, U.; Rouach, N. Astroglial Potassium Clearance Contributes to Short-Term Plasticity of Synaptically Evoked Currents at the Tripartite Synapse. J. Physiol. 2014, 592, 87–102. [Google Scholar] [CrossRef]
  158. Bellot-Saez, A.; Kékesi, O.; Morley, J.W.; Buskila, Y. Astrocytic Modulation of Neuronal Excitability through K + Spatial Buffering. Neurosci. Biobehav. Rev. 2017, 77, 87–97. [Google Scholar] [CrossRef]
  159. Kelley, K.W.; Ben Haim, L.; Schirmer, L.; Tyzack, G.E.; Tolman, M.; Miller, J.G.; Tsai, H.H.; Chang, S.M.; Molofsky, A.V.; Yang, Y.; et al. Kir4.1-Dependent Astrocyte-Fast Motor Neuron Interactions Are Required for Peak Strength. Neuron 2018, 98, 306–319. [Google Scholar] [CrossRef]
  160. Lindström, V.; Gustafsson, G.; Sanders, L.H.; Howlett, E.H.; Sigvardson, J.; Kasrayan, A.; Ingelsson, M.; Bergström, J.; Erlandsson, A. Extensive Uptake of α-Synuclein Oligomers in Astrocytes Results in Sustained Intracellular Deposits and Mitochondrial Damage. Mol. Cell. Neurosci. 2017, 82, 143–156. [Google Scholar] [CrossRef]
  161. Rostami, J.; Holmqvist, S.; Lindström, V.; Sigvardson, J.; Westermark, G.T.; Ingelsson, M.; Bergström, J.; Roybon, L.; Erlandsson, A. Human Astrocytes Transfer Aggregated Alpha-Synuclein via Tunneling Nanotubes. J. Neurosci. 2017, 37, 11835–11853. [Google Scholar] [CrossRef]
  162. Birger, A.; Ben-Dor, I.; Ottolenghi, M.; Turetsky, T.; Gil, Y.; Sweetat, S.; Perez, L.; Belzer, V.; Casden, N.; Steiner, D.; et al. Human IPSC-Derived Astrocytes from ALS Patients with Mutated C9ORF72 Show Increased Oxidative Stress and Neurotoxicity. EBioMedicine 2019, 50, 274–289. [Google Scholar] [CrossRef]
  163. Jiwaji, Z.; Hardingham, G.E. The Consequences of Neurodegenerative Disease on Neuron-Astrocyte Metabolic and Redox Interactions. Neurobiol. Dis. 2023, 185, 106255. [Google Scholar] [CrossRef]
  164. Vargas, M.R.; Johnson, D.A.; Sirkis, D.W.; Messing, A.; Johnson, J.A. Nrf2 Activation in Astrocytes Protects against Neurodegeneration in Mouse Models of Familial Amyotrophic Lateral Sclerosis. J. Neurosci. 2008, 28, 13574–13581. [Google Scholar] [CrossRef]
  165. Gan, L.; Vargas, M.R.; Johnson, D.A.; Johnson, J.A. Astrocyte-Specific Overexpression of Nrf2 Delays Motor Pathology and Synuclein Aggregation throughout the CNS in the Alpha-Synuclein Mutant (A53T) Mouse Model. J. Neurosci. 2012, 32, 17775–17787. [Google Scholar] [CrossRef] [PubMed]
  166. SantaCruz, K.S.; Yazlovitskaya, E.; Collins, J.; Johnson, J.; DeCarli, C. Regional NAD(P)H:Quinone Oxidoreductase Activity in Alzheimer’s Disease. Neurobiol. Aging 2004, 25, 63–69. [Google Scholar] [CrossRef] [PubMed]
  167. Magistretti, P.J.; Allaman, I. Lactate in the Brain: From Metabolic End-Product to Signalling Molecule. Nat. Rev. Neurosci. 2018, 19, 235–249. [Google Scholar] [CrossRef] [PubMed]
  168. Figley, C.R. Lactate Transport and Metabolism in the Human Brain: Implications for the Astrocyte-Neuron Lactate Shuttle Hypothesis. J. Neurosci. 2011, 31, 4768–4770. [Google Scholar] [CrossRef] [PubMed]
  169. Madji Hounoum, B.; Mavel, S.; Coque, E.; Patin, F.; Vourc’h, P.; Marouillat, S.; Nadal-Desbarats, L.; Emond, P.; Corcia, P.; Andres, C.R.; et al. Wildtype Motoneurons, ALS-Linked SOD1 Mutation and Glutamate Profoundly Modify Astrocyte Metabolism and Lactate Shuttling. Glia 2017, 65, 592–605. [Google Scholar] [CrossRef] [PubMed]
  170. Ferraiuolo, L.; Higginbottom, A.; Heath, P.R.; Barber, S.; Greenald, D.; Kirby, J.; Shaw, P.J. Dysregulation of Astrocyte-Motoneuron Cross-Talk in Mutant Superoxide Dismutase 1-Related Amyotrophic Lateral Sclerosis. Brain 2011, 134, 2627–2641. [Google Scholar] [CrossRef] [PubMed]
  171. Allen, S.P.; Hall, B.; Castelli, L.M.; Francis, L.; Woof, R.; Siskos, A.P.; Kouloura, E.; Gray, E.; Thompson, A.G.; Talbot, K.; et al. Astrocyte Adenosine Deaminase Loss Increases Motor Neuron Toxicity in Amyotrophic Lateral Sclerosis. Brain 2019, 142, 586–605. [Google Scholar] [CrossRef] [PubMed]
  172. Allen, S.P.; Hall, B.; Woof, R.; Francis, L.; Gatto, N.; Shaw, A.C.; Myszczynska, M.; Hemingway, J.; Coldicott, I.; Willcock, A.; et al. C9orf72 Expansion within Astrocytes Reduces Metabolic Flexibility in Amyotrophic Lateral Sclerosis. Brain 2019, 142, 3771–3790. [Google Scholar] [CrossRef]
  173. Strittmatter, W.J.; Saunders, A.M.; Schmechel, D.; Pericak-vance, M.; Enghild, J.; Salvesen, G.S.; Roses, A.D.; Bryan Alzheimer, K. Apolipoprotein E: High-Avidity Binding to B-Amyloid and Increased Frequency of Type 4 Allele in Late-Onset Familial Alzheimer Disease. Proc. Natl. Acad. Sci. USA 1993, 90, 1977–1981. [Google Scholar] [CrossRef] [PubMed]
  174. Kim, J.; Basak, J.M.; Holtzman, D.M. The Role of Apolipoprotein E in Alzheimer’s Disease. Neuron 2009, 63, 287–303. [Google Scholar] [CrossRef]
  175. Jansen, I.E.; Savage, J.E.; Watanabe, K.; Bryois, J.; Williams, D.M.; Steinberg, S.; Sealock, J.; Karlsson, I.K.; Hägg, S.; Athanasiu, L.; et al. Genome-Wide Meta-Analysis Identifies New Loci and Functional Pathways Influencing Alzheimer’s Disease Risk. Nat. Genet. 2019, 51, 404–413. [Google Scholar] [CrossRef]
  176. Shi, Y.; Yamada, K.; Liddelow, S.A.; Smith, S.T.; Zhao, L.; Luo, W.; Tsai, R.M.; Spina, S.; Grinberg, L.T.; Rojas, J.C.; et al. ApoE4 Markedly Exacerbates Tau-Mediated Neurodegeneration in a Mouse Model of Tauopathy. Nature 2017, 549, 523–527. [Google Scholar] [CrossRef]
  177. Reiman, E.M.; Chen, K.; Alexander, G.E.; Caselli, R.J.; Bandy, D.; Osborne, D.; Saunders, A.M.; Hardy, J. Correlations between Apolipoprotein E Ε4 Gene Dose and Brain-Imaging Measurements of Regional Hypometabolism. Proc. Natl. Acad. Sci. USA 2005, 102, 8299–8302. [Google Scholar] [CrossRef] [PubMed]
  178. Williams, H.C.; Farmer, B.C.; Piron, M.A.; Walsh, A.E.; Bruntz, R.C.; Gentry, M.S.; Sun, R.C.; Johnson, L.A. APOE Alters Glucose Flux through Central Carbon Pathways in Astrocytes. Neurobiol. Dis. 2020, 136, 104742. [Google Scholar] [CrossRef] [PubMed]
  179. Lin, Y.T.; Seo, J.; Gao, F.; Feldman, H.M.; Wen, H.L.; Penney, J.; Cam, H.P.; Gjoneska, E.; Raja, W.K.; Cheng, J.; et al. APOE4 Causes Widespread Molecular and Cellular Alterations Associated with Alzheimer’s Disease Phenotypes in Human IPSC-Derived Brain Cell Types. Neuron 2018, 98, 1141–1154.e7. [Google Scholar] [CrossRef]
  180. TCW, J.; Qian, L.; Pipalia, N.H.; Chao, M.J.; Liang, S.A.; Shi, Y.; Jain, B.R.; Bertelsen, S.E.; Kapoor, M.; Marcora, E.; et al. Cholesterol and Matrisome Pathways Dysregulated in Astrocytes and Microglia. Cell 2022, 185, 2213–2233.e25. [Google Scholar] [CrossRef]
  181. Valenza, M.; Marullo, M.; Di Paolo, E.; Cesana, E.; Zuccato, C.; Biella, G.; Cattaneo, E. Disruption of Astrocyte-Neuron Cholesterol Cross Talk Affects Neuronal Function in Huntington’s Disease. Cell Death Differ. 2015, 22, 690–702. [Google Scholar] [CrossRef]
  182. Birolini, G.; Verlengia, G.; Talpo, F.; Maniezzi, C.; Zentilin, L.; Giacca, M.; Conforti, P.; Cordiglieri, C.; Caccia, C.; Leoni, V.; et al. SREBP2 Gene Therapy Targeting Striatal Astrocytes Ameliorates Huntington’s Disease Phenotypes. Brain 2021, 144, 3175–3190. [Google Scholar] [CrossRef]
  183. Kim, J.M.; Cha, S.H.; Choi, Y.R.; Jou, I.; Joe, E.H.; Park, S.M. DJ-1 Deficiency Impairs Glutamate Uptake into Astrocytes via the Regulation of Flotillin-1 and Caveolin-1 Expression. Sci. Rep. 2016, 6, 28823. [Google Scholar] [CrossRef] [PubMed]
  184. Kim, K.S.; Kim, J.S.; Park, J.Y.; Suh, Y.H.; Jou, I.; Joe, E.H.; Park, S.M. DJ-1 Associates with Lipid Rafts by Palmitoylation and Regulates Lipid Rafts-Dependent Endocytosis in Astrocytes. Hum. Mol. Genet. 2013, 22, 4805–4817. [Google Scholar] [CrossRef] [PubMed]
  185. Guttenplan, K.A.; Weigel, M.K.; Prakash, P.; Wijewardhane, P.R.; Hasel, P.; Rufen-Blanchette, U.; Münch, A.E.; Blum, J.A.; Fine, J.; Neal, M.C.; et al. Neurotoxic Reactive Astrocytes Induce Cell Death via Saturated Lipids. Nature 2021, 599, 102–107. [Google Scholar] [CrossRef] [PubMed]
  186. Nagelhus, E.A.; Horio, Y.; Inanobe, A.; Fujita, A.; Haug, F.M.; Nielsen, S.; Kuhachi, Y.; Ottersen, O.P. Immunogold Evidence Suggests That Coupling of K+ Siphoning and Water Transport in Rat Retinal Muller Cells Is Mediated by a Coenrichment of Kir4.1 and AQP4 in Specific Membrane Domains. Glia 1999, 26, 47–54. [Google Scholar] [CrossRef]
  187. Nagelhus, E.A.; Mathiisen, T.M.; Ottersen, O.P. Aquaporin-4 in the Central Nervous System: Cellular and Subcellular Distribution and Coexpression with KIR4.1. Neuroscience 2004, 129, 905–913. [Google Scholar] [CrossRef] [PubMed]
  188. Amiry-Moghaddam, M.; Williamson, A.; Palomba, M.; Eid, T.; de Lanerolle, N.C.; Nagelhus, E.A.; Adams, M.E.; Froehner, S.C.; Agre, P.; Ottersen, O.P. Delayed K+ Clearance Associated with Aquaporin-4 Mislocalization: Phenotypic Defects in Brains of α-Syntrophin-Null Mice. Proc. Natl. Acad. Sci. USA 2003, 100, 13615–13620. [Google Scholar] [CrossRef] [PubMed]
  189. Caron, N.S.; Banos, R.; Yanick, C.; Aly, A.E.; Byrne, L.M.; Smith, E.D.; Xie, Y.; Smith, S.E.P.; Potluri, N.; Black, H.F.; et al. Mutant Huntingtin Is Cleared from the Brain via Active Mechanisms in Huntington Disease. J. Neurosci. 2021, 41, 780–796. [Google Scholar] [CrossRef] [PubMed]
  190. Iliff, J.J.; Wang, M.; Liao, Y.; Plogg, B.A.; Peng, W.; Gundersen, G.A.; Benveniste, H.; Vates, G.E.; Deane, R.; Goldman, S.A.; et al. A Paravascular Pathway Facilitates CSF Flow through the Brain Parenchyma and the Clearance of Interstitial Solutes, Including Amyloid β. Sci. Transl. Med. 2012, 4, 147ra111. [Google Scholar] [CrossRef]
  191. Xu, Z.; Xiao, N.; Chen, Y.; Huang, H.; Marshall, C.; Gao, J.; Cai, Z.; Wu, T.; Hu, G.; Xiao, M. Deletion of Aquaporin-4 in APP/PS1 Mice Exacerbates Brain Aβ Accumulation and Memory Deficits. Mol. Neurodegener. 2015, 10, 58. [Google Scholar] [CrossRef]
  192. Hoshi, A.; Yamamoto, T.; Shimizu, K.; Ugawa, Y.; Nishizawa, M.; Takahashi, H.; Kakita, A. Characteristics of Aquaporin Expression Surrounding Senile Plaques and Cerebral Amyloid Angiopathy in Alzheimer Disease. J. Neuropathol. Exp. Neurol. 2012, 71, 750–759. [Google Scholar] [CrossRef] [PubMed]
  193. Castellano, J.M.; Kim, J.; Stewart, F.R.; Jiang, H.; DeMattos, R.B.; Patterson, B.W.; Fagan, A.M.; Morris, J.C.; Mawuenyega, K.G.; Cruchaga, C.; et al. Human ApoE Isoforms Differentially Regulate Brain Amyloid-β Peptide Clearance. Sci. Transl. Med. 2011, 3, 89ra57. [Google Scholar] [CrossRef] [PubMed]
  194. Streubel-Gallasch, L.; Giusti, V.; Sandre, M.; Tessari, I.; Plotegher, N.; Giusto, E.; Masato, A.; Iovino, L.; Battisti, I.; Arrigoni, G.; et al. Parkinson’s Disease-Associated LRRK2 Interferes with Astrocyte-Mediated Alpha-Synuclein Clearance. Mol. Neurobiol. 2021, 58, 3119–3140. [Google Scholar] [CrossRef]
  195. Loria, F.; Vargas, J.Y.; Bousset, L.; Syan, S.; Salles, A.; Melki, R.; Zurzolo, C. α-Synuclein Transfer between Neurons and Astrocytes Indicates That Astrocytes Play a Role in Degradation Rather than in Spreading. Acta Neuropathol. 2017, 134, 789–808. [Google Scholar] [CrossRef]
  196. Dilsizoglu Senol, A.; Samarani, M.; Syan, S.; Guardia, C.M.; Nonaka, T.; Liv, N.; Latour-Lambert, P.; Hasegawa, M.; Klumperman, J.; Bonifacino, J.S.; et al. α-Synuclein Fibrils Subvert Lysosome Structure and Function for the Propagation of Protein Misfolding between Cells through Tunneling Nanotubes. PLoS. Biol. 2021, 19, e3001287. [Google Scholar] [CrossRef]
  197. Mazzulli, J.R.; Xu, Y.-H.; Sun, Y.; Knight, A.L.; McLean, P.J.; Caldwell, G.A.; Sidransky, E.; Grabowski, G.A.; Krainc, D. Gaucher Disease Glucocerebrosidase and α-Synuclein Form a Bidirectional Pathogenic Loop in Synucleinopathies. Cell 2011, 146, 37–52. [Google Scholar] [CrossRef] [PubMed]
  198. Morrone Parfitt, G.; Coccia, E.; Goldman, C.; Whitney, K.; Reyes, R.; Sarrafha, L.; Nam, K.H.; Sohail, S.; Jones, D.R.; Crary, J.F.; et al. Disruption of Lysosomal Proteolysis in Astrocytes Facilitates Midbrain Organoid Proteostasis Failure in an Early-Onset Parkinson’s Disease Model. Nat. Commun. 2024, 15, 447. [Google Scholar] [CrossRef]
  199. Perea, J.R.; López, E.; Diéz-Ballesteros, J.C.; Ávila, J.; Hernández, F.; Bolós, M. Extracellular Monomeric Tau Is Internalized by Astrocytes. Front. Neurosci. 2019, 13, 442. [Google Scholar] [CrossRef]
  200. Grubman, A.; Chew, G.; Ouyang, J.F.; Sun, G.; Choo, X.Y.; McLean, C.; Simmons, R.K.; Buckberry, S.; Vargas-Landin, D.B.; Poppe, D.; et al. A Single-Cell Atlas of Entorhinal Cortex from Individuals with Alzheimer’s Disease Reveals Cell-Type-Specific Gene Expression Regulation. Nat. Neurosci. 2019, 22, 2087–2097. [Google Scholar] [CrossRef]
  201. Rostami, J.; Mothes, T.; Kolahdouzan, M.; Eriksson, O.; Moslem, M.; Bergström, J.; Ingelsson, M.; O’Callaghan, P.; Healy, L.M.; Falk, A.; et al. Crosstalk between Astrocytes and Microglia Results in Increased Degradation of α-Synuclein and Amyloid-β Aggregates. J. Neuroinflammation 2021, 18, 124. [Google Scholar] [CrossRef]
  202. Nakamura, S.; Yoshimori, T. New Insights into Autophagosome–Lysosome Fusion. J. Cell Sci. 2017, 130, 1209–1216. [Google Scholar] [CrossRef]
  203. Kulkarni, A.; Dong, A.; Kulkarni, V.V.; Chen, J.; Laxton, O.; Anand, A.; Maday, S. Differential Regulation of Autophagy during Metabolic Stress in Astrocytes and Neurons. Autophagy 2020, 16, 1651–1667. [Google Scholar] [CrossRef]
  204. Jäger, S.; Bucci, C.; Tanida, I.; Ueno, T.; Kominami, E.; Saftig, P.; Eskelinen, E.L. Role for Rab7 in Maturation of Late Autophagic Vacuoles. J. Cell Sci. 2004, 117, 4837–4848. [Google Scholar] [CrossRef] [PubMed]
  205. Funayama, M.; Nishioka, K.; Li, Y.; Hattori, N. Molecular Genetics of Parkinson’s Disease: Contributions and Global Trends. J. Hum. Genet. 2023, 68, 125–130. [Google Scholar] [CrossRef]
  206. Madureira, M.; Connor-Robson, N.; Wade-Martins, R. LRRK2: Autophagy and Lysosomal Activity. Front. Neurosci. 2020, 14, 498. [Google Scholar] [CrossRef] [PubMed]
  207. di Domenico, A.; Carola, G.; Calatayud, C.; Pons-Espinal, M.; Muñoz, J.P.; Richaud-Patin, Y.; Fernandez-Carasa, I.; Gut, M.; Faella, A.; Parameswaran, J.; et al. Patient-Specific IPSC-Derived Astrocytes Contribute to Non-Cell-Autonomous Neurodegeneration in Parkinson’s Disease. Stem Cell Rep. 2019, 12, 213–229. [Google Scholar] [CrossRef]
  208. Szebényi, K.; Wenger, L.M.D.; Sun, Y.; Dunn, A.W.E.; Limegrover, C.A.; Gibbons, G.M.; Conci, E.; Paulsen, O.; Mierau, S.B.; Balmus, G.; et al. Human ALS/FTD Brain Organoid Slice Cultures Display Distinct Early Astrocyte and Targetable Neuronal Pathology. Nat. Neurosci. 2021, 24, 1542–1554. [Google Scholar] [CrossRef]
  209. Liddelow, S.; Barres, B. SnapShot: Astrocytes in Health and Disease. Cell 2015, 162, 1170.e1. [Google Scholar] [CrossRef] [PubMed]
  210. Nicaise, C.; Soyfoo, M.S.; Authelet, M.; De Decker, R.; Bataveljic, D.; Delporte, C.; Pochet, R. Aquaporin-4 Overexpression in Rat ALS Model. Anat. Rec. 2009, 292, 207–213. [Google Scholar] [CrossRef] [PubMed]
  211. Garbuzova-Davis, S.; Haller, E.; Saporta, S.; Kolomey, I.; Nicosia, S.V.; Sanberg, P.R. Ultrastructure of Blood-Brain Barrier and Blood-Spinal Cord Barrier in SOD1 Mice Modeling ALS. Brain Res. 2007, 1157, 126–137. [Google Scholar] [CrossRef]
  212. Garbuzova-Davis, S.; Saporta, S.; Haller, E.; Kolomey, I.; Bennett, S.P.; Potter, H.; Sanberg, P.R. Evidence of Compromised Blood-Spinal Cord Barrier in Early and Late Symptomatic SOD1 Mice Modeling ALS. PLoS ONE 2007, 2, e1205. [Google Scholar] [CrossRef]
  213. Miyazaki, K.; Ohta, Y.; Nagai, M.; Morimoto, N.; Kurata, T.; Takehisa, Y.; Ikeda, Y.; Matsuura, T.; Abe, K. Disruption of Neurovascular Unit Prior to Motor Neuron Degeneration in Amyotrophic Lateral Sclerosis. J. Neurosci. Res. 2011, 89, 718–728. [Google Scholar] [CrossRef]
  214. Winkler, E.A.; Sengillo, J.D.; Sullivan, J.S.; Henkel, J.S.; Appel, S.H.; Zlokovic, B.V. Blood-Spinal Cord Barrier Breakdown and Pericyte Reductions in Amyotrophic Lateral Sclerosis. Acta Neuropathol. 2013, 125, 111–120. [Google Scholar] [CrossRef]
  215. Halliday, M.R.; Rege, S.V.; Ma, Q.; Zhao, Z.; Miller, C.A.; Winkler, E.A.; Zlokovic, B.V. Accelerated Pericyte Degeneration and Blood-Brain Barrier Breakdown in Apolipoprotein E4 Carriers with Alzheimer’s Disease. J. Cereb. Blood Flow Metab. 2016, 36, 216–227. [Google Scholar] [CrossRef]
  216. Bell, R.D.; Winkler, E.A.; Singh, I.; Sagare, A.P.; Deane, R.; Wu, Z.; Holtzman, D.M.; Betsholtz, C.; Armulik, A.; Sallstrom, J.; et al. Apolipoprotein E Controls Cerebrovascular Integrity via Cyclophilin A. Nature 2012, 485, 512–516. [Google Scholar] [CrossRef] [PubMed]
  217. Nishitsuji, K.; Hosono, T.; Nakamura, T.; Bu, G.; Michikawa, M. Apolipoprotein E Regulates the Integrity of Tight Junctions in an Isoform-Dependent Manner in an in Vitro Blood-Brain Barrier Model. J. Biol. Chem. 2011, 286, 17536–17542. [Google Scholar] [CrossRef] [PubMed]
  218. Spampinato, S.F.; Merlo, S.; Fagone, E.; Fruciano, M.; Barbagallo, C.; Kanda, T.; Sano, Y.; Purrello, M.; Vancheri, C.; Ragusa, M.; et al. Astrocytes Modify Migration of Pbmcs Induced by β-Amyloid in a Blood-Brain Barrier in Vitro Model. Front. Cell Neurosci. 2019, 13, 337. [Google Scholar] [CrossRef] [PubMed]
  219. Hsiao, H.Y.; Chen, Y.C.; Huang, C.H.; Chen, C.C.; Hsu, Y.H.; Chen, H.M.; Chiu, F.L.; Kuo, H.C.; Chang, C.; Chern, Y. Aberrant Astrocytes Impair Vascular Reactivity in Huntington Disease. Ann. Neurol. 2015, 78, 178–192. [Google Scholar] [CrossRef] [PubMed]
  220. de Rus Jacquet, A.; Alpaugh, M.; Denis, H.L.; Tancredi, J.L.; Boutin, M.; Decaestecker, J.; Beauparlant, C.; Herrmann, L.; Saint-Pierre, M.; Parent, M.; et al. The Contribution of Inflammatory Astrocytes to BBB Impairments in a Brain-Chip Model of Parkinson’s Disease. Nat. Commun. 2023, 14, 3651. [Google Scholar] [CrossRef] [PubMed]
  221. D’Ambrosi, N.; Apolloni, S. Fibrotic Scar in Neurodegenerative Diseases. Front. Immunol. 2020, 11, 1394. [Google Scholar] [CrossRef]
  222. Anderson, M.A.; Burda, J.E.; Ren, Y.; Ao, Y.; O’Shea, T.M.; Kawaguchi, R.; Coppola, G.; Khakh, B.S.; Deming, T.J.; Sofroniew, M.V. Astrocyte Scar Formation Aids Central Nervous System Axon Regeneration. Nature 2016, 532, 195–200. [Google Scholar] [CrossRef]
  223. Kamermans, A.; Planting, K.E.; Jalink, K.; van Horssen, J.; de Vries, H.E. Reactive Astrocytes in Multiple Sclerosis Impair Neuronal Outgrowth through TRPM7-Mediated Chondroitin Sulfate Proteoglycan Production. Glia 2019, 67, 68–77. [Google Scholar] [CrossRef]
  224. Shijo, T.; Warita, H.; Suzuki, N.; Kitajima, Y.; Ikeda, K.; Akiyama, T.; Ono, H.; Mitsuzawa, S.; Nishiyama, A.; Izumi, R.; et al. Aberrant Astrocytic Expression of Chondroitin Sulfate Proteoglycan Receptors in a Rat Model of Amyotrophic Lateral Sclerosis. J. Neurosci. Res. 2018, 96, 222–233. [Google Scholar] [CrossRef]
  225. Mizuno, H.; Warita, H.; Aoki, M.; Itoyama, Y. Accumulation of Chondroitin Sulfate Proteoglycans in the Microenvironment of Spinal Motor Neurons in Amyotrophic Lateral Sclerosis Transgenic Rats. J. Neurosci. Res. 2008, 86, 2512–2523. [Google Scholar] [CrossRef] [PubMed]
  226. Peters, S.; Zitzelsperger, E.; Kuespert, S.; Iberl, S.; Heydn, R.; Johannesen, S.; Petri, S.; Aigner, L.; Thal, D.R.; Hermann, A.; et al. The TGF-β System as a Potential Pathogenic Player in Disease Modulation of Amyotrophic Lateral Sclerosis. Front. Neurol. 2017, 8, 669. [Google Scholar] [CrossRef] [PubMed]
  227. Schachtrup, C.; Ryu, J.K.; Helmrick, M.J.; Vagena, E.; Galanakis, D.K.; Degen, J.L.; Margolis, R.U.; Akassoglou, K. Fibrinogen Triggers Astrocyte Scar Formation by Promoting the Availability of Active TGF-β after Vascular Damage. J. Neurosci. 2010, 30, 5843–5854. [Google Scholar] [CrossRef] [PubMed]
  228. Gris, P.; Tighe, A.; Levin, D.; Sharma, R.; Brown, A. Transcriptional Regulation of Scar Gene Expression in Primary Astrocytes. Glia 2007, 55, 1145–1155. [Google Scholar] [CrossRef] [PubMed]
  229. Lian, H.; Zheng, H. Signaling Pathways Regulating Neuron–Glia Interaction and Their Implications in Alzheimer’s Disease. J. Neurochem. 2016, 136, 475–491. [Google Scholar] [CrossRef] [PubMed]
  230. Herrmann, J.E.; Imura, T.; Song, B.; Qi, J.; Ao, Y.; Nguyen, T.K.; Korsak, R.A.; Takeda, K.; Akira, S.; Sofroniew, M.V. STAT3 Is a Critical Regulator of Astrogliosis and Scar Formation after Spinal Cord Injury. J. Neurosci. 2008, 28, 7231–7243. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Common disease hallmarks in neurodegeneration. Figure is created with biorender.com (accessed on 28 January 2024).
Figure 1. Common disease hallmarks in neurodegeneration. Figure is created with biorender.com (accessed on 28 January 2024).
Biomolecules 14 00289 g001
Figure 2. Common mechanisms involved in astrocytic pathology in neurodegenerative diseases. Figure is created with biorender.com (accessed on 28 Janurary 2024).
Figure 2. Common mechanisms involved in astrocytic pathology in neurodegenerative diseases. Figure is created with biorender.com (accessed on 28 Janurary 2024).
Biomolecules 14 00289 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Stoklund Dittlau, K.; Freude, K. Astrocytes: The Stars in Neurodegeneration? Biomolecules 2024, 14, 289. https://doi.org/10.3390/biom14030289

AMA Style

Stoklund Dittlau K, Freude K. Astrocytes: The Stars in Neurodegeneration? Biomolecules. 2024; 14(3):289. https://doi.org/10.3390/biom14030289

Chicago/Turabian Style

Stoklund Dittlau, Katarina, and Kristine Freude. 2024. "Astrocytes: The Stars in Neurodegeneration?" Biomolecules 14, no. 3: 289. https://doi.org/10.3390/biom14030289

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop