Open access peer-reviewed chapter

Astrocytes: Initiators of and Responders to Inflammation

Written By

Allison Soung and Robyn S. Klein

Submitted: 31 August 2019 Reviewed: 16 September 2019 Published: 05 November 2019

DOI: 10.5772/intechopen.89760

From the Edited Volume

Glia in Health and Disease

Edited by Tania Spohr

Chapter metrics overview

1,447 Chapter Downloads

View Full Metrics

Abstract

We are in the midst of a glial renaissance; astrocytes, essential for brain homeostasis and neuroprotection, have experienced resurgence in focused analyses. New roles in synaptic plasticity, innate immunity and control of recruited immune cells have placed astrocytes at the center of central nervous system functions. Astrocytes have been shown to receive and convey information to all neural cell types in a coordinated effort to respond to injury and infection, initiating reparative mechanisms. Astrocytes detect injury and infection signals from neurons, microglia, oligodendrocytes and endothelial cells, responding by secreting cytokines, chemokines and growth factors, which may activate immune defenses. While regional heterogeneity in astrocyte form and function has been appreciated since the early 1990s, technologic advances have allowed scientists to show only that astrocytes may be as individualized as neurons. Adult astrocytes may undergo a morphological and functional transformation referred to as astrogliosis. Newly generated astrocytes exhibit heterogenous phenotypes; thus, some remove toxic molecules, restore blood-brain barrier function, and promote extracellular matrix components to support axonal growth and repair, while others inhibit neuronal repair and regeneration. This chapter will introduce some of the cellular and molecular components involved in astrocyte responses induced by inflammatory mediators or pathogens during neuroinflammation or neuroinfectious diseases.

Keywords

  • astrocyte
  • cytokines
  • chemokines
  • pathogens
  • viruses
  • bacteria
  • astrogliosis

1. Introduction

Astrocytes are a principle participant in central nervous system (CNS) responses to neurological disorders or diseases [1, 2, 3]. During development and homeostasis, astrocytes coordinate immune responses by regulating microglia activation and blood-brain barrier (BBB) formation [4, 5]. Through dedicated molecular cascades, astrocytes also provide growth factors to neurons, support synapse formation, and help regulate extracellular balance of ions and neurotransmitters, making these glial cells essential for brain homeostasis [6, 7]. In response to CNS injury and disease, astrocytes undergo a process termed astrogliosis, a multifactorial and complex remodeling of astrocytes [7, 8, 9, 10]. Despite the use of a single term to describe astrocyte reaction to insult, astrogliosis results in a spectrum of heterogenous changes in a context specific manner that vary with etiology and severity of CNS injury [9, 10, 11, 12, 13]. Classically, this process is characterized by upregulation of glial fibrillary acid protein (GFAP) and vimentin, key astrocyte intermediate filaments, and hypertrophy of astrocyte processes [14] (Figure 1). Changes in astrocyte biochemistry and physiology that may result in the secretion of anti-inflammatory and pro-inflammatory factors also contribute to this process [10, 15, 16, 17].

Figure 1.

Astrogliosis is typically characterized by hypertrophy of astrocyte processes. Expression of GFAP+ astrocytes (green) and DAPI (blue) in a mock-infected mouse CA3 hippocampus (A) and in a West Nile virus-infected CA3 hippocampus 7 days post infection (B). Viral infection triggers reactive astrogliosis in the hippocampus, in which the processes of activated astrocytes show hypertrophy.

Advertisement

2. Astrogliosis

Functionally, astrogliosis results in the expression of molecules that provide neurotrophic support to injured neurons, isolate damaged area and CNS inflammation from healthy CNS tissue, rebuild and maintain a compromised BBB, and contribute circuitry remodeling around the lesioned region [7, 9, 10, 11, 12, 18]. Consistent with this, studies using animal models of traumatic brain injury, spinal cord injury, and autoimmunity, all reveal that the loss of reactive astrocytes during acute processes leads to the exacerbation of clinical symptoms, recruit of immune molecules, changes in BBB integrity, and neuronal death [710, 19]. The overall goals of these functional reactions are therefore beneficial for the CNS. However, past research has also highlighted detrimental and inhibitory effects of astrogliosis, including augmentation of inflammation, as well as inhibition of neuronal repair and axonal growth [20, 21]. The dual outcomes of astrogliosis highlight the time- and context-specific way this process may be regulated. Future studies of this process may ultimately determine mechanisms to manipulate astrogliosis as a therapeutic target to improve CNS injury outcomes [10, 22].

Astrogliosis is induced and regulated by a variety of extracellular molecules, such as neurotransmitters, steroid hormones, cytokines and neurodegeneration-associated molecules (Table 1). Intracellular signaling pathways, such cyclic AMP (cAMP), signal transducer and activator of transcription 3 (STAT3), nuclear factor kappa B (NFκB), Rho-kinase, and calcium have all been observed to induce the expression of GFAP or vimentin [11, 45, 46, 47]. Extracellular signaling pathways, including responses to epidermal growth factor (EGF), fibroblast growth factor (FGF), sonic hedgehog (SHH), and albumin, can also regulate astrocyte proliferation [9, 48, 49, 50]. Specific pro- and anti-inflammatory effects of reactive astrocytes may be regulated separately. Thus, the genetic ablation of STAT3 within astrocytes, or its associated membrane receptor gp130, leads to increased inflammation during autoimmune disease, traumatic injury and infection [24, 37, 38, 39, 51], while genetic deletion of NFκB or the suppressor of cytokine signaling 3 (Soc3) signaling pathway in astrocytes decreases the recruitment of immune cells [31, 32, 37]. Furthermore, recruited immune cells release numerous cytokines that may further stimulate astrocyte activation (Table 1). In addition, recent studies indicate that microglia critically induce astrogliosis via expression of pro-inflammatory cytokines, including interleukin (IL)-1β, tumor necrosis factor (TNF), and interferon (IFN)-γ [26, 52, 53].

Signaling pathwayInjuryChemokines/cytokines releasedImmune/functional outcomeReferences
ERα signalingEAEReduction of leukocyte molecules[23]
Gp103/IL-6 signalingEAEDownregulation of IL-17 and IFNγReduction of T cell infiltration, inhibition of astrocyte apoptosis, improvement of disease course[24, 25]
InfectionDownregulation of IFNγInhibition of astrocyte apoptosis, decrease of pathogen burden[25]
IL-1β signalingTraumatic injury, infectionIncrease of GFAP expression[26, 27]
IL-17 signalingEAEUpregulation of CXCL2Increase of leukocyte infiltration, worsen disease course[28]
IFNγ signalingEAEDownregulation of CCL5, IL-1β and TNFImproved course of disease[24]
Traumatic injuryIncrease of GFAP expression[29]
TNFR1 signalingEAEIncreases of T cell infiltration, worsen disease course[30]
NFκB signalingEAEUpregulation of CCL2, CCL5, CXCL10, IL-1β, IFNγ, and TNF; downregulation of IL-6Reduction of leukocyte molecules, increase in axon pathology, worsen disease course[31]
Ischemia, traumatic injuryUpregulation of CCL2, CCL5, CXCL10, IL-6, TGF-β and TNFIncrease of leukocytes molecules, reduction of GFAP expression, increase of neuronal damage[31, 32, 33]
Notch signalingIschemiaReduction of leukocyte molecules, increase of GFAP expression, increase of astrocyte proliferation[34, 35]
SHH signalingEAEMaintenance of BBB[36]
Soc3 signalingTraumatic injuryIncrease of leukocyte molecules, increase of GFAP expression[37]
STAT3 signalingTraumatic injuryReduction of leukocyte molecules, inhibition of GFAP expression[37, 38]
Traumatic injuryIncrease of GFAP and vimentin expression[39, 40]
TGF-β signalingTraumatic injuryIncrease of leukocyte molecules, increase of GFAP expression, increase of ECM components[41, 42, 43]
InfectionInhibition of NFκB signaling; downregulation of CCL5Reduction of T cell infiltration, decrease of neuronal death[44]

Table 1.

Triggers of reactive astrogliosis.

A variety of intracellular signaling molecules have been shown to induce reactive astrogliosis or to modulate aspects of the reactive astrogliosis process. In response to a range of CNS injuries, all cell types within the CNS, such as neurons, microglia, other astrocytes, endothelium, and pericytes, can release signaling molecules that are able to trigger astrogliosis.

BBB = blood-brain barrier, CCL = chemokine (C-C motif), CXCL = chemokine (C-X-C motif) ligand, ER = estrogen receptor, Gp = glycoprotein, IL = interleukin, IFN = interferon, NFκB = nuclear factor kappa B, EAE = experimental autoimmune encephalomyelitis, ECM = extracellular matrix, SHH = sonic hedgehog, Soc3 = suppressor of cytokine signaling 3, STAT3 = signaling transducer and activator of transcription 3, TGF = transforming growth factor β, TNF = tumor necrosis factor.

In response to injury, reactive astrocytes were previously believed to migrate to the lesion site. Recent live imaging studies, however, indicate that astrocytes do not migrate towards the lesion site [54]. Instead, astrocytes remain in their tiled-domains and become hypertrophic [54, 55]. Neither proliferation nor migration of astrocytes contribute to the total increase of GFAP positive cells observed at lesion sites. This has led to a new focus on identifying other sources for adult astrocytes. Currently, there is evidence that radial glia, neuronal progenitor cells (NPCs) within the subventricular (SVZ) and subgranular (SGZ) zones, locally proliferating glia, in addition to NG2+ cells may all contribute to newly generated pools of reactive astrocytes after injury [56].

Advertisement

3. Astrocytes as the gatekeeper to the CNS

During homeostasis, astrocyte end-feet enwrap the brain microvascular endothelial cells, helping maintain the integrity of the BBB. Their physical interaction with the BBB allows astrocytes to influence the entry of peripheral immune cells into the CNS during injury or disease as well as modulating their activity once entering the CNS parenchyma. In health, astrocytes, along with multiple other cell types, support the BBB as well as express localizing cues that restrict leukocytes access into the CNS parenchyma [17, 57, 58, 59]. However, CNS damage caused by stroke, traumatic injury, infection, autoimmune disease, and neurodegenerative disorders leads to the disruption of the BBB, which may increase the CNS entry of immune cells [24, 25, 38, 39, 60, 61, 62, 63, 64, 65, 66].

During injury or infection, astrocytes detect molecular changes in their extracellular environment and in neighboring cells. In stroke, astrocytes become reactive when oxygen and glucose deprivation occurs [67, 68]. In most neurological disorders, the release of neurotransmitters and adenosine triphosphate (ATP) from damaged neurons is detected by astrocytes via P2X and P2Y purinergic receptors [69, 70]. During viral infections, toll-like receptors (TLRs), such as TLR3, 7, and 9, and retinoic acid-inducible gene I (RIG-I)-like receptors (RLRs), are expressed on neurons, astrocytes, and microglia. These receptors are examples of pattern recognition receptors (PPRs) that are differentially activated by pathogen-associated molecular patterns (PAMPs) derived from invading bacteria, fungi, or viruses [71]. Activation of TLRs and RLRs by PAMPs or damage-associated molecular patterns (DAMPs) have been shown to contribute to neuronal damage, induce microglia and astrocyte activation and production of cytokines, including type I IFNs [72, 73, 74]. Type I IFNs, along with numerous other innate cytokines, such as IL-6, IL-1β, IFN-γ, and TNF, have been shown to regulate BBB integrity through a variety of different mechanisms that include the regulation of Rho GTPases, activation of matrix metallopeptidase 9 (MMP9), and suppression of other pro-inflammatory cytokines, including IL-1β, IL-6, and TNF [75, 76]. In support of this, the genetic astrocyte-specific deletion of the type I IFN receptor, IFNαβR (IFNAR), results in enhanced BBB permeability in a murine viral infection model [77].

The entry of leukocytes into the CNS parenchyma involves their passage across the BBB, whose permeability is regulated by astrocytes and pericytes, as well as multiple other cell types [57, 58, 59]. Once leukocytes traverse the BBB, they localize within perivascular spaces and where they interact with numerous cell types, including astrocytes [78]. Astrocytes, thus, take part in both the recruitment and restriction of leukocytes in the CNS [58, 59, 151]. Their functions, however, occur in a context-specific manner via specific signaling events. It is remarkable how astrocytes are able to respond to a diverse number of signaling mechanisms in the orchestration BBB disruption, the recruitment of leukocytes, and the amplification of their pro-inflammatory effects [17, 57, 79, 80], while also being capable of contributing to BBB repair, restricting leukocyte trafficking, and exerting anti-inflammatory effects that promote the resolution of inflammation [6, 9, 10, 11].

Advertisement

4. Reactive astrocytes as a physical barrier

At the site of injury, newly proliferated astrocytes form scars, in which bundles of reactive astrocytes polarize with extracellular matrix (ECM) components and physically surround the lesioned site [38]. The earliest studies focused on the formation of the astrocyte scar and its importance in repairing the BBB after traumatic brain injury [61, 62]. Astrocyte scars form a physical, functional barrier that restricts the entry of leukocytes after traumatic brain injuries, ischemia, neurodegeneration and autoimmune inflammation [37, 38]. This is achieved through the upregulation of ECM proteins, such as fibronectin and laminin, as well as chondroitin sulfate proteoglycans (CSPGs) [41, 42, 81, 82, 83, 84]. Structural proteins, such as GFAP and vimentin, have also been shown to be important for the formation of the astrocyte scar [14]. Mice with global genetic deletion of these molecules display increased inflammation and pathology as well as worsened functional outcomes in various CNS injury models, such as ischemia, traumatic injury, autoimmune inflammation, infection, and neurodegeneration [11, 63, 64, 85, 86, 87, 88].

The astrocyte scar is also important for localizing immune cells and limiting the invasion of infectious pathogens, to the lesion site. For example, the genetic deletion of GFAP+ cells leads to increases in immune cell infiltrations in murine models of traumatic injury and autoimmunity [60, 61]. Genetic loss of GFAP expression also increases pathogen burden in various infections, including Staphylococcus aureus and Toxoplasma gondii [89]. Multiple studies have shown that the restriction of leukocyte entry and migration after infection, autoimmune inflammation, and traumatic brain injury is mediated by astrocyte anti-inflammatory functions via the JAK2-STAT3 signaling pathway in GFAP+ cells [25, 38, 39]. The genetic deletion of astrocyte derived STAT3 signaling prevents scar formation and limits immune cell infiltration in a spinal cord injury model [39]. These observations suggest that the astrocyte scar serves as a functional barrier to restrict cytotoxic inflammatory molecules and cells.

Studies genetically deleting essential components of the ECM, such as MMP9, or inhibiting signaling pathways, including Rho/ROCK, to block CSPG activity have shown astrogliosis to exacerbate inflammation after traumatic injury or autoimmune inflammation as well as preventing axonal growth and behavioral recovery [81, 90, 91, 92]. The astrocyte scar has also been shown to exhibit a diverse array of molecules known to prevent axonal growth, such as CSPGs, semaphoring 3A, keratan sulfate proteoglycans (KSPGs) and ephrins/Eph receptors [19, 93, 94]. The complexity of astrocytes in producing, recruiting and restricting inflammatory cells and other molecules have made these cells a difficult target for potential therapeutic manipulation.

Advertisement

5. Astrocytes as a regulator of the innate immune response

After CNS injury or infection, reactive astrocytes release molecules that attract, recruit and facilitate the migration of immune cells to the lesion site (Figure 2). Astrocytes express leukocyte adhesion molecules, including vascular cell adhesion and intercellular adhesion molecules, in models of ischemia, autoimmunity, and infection [30, 33, 95]. Specifically, in an ischemia model, astrocytes release NF-κB, which increases both vascular cell adhesion and intercellular adhesion molecules [33, 96]. These adhesion molecules promote intercellular interactions that contribute to the trafficking of immune cells to the lesion site.

Figure 2.

Reactive astrocytes regulate immune responses. Activation of astrocytes, either directly by purinergic receptors or PRRs or indirectly, promote expression of inflammatory cytokines and chemokines as well as growth factors. The astrocyte response to molecular changes in their microenvironment shapes the recruitment and activation of peripheral immune cells, modulation of the BBB, and cell-cell interaction (not shown). ATP = adenosine triphosphate, BAFF = B-cell activating factor, BBB = blood-brain barrier, CCL = chemokine (C-C motif), CXCL = chemokine (C-X-C motif) ligand, DAMPs = damage-associated molecular pattern, IL = interleukin, IFN = interferon, PAMP = pathogen-associated molecular pattern, PRR = pattern recognition receptor, TNF = tumor necrosis factor.

Like microglia, the resident macrophages of the CNS, astrocytes play a role in innate immune responses by producing cytokines and chemokines, such as type I and II IFNs and TNF, that promote the expression of hundreds of interferon-stimulated genes (ISGs), such as those that participate in inflammatory cell infiltration [97, 98]. Microglia also upregulate the expression of numerous receptors and produce various chemokines after CNS injury, such as chemokine (C-X3-C motif) receptor 1 (CX3CR1) and chemokine (C-C motif) receptor 2 (CCR2) [99]. Similarly, reactive astrocytes also express many of these receptors and chemokines, suggesting that astrocytes and microglia communicate via chemokines. In fact, astrocyte release of chemokines has been shown to be important for attracting peripheral and CNS myeloid cells to the lesion site. In models of traumatic injury and parasitic infection, astrocytes are a source of chemokine (C-C motif) ligand 2 (CCL2) [100, 101]. Astrocytes have also been shown to produce chemokine (C-X3-C motif) ligand 1 (CXC3L1) and CXCL1, detected by monocytes and microglia, in response to viral infection and spinal cord injury, respectively [102, 103, 104].

After entry into the brain, or activation within the brain, innate immune cells demonstrate a spectrum of phenotypes, ranging from pro- and anti-inflammatory states, and can express a variety of cytokines and chemokines, including IL-1β, IFN-γ, and TNF, that contribute to neuroinflammation [105]. Reactive astrocytes have a demonstrated role in modulating immune responses by releasing cytokines that stimulate microglia and macrophages to adopt either pro- or anti-inflammatory responses. For example, after injury or infection, astrocytes have been shown to release cytokines, such as IFN-γ, TNF, and IL-12, that shift microglia and macrophages to a more pro-inflammatory phenotype [106, 107]. Under similar conditions, however, astrocytes have also been observed to produced cytokines, including IL-10 and transforming growth factor beta (TGF-β), which can shift monocytes towards a less inflammatory states [108, 109, 110]. These findings support the notion that astrocyte responses may be context dependent.

Advertisement

6. Astrocytes as a regulator of the adaptive immune response

During the adaptive immune response, astrocytes are a major source of T and B cells chemoattractants. Reactive astrocytes express CCL5 as well as CXCL10 in infection models, both chemoattractants of T cells [111, 112, 113]. In viral infection models, CXCL10 has been shown to be an important ligand for CXCR3 on CD8+ T cells [114]. The recruitment of such CXCR3+ T cells results in improved viral control and survival after infection [115]. In brain samples from patients with multiple sclerosis, astrocytes have been shown to express CXCL12, a T cell chemoattractant, and B-cell activating factor (BAFF), a B cell chemoattractant [116, 117]. Like their influence on microglia and macrophages, cytokines released by reactive astrocytes can shift T cells to adopt either a more beneficial or detrimental phenotype. For example, reactive astrocytes during autoimmunity release pro-inflammatory cytokines, including TNF, IFN-γ, and IL-17, which may induce T cells to adopt a more pro-inflammatory state. However, astrocytes have also been shown to release IL-10 that shifts T cells towards the anti-inflammatory spectrum [23, 24, 32, 118, 119]. Similarly, in a murine spinal cord injury model reactive astrocytes have been shown to release anti-inflammatory TGF-β [31]. Further studies examining the influences of reactive astrocytes on T cells are needed to better understand the long-term effects of astrogliosis on adaptive immune cells during CNS recovery after injury.

Advertisement

7. Reactive astrocytes as a pro-inflammatory regulator

Reactive astrocytes can release a variety of molecular signals that contribute to the inflammatory state of the CNS after injury or disease by directly activating immune defenses with the release of cytokines, chemokines, and other growth factors (Table 2). Recent advancements in astrocyte transcriptome analysis have begun to reveal the context specific production of pro-inflammatory molecules by astrocytes as well as molecular triggers that induce their production. Analysis of the astrocyte transcriptome after in vivo exposure to lipopolysaccharide (LPS) or infection significantly promoted the production of a pro-inflammatory, neurotoxic molecular profile [26, 52]. However, the astrocyte transcriptome shifts towards an anti-inflammatory, neuroprotective profile in an in vivo ischemia model [52]. Future studies should utilize single-cell sequencing techniques to transcriptionally define individual astrocyte responses during health and disease.

TypeAstrocyte moleculeImmune outcomeReferences
Anti-inflammatory function
CytokinesIL-6 and IL-10Activation of numerous anti-inflammatory signaling pathways[95, 120]
Growth factorsTGF-βActivation of numerous anti-inflammatory signaling pathways[121, 122]
SHHMaintenance of BBB, increase of astrocyte proliferation[36, 123]
Intracellular signaling moleculesSTAT3Suppression of multiple pro-inflammatory signaling pathways[37, 38, 39]
ReceptorsERα and Gp130Suppression of multiple pro-inflammatory signaling pathways[23, 25]
Pro-inflammatory function
ChemokinesCCL2 and CCL5Increase recruitment of leukocytes[95, 120, 122, 124, 125]
CXCL1, CXCL10, and CXCL12Increase recruitment of leukocytes[95, 120, 122, 126]
CytokinesIL-6, IL-17, IL-1β, and TNFActivation of numerous pro-inflammatory signaling pathways[95, 120, 121, 122]
Growth factorsVEGFIncrease of BBB permeability, increase of leukocyte infiltration[127, 128]
Intracellular signaling moleculesNFκB and Soc3Activation of numerous pro-inflammatory signaling pathways[31, 32, 37]
Act1Activation of IL-17-mediated pro-inflammatory signaling pathways[28]

Table 2.

Pro- and anti-immune responses of astrocyte molecules.

As immune-competent cells, astrocytes can detect injury signals and respond with the release of cytokine, chemokines, and growth factors as well initiating intracellular signaling pathways. Data suggests that the downstream effects of astrocyte activation is context- and time-dependent, and that both factors as well as other microenvironment stimuli can shift reactive astrocytes to either a more beneficial or detrimental phenotype.

Act = actin, CCL = chemokine (C-C motif), CXCL = chemokine (C-X-C motif) ligand, Gp = glycoprotein, IL = interleukin, IFN = interferon, NFκB = nuclear factor kappa B, EAE = experimental autoimmune encephalomyelitis, SHH = sonic hedgehog, Soc3 = suppressor of cytokine signaling 3, STAT3 = signaling transducer and activator of transcription 3, TGF = transforming growth factor β, TNF = tumor necrosis factor, VEGF = vascular endothelial growth factor.

Despite the number of astrocyte transcriptome data available, few studies have attempted to elucidate mechanisms and signaling cascades that mediate astrocyte pro-inflammatory production. Recent studies have indicated NFκB and SOC3 as transcriptional regulators of pro-inflammatory astrocytes after a traumatic brain injury and during autoimmune inflammation [31, 32, 37]. In a model of autoimmunity, genetic deletion of astrocyte derived NFκB results in increased expression of ECM components and pro-inflammatory cytokines [129]. Astrocytes have also been shown to release CCL2 and CXCL10 to recruit perivascular leukocytes during autoimmune inflammation [124, 125, 126]. While the role of CCL2 and CXCL10 is diverse, evidence suggests that these molecules produced by astrocytes promote leukocyte migration in the CNS parenchyma [124]. In an autoimmune inflammation model, IL-17 inflammatory induction has been shown to be mediated by astrocyte Act1 signaling. Genetically deleting Act1/IL-17 signaling from astrocytes in an EAE model prevents the induction of pro-inflammatory cytokines [28]. Reactive astrocytes can also shift towards a more pro-inflammatory state by overexpressing pro-inflammatory cytokines. In spinal cord injury and autoimmune models, the overexpression of IL-6 in astrocytes leads to increased immune cell infiltration. The proinflammatory cytokine, IL-1β, produced by astrocytes, has also been shown to initiate a signaling cascade that releases vasoactive endothelial growth factor (VEGF), leading to increased BBB permeability and leukocyte leakage [127, 128]. In general, there is also evidence that astrocytes contribute to triggering inflammatory responses due to increases in neuronal activity in epilepsy, neuropathic pain, and stress [130].

Advertisement

8. Reactive astrocytes as an anti-inflammatory regulator

Despite the growing body of work that suggests pro-inflammatory roles for astrocytes, there is an equal amount of evidence suggesting these cells limit inflammation. Recent loss-of-function experiments have also revealed essential anti-inflammatory roles of astrocytes after a variety of CNS injury and disease states (Table 2). These studies have also revealed specific molecular mechanisms that mediate these anti-inflammatory roles. The astrocyte TGF-β response seems to selectively affect astrocyte cytokine and chemokine production after ischemia in murine models. The genetic deletion of TGF-β signaling in astrocytes leads to diffused inflammation and enhances myeloid cell activation [43, 44]. After toxoplasmic encephalitis, the genetic loss of astrocyte TGF-β signaling can lead to the increase of infiltrating T cells. Notably, in both examples, astrocyte TGF-β signaling controls infiltration immune cell number but not necessarily the immune response profile. Astrocyte signaling involving gp130, a receptor for IL-6, or estrogen receptor 1α has also been shown to be anti-inflammatory. In autoimmune and infection models, the genetic deletion of gp130 from astrocytes results in increased inflammatory cytokine production [24, 25]. Similar outcomes, such as increased myeloid infiltration and mortality, are observed in autoimmune models when estrogen receptor 1α is conditional deleted from astrocytes [23]. During autoimmunity, mice deficient in functional IFNγ signaling in astrocytes result in exacerbated disease and mortality due to enhanced leukocyte infiltration and an upregulation of inflammatory gene expression, including CCL1, CCL5, CXC10, and TNF [119]. These mice also had a reduction in anti-inflammatory cytokines, such as IL-10 and IL-27, when compared to mice with functional IFNγ in astrocytes [119].

Advertisement

9. Reactive astrocytes as a neuroprotector of the CNS

In addition to astrocyte regulation of the immune response, these glial cells can respond to CNS injury by altering neuronal function or survival. Neuronal insults result in the release of numerous signals, including increased glutamate production, ATP release and vascular damage. During numerous CNS disease states, including stroke, traumatic injury, epilepsy, neurodegeneration, and viral infection, injured and dying neurons release glutamate, which is harmful to neurons [131, 132, 133, 134]. Astrocytes have been shown to take up excessive extracellular glutamate and dampen the neurotransmitter’s excitotoxicity on neurons, resulting in decreased neuronal death [135]. In vitro studies have also shown that glutamate signaling in astrocytes decrease their production of CCL5, a T cell chemoattractant, reducing overall neuroinflammation [136].

Advertisement

10. Reactive astrocytes as a neurotoxin of the CNS

Inflammation itself can unfortunately impair astrocyte uptake of glutamate, which leads to increased neuronal toxicity and a positive feedback of neuroinflammation [137]; for example, in an in vitro study TNF, released by microglia, signals to astrocyte to release glutamate, increasing excitotoxicity [138]. Neuronal injury and death also lead to the release of potassium and ATP. Both potassium and ATP can activate the inflammasome complex, which is an innate immune mechanism that when activated, resulting in the production of proinflammatory cytokines and increased inflammatory responses. The activation of the inflammasome complex, in this case, is through pannexin 1 channels, expressed by astrocytes [139, 140]. Pannexin 1 channels are opened by potassium and ATP, and once opened, activate the inflammasome complex, leading to the increased production of pro-inflammatory mediators, such as IL-1β, reactive oxygen and nitrogen species, and CCL2, a myeloid cell chemoattractant [139, 141, 142, 143]. ATP also can induce the release of glutamate from astrocytes, which can contribute to overall excitotoxicity [144]. During health, astrocytes release stored glycogen which is converted to lactate and transported to metabolically support neurons [145]. Neurons can resist excitotoxicity when astrocytes increase their glycogen uptake and lactate delivery [146]. Pro-inflammatory cytokines, including IL-1β as well as IFN-γ, TNF, and IL-6, negatively impacts this process by reducing glycogen storage and lactate transport in astrocytes that is necessary as an energy source of neurons [147, 148].

11. Conclusions

Despite the recent advances in defining the role of astrocytes in regulating neuroinflammation, our understanding of these complex glial cells is only beginning. A few studies have demonstrated astrocyte polarization after various CNS injuries [26, 52, 95]. In this model, “A1” reactive astrocytes are pro-inflammatory, neurotoxic while “A2” reactive astrocytes are anti-inflammatory, neuroprotective. Future research, however, is needed to determine whether, like the inflammatory microglia and macrophages, reactive astrocytes shift phenotypes along a spectrum of responses. The amount of new technology available to researchers will also make it possible to further dissect the complexity of astrocytes. Single-cell transcriptional profiling techniques, specifically, can be used as a tool to identify astrocyte subtypes as well as intracellular signaling networks. This method has already been utilized to reveal distinct astrocyte types with regionally restricted distribution in the healthy mouse brain [149]. A key goal, however, for researchers in the future will be to elucidate signaling networks that are relevant to CNS injury and disease and how immune pathways influence astrocyte reactivity.

While current research focuses primarily on astrocyte interactions with other CNS cell types, such as neurons, microglia, pathogens and infiltrating immune cells, future studies will need to examine how other biologic variables, including age and sex, influence astrocyte effects within the central and peripheral immune systems. Additionally, there is already some evidence that astrocyte immune regulation is influenced by the gut microbiome [150], but the implications and effects of this process on health and disease are unknown.

In summary, astrocytes exhibit diverse and sometimes conflicting roles in the setting of neuroinflammatory diseases. These multipurpose glia cells not only sense and influence damaged neurons but appear to summate multiple signals to develop specific responses that modulate neuroinflammation. It is our hope that understanding how astrocytes receive and response to information as they perform these differential roles will lead to therapies that specifically target astrocytes during CNS injury and disease.

Abbreviations

ATPadenosine triphosphate
BAFFB-cell activating factor
BBBblood-brain barrier
cAMPcyclic AMP
CCLchemokine (C-C motif) ligand
CCRchemokine (C-C motif) receptor
CNScentral nervous system
CSPGchondroitin sulfate proteoglycan
CXCLchemokine (C-X-C motif) ligand
CXCRchemokine (C-X-C motif) receptor
DAMPdamage-associated molecular pattern
ECMextracellular matrix
EGFepidermal growth factor
FGFfibroblast growth factor
GFAPglial fibrillary acid protein
IFNinterferon
ILinterleukin
ISGinterferon-stimulated gene
KSPGkeratan sulfate proteoglycan
MMP9matric metallopeptidase 9
NFκBnuclear factor kappa B
NPCneural progenitor cell
PAMPpathogen-associated molecular pattern
PPRpattern recognition receptor
RIG-Iretinoic acid-inducible gene I
RLRsRIG-I-like receptor
SGZsubgranular zone
SHHsonic hedge hog
SOC3suppressor of cytokine signaling 3
STAT3signal transducer and activator of transcription 3
SVZsubventricular zone
TGF-βtransforming growth factor beta
TLRtoll-like receptor
TNFtumor necrosis factor
VEGFvasoactive endothelial growth factor

References

  1. 1. Klein RS, Fricker LD. Heterogeneous expression of car ypeptidase E and proenkephalin mRNAs by cultured astrocytes. Brain Research. 1992;569(2):300-310
  2. 2. Bindocci E, Savtchouk I, Liaudet N, Becker D, Carriero G, Volterra A. Neuroscience: Three-dimensional Ca2+ imaging advances understanding of astrocyte biology. Science. 2017;356(6339):eaai8185
  3. 3. Yoon H, Walters G, Paulsen AR, Scarisbrick IA. Astrocyte heterogeneity across the brain and spinal cord occurs developmentally, in adulthood and in response to demyelination. PLoS One. 2017;12(7):e0180697
  4. 4. Bialas AR, Stevens B. TGF-β signaling regulates neuronal C1q expression and developmental synaptic refinement. Nature Neuroscience. 2013;16(12):1773-1782
  5. 5. Hurwitz AA, Berman JW, Rashbaum WK, Lyman WD. Human fetal astrocytes induce the expression of blood-brain barrier specific proteins by autologous endothelial cells. Brain Research. 1993;625(2):238-243
  6. 6. Barres BA. The mystery and magic of glia: A perspective on their roles in health and disease. Neuron. 2008;60(3):430-440
  7. 7. Sofroniew MV, Vinters HV. Astrocytes: Biology and pathology. Acta Neuropathologica. 2010;119(1):7-35
  8. 8. Liddelow SA, Barres BA. Reactive astrocytes: Production, function, and therapeutic potential. Immunity. 2017;46(6):957-967
  9. 9. Sofroniew MV. Molecular dissection of reactive astrogliosis and glial scar formation. Trends in Neurosciences. 2009;32(12):638-647
  10. 10. Sofroniew MV. Astrogliosis. Cold Spring Harbor Perspectives in Biology. 2015;7(2):a020420
  11. 11. Pekny M, Nilsson M. Astrocyte activation and reactive gliosis. Glia. 2005;50(4):427-434
  12. 12. Kang W, Hébert JM. Signaling pathways in reactive astrocytes, a genetic perspective. Molecular Neurobiology. 2011;43(3):147-154
  13. 13. Anderson MA, Ao Y, Sofroniew MV. Heterogeneity of reactive astrocytes. Neuroscience Letters. 2014;565:23-29
  14. 14. Pekny M et al. Abnormal reaction to central nervous system injury in mice lacking glial fibrillary acidic protein and vimentin. The Journal of Cell Biology. 1999;145(3):503-514
  15. 15. Sofroniew MV. Multiple roles for astrocytes as effectors of cytokines and inflammatory mediators. The Neuroscientist. 2014;20(2):160-172
  16. 16. Pekny M et al. Astrocytes: A central element in neurological diseases. Acta Neuropathologica. 2016;131(3):323-345
  17. 17. Burda JE, Sofroniew MV. Reactive gliosis and the multicellular response to CNS damage and disease. Neuron. 2014;81(2):229-248
  18. 18. Sofroniew MV. Reactive astrocytes in neural repair and protection. Neuroscience. 2005;11(5):400-407
  19. 19. Silver J, Miller JH. Regeneration beyond the glial scar. Nature Reviews. Neuroscience. 2004;5(2):146-156
  20. 20. Silver J, Schwab ME, Popovich PG. Central nervous system regenerative failure: Role of oligodendrocytes, astrocytes, and microglia. Cold Spring Harbor Perspectives in Biology. 2015;7(3):a020602
  21. 21. Yiu G, He Z. Glial inhibition of CNS axon regeneration. Nature Reviews Neuroscience. 2006;7(8):617-627
  22. 22. Hamby ME, Sofroniew MV. Reactive astrocytes as therapeutic targets for CNS disorders. Neurotherapeutics : The Journal of the American Society for Experimental NeuroTherapeutics. 2010;7(4):494-506
  23. 23. Spence RD et al. Neuroprotection mediated through estrogen receptor-α in astrocytes. Proceedings of the National Academy of Sciences of the United States of America. 2011;108(21):8867-8872
  24. 24. Haroon F et al. Gp130-dependent astrocytic survival is critical for the control of autoimmune central nervous system inflammation. Journal of Immunology. 2011;186(11):6521-6531
  25. 25. Drögemüller K et al. Astrocyte gp130-expression is critical for the control of toxoplasma encephalitis. BMC Proceedings. 2008;2(Suppl 1):S10
  26. 26. Garber C, Vasek MJ, Vollmer LL, Sun T, Jiang X, Klein RS. Astrocytes decrease adult neurogenesis during virus-induced memory dysfunction via IL-1 article. Nature Immunology. 2018;19(2):151-161
  27. 27. Herx LM, Yong VW. Interleukin-1β is required for the early evolution of reactive astrogliosis following CNS lesion. Journal of Neuropathology and Experimental Neurology. 2001;60(10):961-971
  28. 28. Kang Z et al. Astrocyte-restricted ablation of interleukin-17-induced act1-mediated signaling ameliorates autoimmune encephalomyelitis. Immunity. 2010;32(3):414-425
  29. 29. Balasingam V, Tejada-Berges T, Wright E, Bouckova R, Yong V. Reactive astrogliosis in the neonatal mouse brain and its modulation by cytokines. The Journal of Neuroscience. 1994;14(2):846-856
  30. 30. Gimenez MAT, Sim JE, Russell JH. TNFR1-dependent VCAM-1 expression by astrocytes exposes the CNS to destructive inflammation. Journal of Neuroimmunology. 2004;151(1):116-125
  31. 31. Brambilla R et al. Transgenic inhibition of astroglial NF-κB improves functional outcome in experimental autoimmune encephalomyelitis by suppressing chronic central nervous system inflammation. Journal of Immunology. 2009;182(5):2628-2640
  32. 32. Brambilla R et al. Inhibition of astroglial nuclear factor κB reduces inflammation and improves functional recovery after spinal cord injury. The Journal of Experimental Medicine. 2005;202(1):145-156
  33. 33. Dvoriantchikova G et al. Inactivation of astroglial NF-κB promotes survival of retinal neurons following ischemic injury. The European Journal of Neuroscience. 2009;30(2):175-185
  34. 34. Shimada IS, Borders A, Aronshtam A, Spees JL. Proliferating reactive astrocytes are regulated by notch-1 in the peri-infarct area after stroke. Stroke. 2011;42(11):3231-3237
  35. 35. Shimada IS, LeComte MD, Granger JC, Quinlan NJ, Spees JL. Self-renewal and differentiation of reactive astrocyte-derived neural stem/progenitor cells isolated from the cortical peri-infarct area after stroke. The Journal of Neuroscience. 2012;32(23):7926-7940
  36. 36. Alvarez JI et al. The hedgehog pathway promotes blood-brain barrier integrity and CNS immune quiescence. Science. 2011;334(6063):1727-1731
  37. 37. Okada S et al. Conditional ablation of Stat3 or Socs3 discloses a dual role for reactive astrocytes after spinal cord injury. Nature Medicine. 2006;12(7):829-834
  38. 38. Wanner IB et al. Glial scar borders are formed by newly proliferated, elongated astrocytes that interact to corral inflammatory and fibrotic cells via STAT3-dependent mechanisms after spinal cord injury. The Journal of Neuroscience. 2013;33(31):12870-12886
  39. 39. Herrmann JE et al. STAT3 is a critical regulator of astrogliosis and scar formation after spinal cord injury. The Journal of Neuroscience. 2008;28(28):7231-7243
  40. 40. Anderson MA et al. Astrocyte scar formation AIDS central nervous system axon regeneration. Nature. 2016;532(7598):195-200
  41. 41. Schachtrup C et al. Fibrinogen triggers astrocyte scar formation by promoting the availability of active TGF-β after vascular damage. The Journal of Neuroscience. 2010;30(17):5843-5854
  42. 42. Wang Y, Moges H, Bharucha Y, Symes A. Smad3 null mice display more rapid wound closure and reduced scar formation after a stab wound to the cerebral cortex. Experimental Neurology. 2007;203(1):168-184
  43. 43. Cekanaviciute E, Fathali N, Doyle KP, Williams AM, Han J, Buckwalter MS. Astrocytic transforming growth factor-beta signaling reduces subacute neuroinflammation after stroke in mice. Glia. 2014;62(8):1227-1240
  44. 44. Cekanaviciute E et al. Astrocytic TGF-β signaling limits inflammation and reduces neuronal damage during central nervous system toxoplasma infection. Journal of Immunology. 2014;193(1):139-149
  45. 45. Pekny M, Pekna M. Astrocyte reactivity and reactive astrogliosis: Costs and benefits. Physiological Reviews. 2014;94(4):1077-1098
  46. 46. Messing A, Head MW, Galles K, Galbreath EJ, Goldman JE, Brenner M. Fatal encephalopathy with astrocyte inclusions in GFAP transgenic mice. The American Journal of Pathology. 1998;152(2):391-398
  47. 47. Gao K et al. Traumatic scratch injury in astrocytes triggers calcium influx to activate the JNK/c-Jun/AP-1 pathway and switch on GFAP expression. Glia. 2013;61(12):2063-2077
  48. 48. Gadea A, Schinelli S, Gallo V. Endothelin-1 regulates astrocyte proliferation and reactive gliosis via a JNK/c-Jun signaling pathway. The Journal of Neuroscience. 2008;28(10):2394-2408
  49. 49. Levison SW, Jiang FJ, Stoltzfus OK, Ducceschi MH. IL-6-type cytokines enhance epidermal growth factor-stimulated astrocyte proliferation. Glia. 2000;32(3):328-337
  50. 50. Sirko S et al. Reactive glia in the injured brain acquire stem cell properties in response to sonic hedgehog. Cell Stem Cell. 2013;12(4):426-439
  51. 51. Estrada-Sánchez AM, Rebec GV. Corticostriatal dysfunction and glutamate transporter 1 (GLT1) in Huntington’s disease: Interactions between neurons and astrocytes. Basal Ganglia. 2012;2(2):57-66
  52. 52. Liddelow SA et al. Neurotoxic reactive astrocytes are induced by activated microglia. Nature. 2017;541(7638):481-487
  53. 53. Röhl C, Lucius R, Sievers J. The effect of activated microglia on astrogliosis parameters in astrocyte cultures. Brain Research. 2007;1129:43-52
  54. 54. Bardehle S et al. Live imaging of astrocyte responses to acute injury reveals selective juxtavascular proliferation. Nature Neuroscience. 2013;16(5):580-586
  55. 55. Wilhelmsson U et al. Redefining the concept of reactive astrocytes as cells that remain within their unique domains upon reaction to injury. Proceedings of the National Academy of Sciences of the United States of America. 2006;103(46):17513-17518
  56. 56. Ge W-P, Jia J-M. Local production of astrocytes in the cerebral cortex. Neuroscience. 2016;323:3-9
  57. 57. Ransohoff RM, Engelhardt B. The anatomical and cellular basis of immune surveillance in the central nervous system. Nature Reviews. Immunology. 2012;12(9):623-635
  58. 58. Owens T, Bechmann I, Engelhardt B. Perivascular spaces and the two steps to neuroinflammation. Journal of Neuropathology and Experimental Neurology. 2008;67(12):1113-1121
  59. 59. Wilson EH, Weninger W, Hunter CA. Trafficking of immune cells in the central nervous system. The Journal of Clinical Investigation. 2010;120(5):1368-1379
  60. 60. Voskuhl RR et al. Reactive astrocytes form scar-like perivascular barriers to leukocytes during adaptive immune inflammation of the CNS. The Journal of Neuroscience. 2009;29(37):11511-11522
  61. 61. Bush TG et al. Leukocyte infiltration, neuronal degeneration, and neurite outgrowth after ablation of scar-forming, reactive astrocytes in adult transgenic mice. Neuron. 1999;23(2):297-308
  62. 62. Faulkner JR, Herrmann JE, Woo MJ, Tansey KE, Doan NB, Sofroniew MV. Reactive astrocytes protect tissue and preserve function after spinal cord injury. The Journal of Neuroscience. 2004;24(9):2143-2155
  63. 63. Li L et al. Protective role of reactive astrocytes in brain ischemia. Journal of Cerebral Blood Flow and Metabolism. 2008;28(3):468-481
  64. 64. Macauley SL, Pekny M, Sands MS. The role of attenuated astrocyte activation in infantile neuronal ceroid lipofuscinosis. The Journal of Neuroscience. 2011;31(43):15575-15585
  65. 65. Myer DJ, Gurkoff GG, Lee SM, Hovda DA, Sofroniew MV. Essential protective roles of reactive astrocytes in traumatic brain injury. Brain. 2006;129(10):2761-2772
  66. 66. Nawashiro H, Messing A, Azzam N, Brenner M. Mice lacking GFAP are hypersensitive to traumatic cerebrospinal injury. Neuroreport. 1998;9(8):1691-1696
  67. 67. Romão LF, De Sousa VO, Neto VM, Gomes FCA. Glutamate activates GFAP gene promoter from cultured astrocytes through TGF-β1 pathways. Journal of Neurochemistry. 2008;106(2):746-756
  68. 68. Huang XJ et al. Activation of CysLT receptors induces astrocyte proliferation and death after oxygen-glucose deprivation. Glia. 2008;56(1):27-37
  69. 69. Neary JT, Kang Y, Willoughby KA, Ellis EF. Activation of extracellular signal-regulated kinase by stretch-induced injury in astrocytes involves extracellular ATP and P2 purinergic receptors. The Journal of Neuroscience. 2003;23(6):2348-2356
  70. 70. Wurm A et al. Purinergic signaling involved in Müller cell function in the mammalian retina. Progress in Retinal and Eye Research. 2011;30(5):324-342
  71. 71. Crill EK, Furr-Rogers SR, Marriott I. RIG-I is required for VSV-induced cytokine production by murine glia and acts in combination with DAI to initiate responses to HSV-1. Glia. 2015;63(12):2168-2180
  72. 72. Pfefferkorn C et al. Abortively infected astrocytes appear to represent the main source of interferon beta in the virus-infected brain. Journal of Virology. 2016;90(4):2031-2038
  73. 73. Bianchi MG, Bardelli D, Chiu M, Bussolati O. Changes in the expression of the glutamate transporter EAAT3/EAAC1 in health and disease. Cellular and Molecular Life Sciences. 2014;71(11):2001-2015
  74. 74. Kono H, Rock KL. How dying cells alert the immune system to danger. Nature Reviews. Immunology. 2008;8(4):279-289
  75. 75. Daniels BP, Holman DW, Cruz-Orengo L, Jujjavarapu H, Durrant DM, Klein RS. Viral pathogen-associated molecular patterns regulate blood-brain barrier integrity via competing innate cytokine signals. MBio. 2014;5(5):e01476-14
  76. 76. Wong D, Dorovini-Zis K, Vincent SR. Cytokines, nitric oxide, and cGMP modulate the permeability of an in vitro model of the human blood-brain barrier. Experimental Neurology. 2004;190(2):446-455
  77. 77. Daniels BP et al. Regional astrocyte IFN signaling restricts pathogenesis during neurotropic viral infection. The Journal of Clinical Investigation. 2017;127(3):843-856
  78. 78. Brosnan CF, Raine CS. The astrocyte in multiple sclerosis revisited. Glia. 2013;61(4):453-465
  79. 79. Goverman J. Autoimmune T cell responses in the central nervous system. Nature Reviews. Immunology. 2009;9(6):393-407
  80. 80. Prinz M, Priller J. Microglia and brain macrophages in the molecular age: From origin to neuropsychiatric disease. Nature Reviews. Neuroscience. 2014;15(5):300-312
  81. 81. Hsu JYC et al. Matrix metalloproteinase-9 facilitates glial scar formation in the injured spinal cord. The Journal of Neuroscience. 2008;28(50):13467-13477
  82. 82. Jones LL, Margolis RU, Tuszynski MH. The chondroitin sulfate proteoglycans neurocan, brevican, phosphacan, and versican are differentially regulated following spinal cord injury. Experimental Neurology. 2003;182(2):399-411
  83. 83. Davies SJA, Fitch MT, Memberg SP, Hall AK, Raisman G, Silver J. Regeneration of adult axons in white matter tracts of the central nervous system. Nature. 1997;390(6661):680-683
  84. 84. McKeon RJ, Jurynec MJ, Buck CR. The chondroitin sulfate proteoglycans neurocan and phosphacan are expressed by reactive astrocytes in the chronic CNS glial scar. The Journal of Neuroscience. 1999;19(24):10778-10788
  85. 85. Middeldorp J, Hol EM. GFAP in health and disease. Progress in Neurobiology. 2011;93(3):421-443
  86. 86. Liedtke W, Edelmann W, Chiu FC, Kucherlapati R, Raine CS. Experimental autoimmune encephalomyelitis in mice lacking glial fibrillary acidic protein is characterized by a more severe clinical course and an infiltrative central nervous system lesion. The American Journal of Pathology. 1998;152(1):251-259
  87. 87. Liu Z et al. Beneficial effects of gfap/vimentin reactive astrocytes for axonal remodeling and motor behavioral recovery in mice after stroke. Glia. 2014;62(12):2022-2033
  88. 88. Kraft AW et al. Attenuating astrocyte activation accelerates plaque pathogenesis in APP/PS1 mice. The FASEB Journal. 2012;27(1):187-198
  89. 89. Stenzel W, Soltek S, Schlüter D, Deckert M. The intermediate filament GFAP is important for the control of experimental murine Staphylococcus aureus-induced brain abscess and toxoplasma encephalitis. Journal of Neuropathology and Experimental Neurology. 2004;63(6):631-640
  90. 90. Monnier PP, Sierra A, Schwab JM, Henke-Fahle S, Mueller BK. The Rho/ROCK pathway mediates neurite growth-inhibitory activity associated with the chondroitin sulfate proteoglycans of the CNS glial scar. Molecular and Cellular Neurosciences. 2003;22(3):319-330
  91. 91. Gilbert RJ, McKeon RJ, Darr A, Calabro A, Hascall VC, Bellamkonda RV. CS-4,6 is differentially upregulated in glial scar and is a potent inhibitor of neurite extension. Molecular and Cellular Neurosciences. 2005;29(4):545-558
  92. 92. Gates MA, Fillmore H, Steindler DA. Chondroitin sulfate proteoglycan and tenascin in the wounded adult mouse neostriatum in vitro: Dopamine neuron attachment and process outgrowth. The Journal of Neuroscience. 1996;16(24):8005-8018
  93. 93. Ohtake Y, Li S. Molecular mechanisms of scar-sourced axon growth inhibitors. Brain Research. 2015;1619:22-35
  94. 94. Siebert JR, Conta Steencken A, Osterhout DJ. Chondroitin sulfate proteoglycans in the nervous system: Inhibitors to repair. BioMed Research International. 2014;2014:845323
  95. 95. Zamanian JL et al. Genomic analysis of reactive astrogliosis. The Journal of Neuroscience. 2012;32(18):6391-6410
  96. 96. Ouali Alami N et al. NF-κB activation in astrocytes drives a stage-specific beneficial neuroimmunological response in ALS. The EMBO Journal. 2018;37(16):e98697
  97. 97. Cui J, Chen Y, Wang HY, Wang R-F. Mechanisms and pathways of innate immune activation and regulation in health and cancer. Human Vaccines & Immunotherapeutics. 2015;10(11):3270-3285
  98. 98. Hu X, Chakravarty SD, Ivashkiv LB. Regulation of interferon and Toll-like receptor signaling during macrophage activation by opposing feedforward and feedback inhibition mechanisms. Immunological Reviews. 2008;226:41-56
  99. 99. Cherry JD, Olschowka JA, O’Banion MK. Neuroinflammation and M2 microglia: The good, the bad, and the inflamed. Journal of Neuroinflammation. 2014;11
  100. 100. Glabinski AR et al. Chemokine monocyte chemoattractant protein-1 is expressed by astrocytes after mechanical injury to the brain. Journal of Immunology. 1996;156(11):4363-4368
  101. 101. Strack A, Asensio VC, Campbell IL, Schlüter D, Deckert M. Chemokines are differentially expressed by astrocytes, microglia and inflammatory leukocytes in toxoplasma encephalitis and critically regulated by interferon-γ. Acta Neuropathologica. 2002;103(5):458-468
  102. 102. Pereira CF, Middel J, Jansen G, Verhoef J, Nottet HSLM. Enhanced expression of fractalkine in HIV-1 associated dementia. Journal of Neuroimmunology. 2001;115(1):168-175
  103. 103. Pineau I, Sun L, Bastien D, Lacroix S. Astrocytes initiate inflammation in the injured mouse spinal cord by promoting the entry of neutrophils and inflammatory monocytes in an IL-1 receptor/MyD88-dependent fashion. Brain, Behavior, and Immunity. 2010;24(4):540-553
  104. 104. Hughes PM, Botham MS, Frentzel S, Mir A, Perry VH. Expression of fractalkine (CX3CL1) and its receptor, CX3CR1, during acute and chronic inflammation in the rodent CNS. Glia. 2002;37:314-327
  105. 105. Kigerl KA, Gensel JC, Ankeny DP, Alexander JK, Donnelly DJ, Popovich PG. Identification of two distinct macrophage subsets with divergent effects causing either neurotoxicity or regeneration in the injured mouse spinal cord. The Journal of Neuroscience. 2009;29(43):13435-13444
  106. 106. Fadok VA, Bratton DL, Konowal A, Freed PW, Westcott JY, Henson PM. Macrophages that have ingested apoptotic cells in vitro inhibit proinflammatory cytokine production through autocrine/paracrine mechanisms involving TGF-beta, PGE2, and PAF. The Journal of Clinical Investigation. 1998;101(4):890-898
  107. 107. Martinez FO, Helming L, Gordon S. Alternative activation of macrophages: An immunologic functional perspective. Annual Review of Immunology. 2009;27(1):451-483
  108. 108. Spera PA, Ellison JA, Feuerstein GZ, Barone FC. IL-10 reduces rat brain injury following focal stroke. Neuroscience Letters. 1998;251(3):189-192
  109. 109. Xiaoxing X, E BG, Lijun X, Bing OY, Xinmin X, G GR. Increased brain injury and worsened neurological outcome in interleukin-4 knockout mice after transient focal cerebral ischemia. Stroke. 2011;42(7):2026-2032
  110. 110. Ruocco A et al. A transforming growth factor-β antagonist unmasks the neuroprotective role of this endogenous cytokine in excitotoxic and ischemic brain injury. Journal of Cerebral Blood Flow and Metabolism. 1999;19(12):1345-1353
  111. 111. Babcock AA, Kuziel WA, Rivest S, Owens T. Chemokine expression by glial cells directs leukocytes to sites of axonal injury in the CNS. The Journal of Neuroscience. 2003;23(21):7922-7930
  112. 112. Sauder C et al. Chemokine gene expression in astrocytes of borna disease virus-infected rats and mice in the absence of inflammation. Journal of Virology. 2000;74(19):9267-9280
  113. 113. Conant K et al. Induction of monocyte chemoattractant protein-1 in HIV-1 Tat-stimulated astrocytes and elevation in AIDS dementia. Proceedings of the National Academy of Sciences. 1998;95(6):3117-3121
  114. 114. Zhang B, Chan YK, Lu B, Diamond MS, Klein RS. CXCR3 mediates region-specific antiviral T cell trafficking within the central nervous system during West Nile virus encephalitis. Journal of Immunology. 2008;180(4):2641-2649
  115. 115. Christensen JE et al. Fulminant lymphocytic choriomeningitis virus-induced inflammation of the CNS involves a cytokine-chemokine-cytokine-chemokine cascade. Journal of Immunology. 2009;182(2):1079-1087
  116. 116. Krumbholz M et al. Chemokines in multiple sclerosis: CXCL12 and CXCL13 up-regulation is differentially linked to CNS immune cell recruitment. Brain. 2006;129(1):200-211
  117. 117. Krumbholz M et al. BAFF is produced by astrocytes and up-regulated in multiple sclerosis lesions and primary central nervous system lymphoma. The Journal of Experimental Medicine. 2005;201(2):195-200
  118. 118. Stüve O et al. The role of the MHC class II transactivator in class II expression and antigen presentation by astrocytes and in susceptibility to central nervous system autoimmune disease. Journal of Immunology. 2002;169(12):6720-6732
  119. 119. Hindinger C et al. IFN-γ signaling to astrocytes protects from autoimmune mediated neurological disability. PLoS One. 2012;7(7):e42088
  120. 120. Hamby ME, Coppola G, Ao Y, Geschwind DH, Khakh BS, Sofroniew MV. Inflammatory mediators alter the astrocyte transcriptome and calcium signaling elicited by multiple G-protein-coupled receptors. The Journal of Neuroscience. 2012;32(42):14489-14510
  121. 121. Meeuwsen S, Persoon-Deen C, Bsibsi M, Ravid R, Van Noort JM. Cytokine, chemokine and growth factor gene profiling of cultured human astrocytes after exposure to proinflammatory stimuli. Glia. 2003;43(3):243-253
  122. 122. Jensen CJ, Massie A, De Keyser J. Immune players in the CNS: The astrocyte. Journal of Neuroimmune Pharmacology. 2013;8(4):824-839
  123. 123. Pitter KL et al. The SHH/Gli pathway is reactivated in reactive glia and drives proliferation in response to neurodegeneration-induced lesions. Glia. 2014;62(10):1595-1607
  124. 124. Kim RY et al. Astrocyte CCL2 sustains immune cell infiltration in chronic experimental autoimmune encephalomyelitis. Journal of Neuroimmunology. 2014;274(1):53-61
  125. 125. Moreno M et al. Conditional ablation of astroglial CCL2 suppresses CNS accumulation of M1 macrophages and preserves axons in mice with MOG peptide EAE. The Journal of Neuroscience. 2014;34(24):8175-8185
  126. 126. Mills Ko E et al. Deletion of astroglial CXCL10 delays clinical onset but does not affect progressive axon loss in a murine autoimmune multiple sclerosis model. Journal of Neuroinflammation. 2014;11(1):105
  127. 127. Argaw AT et al. Astrocyte-derived VEGF-A drives blood-brain barrier disruption in CNS inflammatory disease. The Journal of Clinical Investigation. 2012;122(7):2454-2468
  128. 128. Argaw AT, Gurfein BT, Zhang Y, Zameer A, John GR. VEGF-mediated disruption of endothelial CLN-5 promotes blood-brain barrier breakdown. Proceedings of the National Academy of Sciences. 2009;106(6):1977-1982
  129. 129. Brambilla R, Morton PD, Ashbaugh JJ, Karmally S, Lambertsen KL, Bethea JR. Astrocytes play a key role in EAE pathophysiology by orchestrating in the CNS the inflammatory response of resident and peripheral immune cells and by suppressing remyelination. Glia. 2014;62(3):452-467
  130. 130. Xanthos DN, Sandkühler J. Neurogenic neuroinflammation: Inflammatory CNS reactions in response to neuronal activity. Nature Reviews Neuroscience. 2014;15(1):43-53
  131. 131. Kaul M, Garden GA, Lipton SA. Pathways to neuronal injury and apoptosis in HIV-associated dementia. Nature. 2001;410(6831):988-994
  132. 132. Maragakis NJ, Rothstein JD. Mechanisms of disease: Astrocytes in neurodegenerative disease. Nature Clinical Practice. Neurology. 2006;2(12):679-689
  133. 133. Dong XX, Wang Y, Qin ZH. Molecular mechanisms of excitotoxicity and their relevance to pathogenesis of neurodegenerative diseases. Acta Pharmacologica Sinica. 2009;30(4):379-387
  134. 134. Goodrich GS, Kabakov AY, Hameed MQ , Dhamne SC, Rosenberg PA, Rotenberg A. Ceftriaxone treatment after traumatic brain injury restores expression of the glutamate transporter, GLT-1, reduces regional gliosis, and reduces post-traumatic seizures in the rat. Journal of Neurotrauma. 2013;30(16):1434-1441
  135. 135. Lepore AC et al. Reduction in expression of the astrocyte glutamate transporter, GLT1, worsens functional and histological outcomes following traumatic spinal cord injury. Glia. 2011;59(12):1996-2005
  136. 136. Besong G et al. Activation of group III metabotropic glutamate receptors inhibits the production of RANTES in glial cell cultures. The Journal of Neuroscience. 2002;22(13):5403-5411
  137. 137. Prow NA, Irani DN. The inflammatory cytokine, interleukin-1 beta, mediates loss of astroglial glutamate transport and drives excitotoxic motor neuron injury in the spinal cord during acute viral encephalomyelitis. Journal of Neurochemistry. 2008;105(4):1276-1286
  138. 138. Bezzi P et al. CXCR4-activated astrocyte glutamate release via TNFa: Amplification by microglia triggers neurotoxicity. Nature Neuroscience. 2001;4(7):702-710
  139. 139. Minkiewicz J, de Rivero Vaccari JP, Keane RW. Human astrocytes express a novel NLRP2 inflammasome. Glia. 2013;61(7):1113-1121
  140. 140. Kido Y et al. Regulation of activity of P2X7 receptor by its splice variants in cultured mouse astrocytes. Glia. 2014;62(3):440-451
  141. 141. Pelegrin P, Surprenant A. Pannexin-1 mediates large pore formation and interleukin-1beta release by the ATP-gated P2X7 receptor. The EMBO Journal. 2006;25(21):5071-5082
  142. 142. Silverman WR et al. The pannexin 1 channel activates the inflammasome in neurons and astrocytes. The Journal of Biological Chemistry. 2009;284(27):18143-18151
  143. 143. Xia M, Zhu Y. FOXO3a involvement in the release of TNF-α stimulated by ATP in spinal cord astrocytes. Journal of Molecular Neuroscience. 2013;51(3):792-804
  144. 144. Pascual O, Ben Achour S, Rostaing P, Triller A, Bessis A. Microglia activation triggers astrocyte-mediated modulation of excitatory neurotransmission. Proceedings of the National Academy of Sciences. 2012;109(4):E197-E205
  145. 145. Pellerin L et al. Evidence supporting the existence of an activity-dependent astrocyte-neuron lactate shuttle. Developmental Neuroscience. 1998;20(4-5):291-299
  146. 146. Bliss TM et al. Dual-gene, dual-cell type therapy against an excitotoxic insult by bolstering neuroenergetics. The Journal of Neuroscience. 2004;24(27):6202-6208
  147. 147. Bélanger M, Allaman I, Magistretti PJ. Brain energy metabolism: Focus on astrocyte-neuron metabolic cooperation. Cell Metabolism. 2011;14(6):724-738
  148. 148. Gavillet M, Allaman I, Magistretti PJ. Modulation of astrocytic metabolic phenotype by proinflammatory cytokines. Glia. 2008;56(9):975-989
  149. 149. Zeisel A et al. Molecular architecture of the mouse nervous system resource molecular architecture of the mouse nervous system. Cell. 2018;174:999-1014
  150. 150. Rothhammer V et al. Type i interferons and microbial metabolites of tryptophan modulate astrocyte activity and central nervous system inflammation via the aryl hydrocarbon receptor. Nature Medicine. 2016;22(6):586-597
  151. 151. Engelhardt B, Coisne C. Fluids and barriers of the CNS establish immune privilege by confining immune surveillance to a two-walled castle moat surrounding the CNS castle. Fluids and Barriers of the CNS. 2011;8(1):4

Written By

Allison Soung and Robyn S. Klein

Submitted: 31 August 2019 Reviewed: 16 September 2019 Published: 05 November 2019