Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2018 Nov 6.
Published in final edited form as: Toxicol Pathol. 2010 Nov 12;39(1):115–123. doi: 10.1177/0192623310385254

Role of Astrocytes in Brain Function and Disease

Marta sidoryk-wegrzynowicz 1, Michal Wegrzynowicz 2,3, Eunsook Lee 4, Aaron B Bowman 2,3, Michael Aschner 1,3,4
PMCID: PMC6218934  NIHMSID: NIHMS993709  PMID: 21075920

Abstract

Astrocytes assume multiple roles in maintaining an optimally suited milieu for neuronal function. Select astrocytic functions include the maintenance of redox potential, the production of trophic factors, the regulation of neurotransmitter and ion concentrations, and the removal of toxins and debris from the cerebrospinal fluid (CSF). Impairments in these and other functions, as well as physiological reactions of astrocytes to injury, can trigger or exacerbate neuronal dysfunction. This review addresses select metabolic interactions between neurons and astrocytes and emphasizes the role of astrocytes in mediating and amplifying the progression of several neurodegenerative disorders, such as Parkinson’s disease (PD), hepatic encephalopathy (HE), hyperammonemia (HA), Alzheimer’s disease (AD), and ischemia.

Keywords: neurodegenerative diseases, nervous system, astrocyte

1. Introduction

The past several decades have given rise to many important discoveries and novel insights into the role of astrocytes in normal brain function and disease, firmly establishing concepts that describe the dynamic and reciprocal signaling networks between astrocytes and neurons. This article briefly delineates a select set of astrocytic functions within the mature central nervous system (CNS), followed by a short discussion emphasizing the astrocytic modulation of neurodegenerative injuries, including Parkinson’s disease (PD), Alzheimer’s disease (AD), hepatic encephalopathy (HE), hyperammonemia (HA), and ischemia. For an excellent and thorough review on the various functions of mature astrocytes, the reader is referred to a recent article by Kimelberg (2010). For a review on astrocytic glial toxicants and their role in the pathogenesis of human neurode-generative diseases such as AD and PD, please refer to Aschner and LoPachin (2001).

1A. Select Functions of Mature Astrocytes and Metabolic Relationships to Neurons

The metabolic relationship between neurons and astrocytes is perhaps best exemplified by the neuronal dependence on astrocyte-derived thiols for the maintenance of stable concentrations of glutathione (GSH) (Dringen and Hirrlinger 2003). GSH is synthesized in a two-step process. First, the action of γ-glutamylcysteine (γ-GS) synthase forms γ-glutamylcysteine from cysteine and glutamate; then glycine derived from GSH synthetase is added, completing the process. GSH is a major antioxidant that constitutes ~90% of the intracellular nonprotein thiols. Conjugation with GSH detoxifies and eliminates toxic molecules from cells, and de novo GSH synthesis ensures the maintenance of intracellular redox status. In general, GSH levels are lower in neurons than in astrocytes (Sagara, Miura, and Bannai 1993), thus rendering neurons more susceptible to injury as a result of oxidative stress. Furthermore, cysteine derived from astrocytes is essential for the maintenance of stable GSH levels in neurons (Dringen, Pfeiffer, and Hamprecht 1999; Kaur, Aschner, and Syversen 2007; Shanker et al. 2001; Wang and Cynader 2000).

The metabolic interaction between astrocytes and neurons is also critical for energy metabolism as well as for the synthesis of de novo glutamate and γ-aminobutyric acid (GABA). Furthermore, this important interaction is also responsible for the termination of glutamatergic and GABAergic activity, which is achieved by the re-uptake of both neurotransmitters, especially glutamate, into astrocytes (see below). Optimal synaptic glutamate concentrations are maintained by glutamate aspartate transporter (GLAST) (Storck et al. 1992) and gluta-mate transporter 1 (GLT1) (Lehre et al. 1995), both of which are preferentially localized on astrocytes. This re-uptake ensures low synaptic extracellular glutamate concentrations, thus protecting neurons from excitotoxicity (Rothstein et al. 1996).

In addition to expressing glutamate-specific transporters (see above), astrocytes and neurons express enzymes that are specific to each cell type (Hertz et al. 1992). This unique compartmentalization originally advanced the hypothesis purporting that these two cell types engage in the exchange of metabolites (Berl, Lajtha, and Waelsch 1961; Lajtha, Berl, and Waelsch 1959). Also derived from this theory is the concept of the glutamate-glutamine cycle, which links glutamatergic neurons and astrocytes (van den Berg and Garfinkel 1971). In the glutamate-glutamine cycle, glutamate released from neurons is predominantly removed by astrocytic GLT1 and GLAST (Gegelashvili and Schousboe 1997, 1998), thereby ensuring a constant flow of glutamine (catalyzed from glutamate by the astrocyte-specific enzyme, GS) from astrocytes to neurons.

Another important concept is that anaplerosis, a requisite reaction for the operation of the tricarboxylic acid (TCA) cycle in the CNS, is exclusively inherent to the astrocyte-specific enzyme, pyruvate carboxylase (PC) (Cesar and Hamprecht 1995; Shank et al. 1985; Yu et al. 1983). PC is a mitochondrial ATP-dependent enzyme containing a biotin prosthetic group, requiring magnesium or manganese and acetyl coenzyme A (CoA). High levels of ADP inhibit the phosphorylation of the enzyme, while acetyl-CoA acts as an allosteric activator of the enzyme. Anaplerosis generates a molecule of oxaloacetate by de novo synthesis. This oxaloacetate molecule then condenses with acetyl-CoA, resulting in the net synthesis of the TCA cycle intermediate, α-ketoglutarate, from which glutamate is formed by transamination (Westergaard et al. 1996). Subsequently, glutamine is synthesized from glutamate in a reaction catalyzed by glutamine synthase (GS), which, analogous to PC, in the CNS, is exclusively expressed in astrocytes (Martinez-Hernandez, Bell, and Norenberg 1977).

Neurotransmitter-mediated metabolic coupling between astrocytes and neurons also invokes lactate release from astrocytes for utilization as an energy source in neurons. The coupling, referred to as the astrocyte neuronal lactate shuttle hypothesis (ANLSH), proposes that glucose enters the CNS via the astrocytic processes. Once in the CNS, glucose then unsheathes the capillaries where it is catabolized to lactate by aerobic glycolysis. Lactate, in turn, can then be shuttled into neurons as an energy source (Magistretti et al. 1994; Pellerin and Magistretti 2004). While remaining somewhat controversial, support for a net lactate transfer between astrocytes and neurons in vivo exists, as recently demonstrated (Pellerin et al. 2007). Enhanced neuronal metabolism occurring in conjunction with elevated levels of CNS electrical activity has been shown to be associated with lactate generated from a non-neuronal compartment, most likely astrocytes (Pellerin et al. 2007; Serres et al. 2004, 2005, 2003). The fact that such a transfer increases with the level of activity is consistent with in vitro observations that have described the redistribution of glucose away from neurons and toward astrocytes upon increased demand. This redistribution reflects enhanced astrocytic glycolysis upon sustained activation to ensure the requisite lactate necessary to maintain ongoing neuronal energy needs (Pellerin et al. 2007). Table 1 summarizes some properties of astrocytes, including their physiological and supportive roles for neurons.

TABLE 1.—

Major functions of astrocytes.

References
Production and release of growth factor Raff et al. (1988); Rudge (1993)
 NGF, BDNF, FGF-2, PDGF, GDNF, TGFp
Regulation of extracellular environment Norenberg (1987); Walz (1989); Copin, Ledig, and Tholey (1992); Ransom and Sontheimer (1992); Makar et al. (1994)
 Homeostasis of H+
 Detoxification of ammonia
 Free radical scavenging
 K+ buffering
Support for neurons Schousboe (1981); Cataldo and Broadwell (1986); Martin (1992); Wang and Cynader (2000); Shanker et al. (2001)
 Supply of TCA cycle intermediates
 Neurotransmitter uptake
 Maintenance of stable GSH levels

NGF, nerve growth factor; BDNF, brain-derived neurotrophic factor; FGF-2, fibroblast growth factor 2; PDGF, platelet-derived growth factor; GDNF, glial cell-derived neurotrophic factor; TGFβ, transforming growth factor beta; TCA, the citric acid cycle.

2. What Is the Evidence in Favor of Astrocytic Modulation of Neurodegeneration?

2A. Role of Astrocytes in Parkinson’s Disease

Parkinson’s disease (PD) is a neurodegenerative disorder associated with the loss of dopamine neurons in the substantia nigra pars compacta (SNpc). It is characterized by slowed movement, rigidity, rest tremor, and bradykinesia (Hornykiewicz and Kish 1987; Lang and Lozano 1998). Metabolism via monoamine oxidase-B (MAO-B) of 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP), a synthetic heroin analog to its active metabolite, MPP+, is known to occur within astrocytes. This pathway has been invoked as the major pathway of MPTP bioactivation and dopamine-specific cell damage, resulting in a Parkinsonian-like syndrome (Heikkila et al. 1989). However, other studies have also invoked astrocytic neuroprotection in the course of exposure to MPTP. Such protection is exemplified by the increased immunoreactivity of the astrocytic marker, glial fibrillary acidic protein (GFAP), in the striatum (Dervan et al. 2004). Moreover, increased numbers of astrocytes and GFAP immunoreactivity have been found in the SNpc of PD postmortem cases (Mirza et al. 2000). Notably, a recent study also showed that α-synuclein immunoreactivity, a major component of Lewy bodies and Lewy neurites appearing in the postmortem brain of PD, is restricted to GFAP-expressing astrocytes (Gu et al. 2010). Astrocytes exert a neuroprotective effect on dopaminergic neurons by secreting a number of neurotrophic factors, such as brain-derived neurotrophic factor (BDNF), glial cell-line-derived neurotrophic factor (GDNF), and mesencephalic astrocyte-derived neurotrophic factor (MANF), as well as numerous antioxidants (Knott et al. 2002; L. F. Lin et al. 1993; Petrova et al. 2003). Release of neurotrophic factors from astrocytes protects dopaminergic neurons in midbrain neuronal/glial cultures (P. S. Chen et al. 2006; Zhang et al. 2005). Astrocytes may also act as scavengers for reactive oxygen species (ROS). Dopamine derived from neurons can be metabolized by astrocytic MAO-B or catechol-O-methyl transferase (COMT), and the resultant free radicals are eliminated by glutathione peroxidase (GPX) (Hirsch et al. 1999; Przedborski and Jackson-Lewis 2000). Moreover, the upregulation of astrocytic protease-activated receptor-1 (PAR-1) in PD has been shown to exert a neuroprotective effect that is mediated by increased levels of GPX (Ishida et al. 2006).

Increased oxidative stress is associated with neuronal cell death in PD (Navarro and Boveris 2009). The transcription factor, NF-E2-related factor (Nrf2), binds to a DNA consensus sequence, antioxidant response element (ARE), and initiates the transcription of genes encoding phase II detoxication enzymes and factors essential for neuronal survival under conditions of oxidative stress (Lee et al. 2005; Rushmore, Morton, and Pickett 1991). Recent reports also indicate that Nrf2 expression that is restricted to astrocytes mediates neuroprotection in the MPTP model (P. C. Chen et al. 2009). Modulation of the Nrf2-ARE pathway in astrocytes may therefore represent a promising therapeutic strategy for the treatment of PD.

2B. Role of Astrocytes in Alzheimer’s Disease

Alzheimer’s disease (AD) is a neurodegenerative disorder and is the most common form of dementia in later life. AD is manifested by the progressive deterioration of cognitive functions such as memory and mental processing (McKhann et al. 1984). Histopathological features of AD include large extracellular senile plaques (SPs) composed of the amyloid-b (Aβ) plagues and neurofibrillary tangles, which are intracellular inclusions of hyperphosphorylated tau protein in selective regions of the brain (Koistinaho et al. 2004; Nagele et al. 2003). Aβ is a peptide of 42 amino acid residues produced by the selective proteolytic cleavage of transmembrane amyloid precursor proteins (APP) by β- and γ-secretases (Haass and Selkoe 1993). Aβ can directly induce neuronal cytotoxi-city, but the relevance of such toxicity to the disease is controversial (Pimplikar 2009; Yankner, Duffy, and Kirschner 1990). Morphological characterization of GFAP-positive astroglial cells performed on AD mouse model at different ages showed an age-dependent reduction in GFAP expression (Rodríguez et al. 2009). These authors suggested that in an AD transgenic, reactive hypertrophic astrocytes surround the neuritic plaques, whereas astroglial cells in other brain regions undergo atrophy, which may account for early changes in synaptic plasticity and cognitive impairments inherent to AD. In the AD human tissue, prominent astrogliosis occurs in the cells surrounding amyloid plaques, and these activated astrocytes accumulate large amounts of Aβ42, which are derived from neuronal debris and associated with plaques (Nagele et al. 2003). Moreover, astrocytes from patients with dementia show significantly decreased complexity compared to the healthy brain (Senitz, Reichenbach, and Smith 1995). In the 3xTg-AD transgenic animal model, which closely resembles the human AD pathology, astrocytes undergo reactive hypertrophy surround the neuritic plaques; whereas throughout the brain parenchyma astrocytes undergo atrophy (Rodríguez et al. 2009; Olabarria et al. 2010).

Astrocytes play an important protective role in AD. While microglial cells are the driving force in SPs formation, astrocytes are crucial in plaque degradation as evidenced by the ultrastructural three-dimensional reconstruction of human classical plaques in different stages of development (Wegiel et al. 2000). Aβ peptides are preferentially internalized by astrocytes, and astrocytic hypertrophic processes degrade Aβ-containing plaques (Kurt, Davies, and Kidd 1999), thus preventing the formation of the deposits of extracellular Aβ (Wyss-Coray et al. 2003). The precise mechanism by which astrocytes recognize and degrade Aβ is not known, but apolipo-protein E (ApoE), which is almost exclusively expressed in astrocytes, has been proposed to be responsible for this cellular action. ApoE is essential for astrocytes to attract chemically, internalize, and degrade Aβ deposits in brain sections in vitro (Koistinaho et al. 2004). Astrocytes also exert protective effects in AD by inhibiting activated microglia. Aβ-induced TGF-β derived from astrocytes can suppress inducible nitric oxide synthase (iNOS) activity in microglia (Vincent, Tilders, and Van Dam 1997). Moreover, astrocyte-conditioned medium from proliferative cultures suppresses activated microglia-induced NO production and phagocytosis of SP cores (DeWitt et al. 1998).

However, the failure of astrocytes to properly degrade Aβ results in the accumulation of Aβ-containing neuronal debris in astrocytes and astrocytic plaque formation (Nagele et al. 2003). Additionally, astrocytes are activated by accumulated Aβ and produce inflammatory mediators, such as interleukin 1β (IL-1β) and tumor necrosis factor-α (TNF-α), which may induce neuronal injury (Johnstone, Gearing, and Miller 1999). Aβ-induced TNF-α increases calcium-dependent gluta-mate release, which may lead to neuronal death in AD (Rossi et al. 2005). Oxidative stress has also been linked to Aβ toxicity as Aβ stimulates ROS production and decreases GSH levels in these cells (Abramov, Canevari, and Duchen 2004; Canevari, Abramov, and Duchen 2004). In addition, Aβ disturbs glucose metabolism in astrocytes, leading to the impairment of neuronal viability (Allaman et al. 2010).

Impairment of glutamatergic neurotransmission associated with excitotoxicity has been implicated in the progression of AD. The glutamate transporter, GLT1, which is preferentially localized in astrocytes, is the major mediator of glutamate clearance in humans. Loss of GLT1 has been reported in the brains of patients with AD (Tian et al. 2010). This effect may be partially mediated by oxidative stress and the differential activity and complex balance between the MAP kinase signaling pathways (Matos et al. 2008).

3. Astrocytes in Neurological Diseases/Conditions

3A. Role of Astrocytes in Ischemia

Prolonged occlusion of cerebral vessels due to cardiac arrest, stroke, or head trauma initiates processes inherent to brain ischemia. Astrocyte swelling is a prominent as well as the earliest response in anoxia-ischemia (Petito et al. 1990). Moreover, the cytoplasm of these astrocytes contains increased numbers of mitochondria and rough endoplasmic reticulum, and the nuclei are enlarged and pale (Norenberg 1981). Astrocytes are known to be critically involved in the pathophysiology of ischemia (Anderson et al. 2003). However, their response to stroke and their role in neuroprotection are not completely understood. Formation of the glial scar during ischemia significantly inhibits functional recovery and processes of regeneration (Fawcett and Asher 1999). Furthermore, focal cerebral ischemia induces astrocytic swelling and leads to an increase in intracerebral pressure, thereby significantly intensifying the ischemic event (Ayata and Ropper 2002). Astrocytic swelling may also reduce the uptake and release of glutamate, potentially initiating excitotoxicity (Anderson et al. 2003). It is known that under ischemic conditions, astrocytic gap junction channels, small pores responsible for homeostasis modulation, may be involved in mediating brain damage by diffusing proapoptotic substances and calcium ions to surrounding healthy cells (Budd and Lipton 1998; J. H. Lin et al. 1998; Rouach et al. 2002). These gap junctions may also contribute to the induction of spreading depression, which has been implicated in infarct expansion (Largo, Cuevas, and Herreras 1996). Notably, rats treated with gap junction blockers, such as octanol and halothane, exhibit reduced infarct volume as well as neuronal death in the permanent focal ischemia model (Saito et al. 1997). Connexin43 (CX43), a principal gap junction forming protein of astrocytes, is associated with protection from ischemic injury (Thompson and MacVicar, 2008). Mice lacking Cx43 in astrocytes showed a significantly increased infarct volume and amplified inflammatory response and apoptosis (Siushansian et al. 2001; Nakase et al. 2004). Astrocytes, on the other hand, have been shown to play a significant role in regeneration during the chronic phase after injury. Astrocytes support neurons by scavenging transmitters released during synaptic activity, controlling ion and water homeostasis and secreting a number of neurotrophic and neuroprotective factors (Y. Chen and Swanson 2003). Many studies also provide evidence for astrocyte-mediated neuroprotection from oxidative stress via a GSH-dependent mechanism (Dringen 2000; Haberg et al. 2001; Iwata-Ichikawa et al. 1999; see also above, Section 1). The inhibition of GSH synthesis increases cortical infarction and edema after ischemia (Mizui, Kinouchi, and Chan 1992). Furthermore, the astrocyte-targeted overexpression of heat shock protein 72 (Hsp72) or superoxide dismutase 2 (SOD2) significantly reduces the loss of CA1 hippocampal neurons in a forebrain ischemia model (Xu et al. 2010). Astrocytes may extend neuronal damage as well as provide neuronal protection under ischemic conditions. Therefore, future efforts aimed at understanding their underlying mechanisms during ischemia are necessary to provide valuable insight into potential therapies.

Formation of the glial scar during ischemia significantly inhibits functional recovery and processes of regeneration (Fawcett and Asher 1999). Furthermore, focal cerebral ischemia induces astrocytic swelling and leads to an increase in intracerebral pressure, thereby significantly intensifying the ischemic event (Ayata and Ropper 2002). Astrocytic swelling may also reduce the uptake and release of glutamate, potentially triggering excitotoxicity (Anderson et al. 2003).

3B. Role of Astrocytes in Hyperammonemia and Hepatic Encephalopathy

The impairment of detoxification processes in chronic or acute liver failure results in increased blood levels of several toxic compounds. One of these compounds, ammonia, readily crosses the blood-brain barrier and accumulates in the central nervous system, where it evokes a number of neuropsychiatric disturbances collectively known as hepatic encephalopathy (HE) (Ferenci et al. 2002; Mullen 2007). Although other toxins involved in HE have been described (Baraldi et al. 1984; Blom et al. 1991; Dejong et al. 2007; Mizoguchi et al. 2001; Montes et al. 2001; Montoliu et al. 2009; Pares et al. 2009), hyperammonemia (HA) is considered to be the primary cause of this disease (Butterworth 2002; Shawcross and Jalan 2005). Cognitive, intellectual, emotional, and behavioral symptoms characterizing HE include the following: circadian rhythm alterations; loss of concentration; depression or euphoria; forgetfulness; confusion; irritability; somnolence; loss of consciousness; and coma, which is the final stage of the disease and usually precedes death (Conn 1994).

The exact mechanisms of ammonia neurotoxicity are not completely known. However it is commonly accepted that astrocytes are the cells that are primarily affected in HE, and neuronal pathology is, to a great extent, secondary to glial dysfunction (Albrecht 2005). The most prominent histopathological changes found in HE that accompanies chronic liver failure include Alzheimer’s type II astrocytosis, (enlarged astrocytes with pale, large nuclei and prominent nucleoli) (Norenberg 1977; Pilbeam, Anderson, and Bhathal 1983) and pronounced astrocytic swelling, leading to brain edema in cases of acute HE (Traber et al. 1987; Wright et al. 2010). Morphological alterations in neurons are observed much less frequently (Butterworth 2007). The high susceptibility of astrocytes to HA may be explained by the fact that when the urea cycle is dysfunctional in the brain, ammonia is detoxified through its condensation with glutamate to form glutamine (Gln). This reaction is catalyzed by the astrocyte-specific enzyme, glutamine synthase (GS) (Cooper et al. 1979; Martinez-Hernandez, Bell, and Norenberg 1977). Increased levels of brain Gln are found in patients suffering from both acute (McConnell et al. 1995; Record et al. 1976) and chronic HE (Laubenberger et al. 1997; Lavoie et al. 1987), as well as in many animal models (Cordoba, Gottstein, and Blei 1996; Hawkins et al. 1993; Hilgier et al. 2008; Zielinska et al. 2004), and are considered to be a key factor in the pathogenesis of this syndrome (Shawcross et al. 2004; Warren and Schenker 1964).

Additionally, a correlation between Gln accumulation and astrocytic swelling has been observed in vivo (Blei et al. 1994; Rama Rao et al. 2010; Takahashi et al. 1991) and in vitro (Norenberg and Bender 1994). However, the hypothesis supporting a direct osmotic effect of Gln (Olafsson, Gottstein, and Blei 1995) appears unlikely (Cordoba et al. 1999; Jayakumar et al. 2006; Zwingmann et al. 2004). Experiments with inhibitors of different mitochondria-related events have shown that the Gln-induced dysfunction of mitochondria may play a key role in astrocytic swelling (Jayakumar et al. 2006; Pichili et al. 2007; Rama Rao et al. 2003). Gln is degraded by the mitochondrial enzyme, phosphate activated glutaminase (PAG) (Bak et al. 2008), and acceleration of this process in HA (Dolinska, Hilgier, and Albrecht 1996; Romero-Gomez et al. 2006) may cause a significant elevation of ammonia levels in astrocytic mitochondria (Albrecht and Norenberg 2006; Kosenko et al. 1996). Increased ammonia concentrations in these organelles lead to the impairment of their functionality as reflected by mitochondrial permeability transition, the loss of mitochondrial transmembrane potential (Bai et al. 2001; Pichili et al. 2007; Rama Rao et al. 2003), a decrease in Krebs cycle activity (Diaz-Munoz and Tapia 1989; Faff-Michalak and Albrecht 1991, 1993; Hindfelt, Plum, and Duffy 1977; Zwingmann et al. 2003), and the loss of ATP (Kosenko et al. 1994; Pichili et al. 2007).

Furthermore, the dysfunction of astrocytic mitochondria may result in oxidative stress. Increased ROS generation has been found in primary astrocytes exposed to ammonia (Murthy et al. 2001) and has been shown to be closely related to accelerated astrocytic Gln metabolism (Jayakumar et al. 2004; Pichili et al. 2007). Oxidative stress has also been observed in animal models of HA and HE (Hilgier et al. 2003; Jiang, Desjardins, and Butterworth 2009; Kosenko et al. 1997). However, very few studies have addressed the role of oxidative stress in HE in humans (Harrison et al. 1991; Jones 1998). Under conditions of HA, the activity of astrocytes in the detoxification of ammonia leads to the overproduction of Gln, the amino acid responsible for a number of pathological cellular processes, thereby significantly contributing to the pathogenesis of HE.

4. Conclusions

Astrocytes play a critical role in normal function of the mammalian nervous system. Astrocytes regulate synaptic transmission and plasticity, protect neurons against toxic compounds, and support metabolically to ensure their optimal functioning. In numerous pathological states, such as AD, PD, or ischemia, astrocytes are involved in neuroprotective mechanisms. As discussed in this review, they support neurons by providing growth factors, cytokines, as well as extracellular matrix molecules, all of which are essential for repair and regeneration. In other conditions, such as HE, disturbances in astrocytic metabolism are implicated in disease pathogenesis. Therefore, modulation of astrocyte functioning may prove to be an efficient therapeutic strategy in many CNS disorders.

Acknowledgments

This article was supported in part by a Public Health Service Grant from the National Institutes of Health, NIEHS 07331, NIEHS 016931, and NIEHS 10563.

Abbreviations:

amyloid-β

AD

Alzheimer’s disease

ANLSH

astrocyte neuronal lactate shuttle hypothesis

ApoE

apolipoprotein E

APP

amyloid precursor proteins

ARE

antioxidant response element

BDNF

brain-derived neurotrophic factor

CNS

central nervous system

CoA

coenzyme A

COMT

catechol-O-methyl transferase

CX43

Connexin43

CSF

cerebrospinal fluid

GABA

γ-aminobutyric acid

GDNF

glial cell-line-derived neurotrophic factor

GFAP

glial fibrillary acid protein

GLAST

glutamate aspartate transporter

Gln

glutamine

GLT1

glutamate transporter 1

GPX

glutathione peroxidase

γ-GS synthase

g-glutamylcysteine synthase

GS

glutamine synthase

GSH

γ glutamylcysteinylglycine

HA

hyperammonemia

HE

hepatic encephalopathy

Hsp72

heat shock protein 72

IL-1β

interleukin 1β

iNOS

inducible nitric oxide synthase

MANF

mesencephalic astrocyte-derived neurotrophic factor

MAO-B

monoamine oxidase-B

MPTP

1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine

Nrf2

NF-E2-related factor

PAG

phosphate activated glutaminase

PAR-1

protease-activated receptor-1

PC

pyruvate carboxylase

PD

Parkinson’s disease

ROS

reactive oxygen species

SNpc

substantia nigra pars compacta

SOD2

superoxide dismutase 2

SPs

senile plaques

TCA

tricarboxylic acid

TNF-α

tumor necrosis factor-α

REFERENCES

  1. Abramov AY, Canevari L, and Duchen MR (2004). Calcium signals induced by amyloid beta peptide and their consequences in neurons and astrocytes in culture. Biochim Biophys Acta 1742, 81–87. [DOI] [PubMed] [Google Scholar]
  2. Albrecht J (2005). Astrocytes in ammonia neurotoxicity: A target, a mediator and a shield In The Role of Glia in Neurotoxicity (Aschner M and Costa LG, eds.), pp. 329–42. CRC Press, Boca Raton, FL. [Google Scholar]
  3. Albrecht J, and Norenberg MD (2006). Glutamine: A Trojan horse in ammonia neurotoxicity. Hepatology 44, 788–94. [DOI] [PubMed] [Google Scholar]
  4. Allaman I, Gavillet M, Belanger M, Laroche T, Viertl D, Lashuel HA, and Magistretti PJ (2010). Amyloid-beta aggregates cause alterations of astrocytic metabolic phenotype: Impact on neuronal viability. J Neurosci 30, 3326–38. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Anderson MF, Blomstrand F, Blomstrand C, Eriksson PS, and Nilsson M (2003). Astrocytes and stroke: Networking for survival? Neurochem Res 28, 293–305. [DOI] [PubMed] [Google Scholar]
  6. Aschner M, and LoPachin RM (2001). Neurotoxic injury and astrocytes In Neuroglia in the Aging Brain (de Vellis J, Ed.), pp. 259–74. Totowa, NJ: Humana. [Google Scholar]
  7. Ayata C, and Ropper AH (2002). Ischaemic brain oedema. J Clin Neurosci 9, 113–24. [DOI] [PubMed] [Google Scholar]
  8. Bai G, Rama Rao KV, Murthy CR, Panickar KS, Jayakumar AR, and Norenberg MD (2001). Ammonia induces the mitochondrial permeability transition in primary cultures of rat astrocytes. J Neurosci Res 66, 981–91. [DOI] [PubMed] [Google Scholar]
  9. Bak LK, Zieminska E, Waagepetersen HS, Schousboe A, and Albrecht J (2008). Metabolism of [U-13C]glutamine and [U-13C]glutamate in isolated rat brain mitochondria suggests functional phosphate-activated glutaminase activity in matrix. Neurochem Res 33, 273–78. [DOI] [PubMed] [Google Scholar]
  10. Baraldi M, Pinelli G, Ricci P, and Zeneroli ML (1984). Toxins in hepatic encephalopathy: The role of the synergistic effect of ammonia, mercaptans and short chain fatty acids. Arch Toxicol Suppl 7, 103–5. [DOI] [PubMed] [Google Scholar]
  11. Berl S, Lajtha A, and Waelsch H (1961). Amino acid and protein metabolism of the brain—VI. Cerebral compartments of glutamic acid metabolism. J Neurochem 7, 322–32. [DOI] [PubMed] [Google Scholar]
  12. Blei AT, Olafsson S, Therrien G, and Butterworth RF (1994). Ammonia-induced brain edema and intracranial hypertension in rats after portacaval anastomosis. Hepatology 19, 1437–44. [PubMed] [Google Scholar]
  13. Blom HJ, Ferenci P, Grimm G, Yap SH, and Tangerman A (1991). The role of methanethiol in the pathogenesis of hepatic encephalopathy. Hepatology 13, 445–54. [PubMed] [Google Scholar]
  14. Budd SL, and Lipton SA (1998). Calcium tsunamis: Do astrocytes transmit cell death messages via gap junctions during ischemia? Nat Neurosci 1, 431–32. [DOI] [PubMed] [Google Scholar]
  15. Butterworth RF (2002). Pathophysiology of hepatic encephalopathy: A new look at ammonia. Metab Brain Dis 17, 221–27. [DOI] [PubMed] [Google Scholar]
  16. Butterworth RF (2007). Neuronal cell death in hepatic encephalopathy. Metab Brain Dis 22, 309–20. [DOI] [PubMed] [Google Scholar]
  17. Canevari L, Abramov AY, and Duchen MR (2004). Toxicity of amyloid beta peptide: Tales of calcium, mitochondria, and oxidative stress. Neurochem Res 29, 637–50. [DOI] [PubMed] [Google Scholar]
  18. Cataldo AM, and Broadwell RD (1986). Cytochemical identification of cerebral glycogen and glucose-6-phosphatase activity under normal and experimental conditions. II. Choroid plexus and ependymal epithelia, endothelia and pericytes. J Neurocytol 15, 511–24. [DOI] [PubMed] [Google Scholar]
  19. Cesar M, and Hamprecht B (1995). Immunocytochemical examination of neural rat and mouse primary cultures using monoclonal antibodies raised against pyruvate carboxylase. J Neurochem 64, 2312–18. [DOI] [PubMed] [Google Scholar]
  20. Chen PC, Vargas MR, Pani AK, Smeyne RJ, Johnson DA, Kan YW, and Johnson JA (2009). Nrf2-mediated neuroprotection in the MPTP mouse model of Parkinson’s disease: Critical role for the astrocyte. Proc Natl Acad Sci U S A 106, 2933–38. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Chen PS, Peng GS, Li G, Yang S, Wu X, Wang CC, Wilson B, Lu RB, Gean PW, Chuang DM, and Hong JS (2006). Valproate protects dopaminergic neurons in midbrain neuron/glia cultures by stimulating the release of neurotrophic factors from astrocytes. Mol Psychiatry 11, 1116–25. [DOI] [PubMed] [Google Scholar]
  22. Chen Y, and Swanson RA (2003). Astrocytes and brain injury. J Cereb Blood Flow Metab 23, 137–49. [DOI] [PubMed] [Google Scholar]
  23. Conn HO (1994). The hepatic encephalopathies. In Hepatic Encephalopathy: Syndromes and Therapies (Conn HO and Bircher J, eds.), pp. 1–12. Medi-Ed Press, Bloomington, IN. [Google Scholar]
  24. Cooper AJ, McDonald JM, Gelbard AS, Gledhill RF, and Duffy TE (1979). The metabolic fate of 13N-labeled ammonia in rat brain. J Biol Chem 254, 4982–92. [PubMed] [Google Scholar]
  25. Copin JC, Ledig M, and Tholey G (1992). Free radical scavenging systems of rat astroglial cells in primary culture: Effects of anoxia and drug treatment. Neurochem Res 17, 677–82. [DOI] [PubMed] [Google Scholar]
  26. Cordoba J, Crespin J, Gottstein J, and Blei AT (1999). Mild hypothermia modifies ammonia-induced brain edema in rats after portacaval anastomosis. Gastroenterology 116, 686–93. [DOI] [PubMed] [Google Scholar]
  27. Cordoba J, Gottstein J, and Blei AT (1996). Glutamine, myo-inositol, and organic brain osmolytes after portocaval anastomosis in the rat: Implications for ammonia-induced brain edema. Hepatology 24, 919–23. [DOI] [PubMed] [Google Scholar]
  28. Dejong CH, van de Poll MC, Soeters PB, Jalan R, and Olde Damink SW (2007). Aromatic amino acid metabolism during liver failure. J Nutr 137, 1579S–85S; discussion 1597S–98S. [DOI] [PubMed] [Google Scholar]
  29. Dervan AG, Meshul CK, Beales M, McBean GJ, Moore C, Totterdell S, Snyder AK, and Meredith GE (2004). Astroglial plasticity and glutamate function in a chronic mouse model of Parkinson’s disease. Exp Neurol 190, 145–56. [DOI] [PubMed] [Google Scholar]
  30. DeWitt DA, Perry G, Cohen M, Doller C, and Silver J (1998). Astrocytes regulate microglial phagocytosis of senile plaque cores of Alzheimer’s disease. Exp Neurol 149, 329–40. [DOI] [PubMed] [Google Scholar]
  31. Diaz-Munoz M, and Tapia R (1989). Functional changes of brain mitochondria during experimental hepatic encephalopathy. Biochem Pharmacol 38, 3835–41. [DOI] [PubMed] [Google Scholar]
  32. Dolinska M, Hilgier W, and Albrecht J (1996). Ammonia stimulates gluta-mine uptake to the cerebral non-synaptic mitochondria of the rat. Neurosci Lett 213, 45–48. [DOI] [PubMed] [Google Scholar]
  33. Dringen R (2000). Metabolism and functions of glutathione in brain. Prog Neurobiol 62, 649–71. [DOI] [PubMed] [Google Scholar]
  34. Dringen R, and Hirrlinger J (2003). Glutathione pathways in the brain. Biol Chem 384, 505–16. [DOI] [PubMed] [Google Scholar]
  35. Dringen R, Pfeiffer B, and Hamprecht B (1999). Synthesis of the antioxidant glutathione in neurons: Supply by astrocytes of CysGly as precursor for neuronal glutathione. J Neurosci 19, 562–69. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Faff-Michalak L, and Albrecht J (1991). Aspartate aminotransferase, malate dehydrogenase, and pyruvate carboxylase activities in rat cerebral synaptic and nonsynaptic mitochondria: Effects of in vitro treatment with ammonia, hyperammonemia and hepatic encephalopathy. Metab Brain Dis 6, 187–97. [DOI] [PubMed] [Google Scholar]
  37. Faff-Michalak L, and Albrecht J (1993). The two catalytic components of the 2-oxoglutarate dehydrogenase complex in rat cerebral synaptic and nonsynaptic mitochondria: Comparison of the response to in vitro treatment with ammonia, hyperammonemia, and hepatic encephalopathy. Neurochem Res 18, 119–23. [DOI] [PubMed] [Google Scholar]
  38. Fawcett JW, and Asher RA (1999). The glial scar and central nervous system repair. Brain Res Bull 49, 377–91. [DOI] [PubMed] [Google Scholar]
  39. Ferenci P, Lockwood A, Mullen K, Tarter R, Weissenborn K, and Blei AT (2002). Hepatic encephalopathy—definition, nomenclature, diagnosis, and quantification: Final report of the working party at the 11th World Congresses of Gastroenterology, Vienna, 1998. Hepatology 35, 716–21. [DOI] [PubMed] [Google Scholar]
  40. Gegelashvili G, and Schousboe A (1997). High affinity glutamate transporters: Regulation of expression and activity. Mol Pharmacol 52, 6–15. [DOI] [PubMed] [Google Scholar]
  41. Gegelashvili G, and Schousboe A (1998). Cellular distribution and kinetic properties of high-affinity glutamate transporters. Brain Res Bull 45, 233–38. [DOI] [PubMed] [Google Scholar]
  42. Gu XL, Long CX, Sun L, Xie C, Lin X, and Cai H (2010). Astrocytic expression of Parkinson’s disease-related A53T alpha-synuclein causes neurodegeneration in mice. Mol Brain 21, 3–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Haass C, and Selkoe DJ (1993). Cellular processing of beta-amyloid precursor protein and the genesis of amyloid beta-peptide. Cell 75, 1039–42. [DOI] [PubMed] [Google Scholar]
  44. Haberg A, Qu H, Saether O, Unsgard G, Haraldseth O, and Sonnewald U (2001). Differences in neurotransmitter synthesis and intermediary metabolism between glutamatergic and GABAergic neurons during 4 hours of middle cerebral artery occlusion in the rat: The role of astrocytes in neuronal survival. J Cereb Blood Flow Metab 21, 1451–63. [DOI] [PubMed] [Google Scholar]
  45. Harrison PM, Wendon JA, Gimson AE, Alexander GJ, and Williams R (1991). Improvement by acetylcysteine of hemodynamics and oxygen transport in fulminant hepatic failure. N Engl J Med 324, 1852–57. [DOI] [PubMed] [Google Scholar]
  46. Hawkins RA, Jessy J, Mans AM, and De Joseph MR (1993). Effect of reducing brain glutamine synthesis on metabolic symptoms of hepatic encephalopathy. J Neurochem 60, 1000–1006. [DOI] [PubMed] [Google Scholar]
  47. Heikkila RE, Sieber BA, Manzino L, and Sonsalla PK (1989). Some features of the nigrostriatal dopaminergic neurotoxin 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) in the mouse. Mol Chem Neuropathol 10, 171–83. [DOI] [PubMed] [Google Scholar]
  48. Hertz L, Peng L, Westergaard N, Yudkoff M, and Schousboe A (1992). Neuronal–astrocytic interactions in metabolism of transmitter amino acids of the glutamate family In Alfred Benzon Symposium 32, Drug Research Related to Neuroactive Amino Acids (Schousboe A, Diemer NH, and Kofod H, eds.), pp. 30–48. Copenhagen, Denmark: Munksgaard. [Google Scholar]
  49. Hilgier W, Anderzhanova E, Oja SS, Saransaari P, and Albrecht J (2003). Taurine reduces ammonia- and N-methyl-D-aspartate-induced accumulation of cyclic GMP and hydroxyl radicals in microdialysates of the rat striatum. Eur J Pharmacol 468, 21–25. [DOI] [PubMed] [Google Scholar]
  50. Hilgier W, Wegrzynowicz M, Maczewski M, Beresewicz A, Oja SS, Saransaari P, and Albrecht J (2008). Effect of glutamine synthesis inhibition with methionine sulfoximine on the nitric oxide-cyclic GMP pathway in the rat striatum treated acutely with ammonia: A microdialysis study. Neurochem Res 33, 267–72. [DOI] [PubMed] [Google Scholar]
  51. Hindfelt B, Plum F, and Duffy TE (1977). Effect of acute ammonia intoxication on cerebral metabolism in rats with portacaval shunts. J Clin Invest 59, 386–96. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Hirsch EC, Hunot S, Damier P, Brugg B, Faucheux BA, Michel PP, Ruberg M, Muriel MP, Mouatt-Prigent A, and Agid Y (1999). Glial cell participation in the degeneration of dopaminergic neurons in Parkinson’s disease. Adv Neurol 80, 9–18. [PubMed] [Google Scholar]
  53. Hornykiewicz O, and Kish SJ (1987). Biochemical pathophysiology of Parkinson’s disease. Adv Neurol 45, 19–34. [PubMed] [Google Scholar]
  54. Ishida Y, Nagai A, Kobayashi S, and Kim SU (2006). Upregulation of protease-activated receptor-1 in astrocytes in Parkinson disease: Astrocyte-mediated neuroprotection through increased levels of glutathione peroxidase. J Neuropathol Exp Neurol 65, 66–77. [DOI] [PubMed] [Google Scholar]
  55. Iwata-Ichikawa E, Kondo Y, Miyazaki I, Asanuma M, and Ogawa N (1999). Glial cells protect neurons against oxidative stress via transcriptional up-regulation of the glutathione synthesis. J Neurochem 72, 2334–44. [DOI] [PubMed] [Google Scholar]
  56. Jayakumar AR, Rama Rao KV, Schousboe A, and Norenberg MD (2004). Glutamine-induced free radical production in cultured astrocytes. Glia 46, 296–301. [DOI] [PubMed] [Google Scholar]
  57. Jayakumar AR, Rao KV, Murthy Ch R, and Norenberg MD (2006). Glutamine in the mechanism of ammonia-induced astrocyte swelling. Neurochem Int 48, 623–28. [DOI] [PubMed] [Google Scholar]
  58. Jiang W, Desjardins P, and Butterworth RF (2009). Minocycline attenuates oxidative/nitrosative stress and cerebral complications of acute liver failure in rats. Neurochem Int 55, 601–5. [DOI] [PubMed] [Google Scholar]
  59. Johnstone M, Gearing AJ, and Miller KM (1999). A central role for astrocytes in the inflammatory response to beta-amyloid; chemokines, cytokines and reactive oxygen species are produced. J Neuroimmunol 93, 182–93. [DOI] [PubMed] [Google Scholar]
  60. Jones AL (1998). Mechanism of action and value of N-acetylcysteine in the treatment of early and late acetaminophen poisoning: A critical review. J Toxicol Clin Toxicol 36, 277–85. [DOI] [PubMed] [Google Scholar]
  61. Kaur P, Aschner M, and Syversen T (2007). Role of glutathione in determining the differential sensitivity between the cortical and cerebellar regions towards mercury-induced oxidative stress. Toxicology 230, 164–77. [DOI] [PubMed] [Google Scholar]
  62. Kimelberg HK (2010). Functions of mature mammalian astrocytes: A current view. Neuroscientist 16, 79–106. [DOI] [PubMed] [Google Scholar]
  63. Knott C, Stern G, Kingsbury A, Welcher AA, and Wilkin GP (2002). Elevated glial brain-derived neurotrophic factor in Parkinson’s diseased nigra. Parkinsonism Relat Disord 8, 329–41. [DOI] [PubMed] [Google Scholar]
  64. Koistinaho M, Lin S, Wu X, Esterman M, Koger D, Hanson J, Higgs R, Liu F, Malkani S, Bales KR, and Paul SM (2004). Apolipoprotein E promotes astrocyte colocalization and degradation of deposited amyloid-beta peptides. Nat Med 10, 719–26. [DOI] [PubMed] [Google Scholar]
  65. Kosenko E, Felipo V, Montoliu C, Grisolía S, and Kaminsky Y (1996). Effects of acute hyperammonemia in vivo on oxidative metabolism in nonsynaptic rat brain mitochondria. Metab Brain Dis 12, 69–82. [DOI] [PubMed] [Google Scholar]
  66. Kosenko E, Kaminsky Y, Grau E, Minana MD, Marcaida G, Grisolia S, and Felipo V (1994). Brain ATP depletion induced by acute ammonia intoxication in rats is mediated by activation of the NMDA receptor and Na+,K(+)-ATPase. J Neurochem 63, 2172–78. [DOI] [PubMed] [Google Scholar]
  67. Kosenko E, Kaminsky Y, Kaminsky A, Valencia M, Lee L, Hermenegildo C, and Felipo V (1997). Superoxide production and antioxidant enzymes in ammonia intoxication in rats. Free Radic Res 27, 637–44. [DOI] [PubMed] [Google Scholar]
  68. Kurt MA, Davies DC, and Kidd M (1999). beta-Amyloid immunoreactivity in astrocytes in Alzheimer’s disease brain biopsies: An electron microscope study. Exp Neurol 158, 221–28. [DOI] [PubMed] [Google Scholar]
  69. Lajtha A, Berl S, and Waelsch H (1959). Amino acid and protein metabolism of the brain. IV. The metabolism of glutamic acid. J Neurochem 3, 322–32. [DOI] [PubMed] [Google Scholar]
  70. Lang AE, and Lozano AM (1998). Parkinson’s disease. First of two parts. N Engl J Med 339, 1044–53. [DOI] [PubMed] [Google Scholar]
  71. Largo C, Cuevas P, and Herreras O (1996). Is glia disfunction the initial cause of neuronal death in ischemic penumbra? Neurol Res 18, 445–48. [DOI] [PubMed] [Google Scholar]
  72. Laubenberger J, Haussinger D, Bayer S, Gufler H, Hennig J, and Langer M (1997). Proton magnetic resonance spectroscopy of the brain in symptomatic and asymptomatic patients with liver cirrhosis. Gastroenterology 112, 1610–16. [DOI] [PubMed] [Google Scholar]
  73. Lavoie J, Giguere JF, Layrargues GP, and Butterworth RF (1987). Amino acid changes in autopsied brain tissue from cirrhotic patients with hepatic encephalopathy. J Neurochem 49, 692–97. [DOI] [PubMed] [Google Scholar]
  74. Lee JM, Li J, Johnson DA, Stein TD, Kraft AD, Calkins MJ, Jakel RJ, and Johnson JA (2005). Nrf2, a multi-organ protector? FASEB J 19, 1061–66. [DOI] [PubMed] [Google Scholar]
  75. Lehre KP, Levy LM, Ottersen OP, Storm-Mathisen J, and Danbolt NC (1995). Differential expression of two glial glutamate transporters in the rat brain: Quantitative and immunocytochemical observations. J Neurosci 15, 1835–53. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Lin JH, Weigel H, Cotrina ML, Liu S, Bueno E, Hansen AJ, Hansen TW, Goldman S, and Nedergaard M (1998). Gap-junction-mediated propagation and amplification of cell injury. Nat Neurosci 1, 494–500. [DOI] [PubMed] [Google Scholar]
  77. Lin LF, Doherty DH, Lile JD, Bektesh S, and Collins F (1993). GDNF: A glial cell line-derived neurotrophic factor for midbrain dopaminergic neurons. Science 260, 1130–32. [DOI] [PubMed] [Google Scholar]
  78. Magistretti PJ, Sorg O, Naichen Y, Pellerin L, de Rham S, and Martin JL (1994). Regulation of astrocyte energy metabolism by neurotransmitters. Ren Physiol Biochem 17, 168–71. [DOI] [PubMed] [Google Scholar]
  79. Makar TK, Nedergaard M, Preuss A, Gelbard AS, Perumal AS, and Cooper AJ (1994). Vitamin E, ascorbate, glutathione, glutathione disulfide, and enzymes of glutathione metabolism in cultures of chick astrocytes and neurons: Evidence that astrocytes play an important role in antioxidative processes in the brain. J Neurochem 62, 45–53. [DOI] [PubMed] [Google Scholar]
  80. Martin DL (1992). Synthesis and release of neuroactive substances by glial cells. Glia 5, 81–94. [DOI] [PubMed] [Google Scholar]
  81. Martinez-Hernandez A, Bell KP, and Norenberg MD (1977). Glutamine synthetase: Glial localization in brain. Science 195, 1356–58. [DOI] [PubMed] [Google Scholar]
  82. Matos M, Augusto E, Oliveira CR, and Agostinho P (2008). Amyloid-beta peptide decreases glutamate uptake in cultured astrocytes: Involvement of oxidative stress and mitogen-activated protein kinase cascades. Neuroscience 156, 898–910. [DOI] [PubMed] [Google Scholar]
  83. McConnell JR, Antonson DL, Ong CS, Chu WK, Fox IJ, Heffron TG, Langnas AN, and Shaw BW Jr. (1995). Proton spectroscopy of brain glutamine in acute liver failure. Hepatology 22, 69–74. [PubMed] [Google Scholar]
  84. McKhann G, Drachman D, Folstein M, Katzman R, Price D, and Stadlan EM (1984). Clinical diagnosis of Alzheimer’s disease: Report of the NINCDS-ADRDA Work Group under the auspices of Department of Health and Human Services Task Force on Alzheimer’s Disease. Neurology 34, 939–44. [DOI] [PubMed] [Google Scholar]
  85. Mirza B, Hadberg H, Thomsen P, Moos T (2000). The absence of reactive astrocytosis is indicative of a unique inflammatory process in Parkinson’s disease. Neuroscience 95, 425–32. [DOI] [PubMed] [Google Scholar]
  86. Mizoguchi N, Nishimura Y, Ono H, and Sakura N (2001). Manganese elevations in blood of children with congenital portosystemic shunts. Eur J Pediatr 160, 247–50. [DOI] [PubMed] [Google Scholar]
  87. Mizui T, Kinouchi H, and Chan PH (1992). Depletion of brain glutathione by buthionine sulfoximine enhances cerebral ischemic injury in rats. Am J Physiol 262, H313–17. [DOI] [PubMed] [Google Scholar]
  88. Montes S, Alcaraz-Zubeldia M, Muriel P, and Rios C (2001). Striatal manganese accumulation induces changes in dopamine metabolism in the cirrhotic rat. Brain Res 891, 123–29. [DOI] [PubMed] [Google Scholar]
  89. Montoliu C, Piedrafita B, Serra MA, del Olmo JA, Urios A, Rodrigo JM, and Felipo V (2009). IL-6 and IL-18 in blood may discriminate cirrhotic patients with and without minimal hepatic encephalopathy. J Clin Gastroenterol 43, 272–79. [DOI] [PubMed] [Google Scholar]
  90. Mullen KD (2007). Review of the final report of the 1998 Working Party on Definition, Nomenclature and Diagnosis of Hepatic Encephalopathy. Aliment Pharmacol Ther 25 (Suppl. 1), 11–16. [DOI] [PubMed] [Google Scholar]
  91. Murthy CR, Rama Rao KV, Bai G, and Norenberg MD (2001). Ammonia-induced production of free radicals in primary cultures of rat astrocytes. J Neurosci Res 66, 282–88. [DOI] [PubMed] [Google Scholar]
  92. Nagele RG, D’Andrea MR, Lee H, Venkataraman V, and Wang HY (2003). Astrocytes accumulate A beta 42 and give rise to astrocytic amyloid plaques in Alzheimer disease brains. Brain Res 971, 197–209. [DOI] [PubMed] [Google Scholar]
  93. Nakase T, Sohl G, Theis M, Willecke K, and Naus CCG (2004). Increased apoptosis and inflammation after focal brain ischemia in mice lacking connexin43 in astrocytes. American J Pathol 164, 2067–75. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Navarro A, and Boveris A (2009). Brain mitochondrial dysfunction and oxidative damage in Parkinson’s disease. J Bioenerg Biomembr 41, 517–21. [DOI] [PubMed] [Google Scholar]
  95. Norenberg MD (1977). A light and electron microscopic study of experimental portal-systemic (ammonia) encephalopathy. Progression and reversal of the disorder. Lab Invest 36, 618–27. [PubMed] [Google Scholar]
  96. Norenberg MD (1981). The astrocyte in liver disease In Advances in Cellular Neurobiology 2 (Fedoroff S and Hertz L, eds.), pp. 303–52. San Diego, CA: Academic Press. [Google Scholar]
  97. Norenberg MD (1987). The role of astrocytes in hepatic encephalopathy. Neurochem Pathol 6, 13–33. [DOI] [PubMed] [Google Scholar]
  98. Norenberg MD, and Bender AS (1994). Astrocyte swelling in liver failure: Role of glutamine and benzodiazepines. Acta Neurochir Suppl (Wien) 60, 24–27. [DOI] [PubMed] [Google Scholar]
  99. Olabarria M, Noristani HN, Verkhratsky A, and Rodríguez JJ (2010). Concomitant astroglial atrophy and astrogliosis in a triple transgenic animal model of Alzheimer’s disease. Glia 58, 831–38. [DOI] [PubMed] [Google Scholar]
  100. Olafsson S, Gottstein J, and Blei AT (1995). Brain edema and intracranial hypertension in rats after total hepatectomy. Gastroenterology 108, 1097–1103. [DOI] [PubMed] [Google Scholar]
  101. Pares A, Deulofeu R, Cisneros L, Escorsell A, Salmeron JM, Caballeria J, and Mas A (2009). Albumin dialysis improves hepatic encephalopathy and decreases circulating phenolic aromatic amino acids in patients with alcoholic hepatitis and severe liver failure. Crit Care 13, R8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Pellerin L, Bouzier-Sore AK, Aubert A, Serres S, Merle M, Costalat R, and Magistretti PJ (2007). Activity-dependent regulation of energy metabolism by astrocytes: An update. Glia 55, 1251–62. [DOI] [PubMed] [Google Scholar]
  103. Pellerin L, and Magistretti PJ (2004). Neuroenergetics: Calling upon astrocytes to satisfy hungry neurons. Neuroscientist 10, 53–62. [DOI] [PubMed] [Google Scholar]
  104. Petito CK, Morgello S, Felix JC, and Lesser ML (1990). The two patterns of reactive astrocytosis in postischemic rat brain. J Cereb Blood Flow Metab 10, 850–59. [DOI] [PubMed] [Google Scholar]
  105. Petrova P, Raibekas A, Pevsner J, Vigo N, Anafi M, Moore MK, Peaire AE, Shridhar V, Smith DI, Kelly J, Durocher Y, and Commissiong JW (2003). MANF: A new mesencephalic, astrocyte-derived neurotrophic factor with selectivity for dopaminergic neurons. J Mol Neurosci 20, 173–88. [DOI] [PubMed] [Google Scholar]
  106. Pichili VB, Rao KV, Jayakumar AR, and Norenberg MD (2007). Inhibition of glutamine transport into mitochondria protects astrocytes from ammonia toxicity. Glia 55, 801–9. [DOI] [PubMed] [Google Scholar]
  107. Pilbeam CM, Anderson RM, and Bhathal PS (1983). The brain in experimental portal-systemic encephalopathy. I. Morphological changes in three animal models. J Pathol 140, 331–45. [DOI] [PubMed] [Google Scholar]
  108. Pimplikar SW (2009). Reassessing the amyloid cascade hypothesis of Alzheimer’s disease. Int J Biochem Cell Biol 41, 1261–68. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Przedborski S, and Jackson-Lewis V (2000). ROS and Parkinson’s disease: A view to a kill In Free Radicals in Brain Pathophysiology (Poli G, Cadenas E, and Packer L, eds.), pp. 273–90. New York, NY: Marcel Dekker. [Google Scholar]
  110. Raff MC, Lillien LE, Richardson WD, Burne JF, and Noble MD (1988). Platelet-derived growth factor from astrocytes drives the clock that times oligodendrocyte development in culture. Nature 333, 562–65. [DOI] [PubMed] [Google Scholar]
  111. Rama Rao KV, Chen M, Simard JM, and Norenberg MD (2003). Suppression of ammonia-induced astrocyte swelling by cyclosporin A. J Neurosci Res 74, 891–97. [DOI] [PubMed] [Google Scholar]
  112. Rama Rao KV, Reddy PV, Tong X, and Norenberg MD (2010). Brain edema in acute liver failure: Inhibition by L-histidine. Am J Pathol 176, 1400–1408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Ransom BR, and Sontheimer H (1992). The neurophysiology of glial cells. J Clin Neurophysiol 9, 224–51. [DOI] [PubMed] [Google Scholar]
  114. Record CO, Buxton B, Chase RA, Curzon G, Murray-Lyon IM, and Williams R (1976). Plasma and brain amino acids in fulminant hepatic failure and their relationship to hepatic encephalopathy. Eur J Clin Invest 6, 387–94. [DOI] [PubMed] [Google Scholar]
  115. Rodríguez JJ, Olabarria M, Chvatal A, and Verkhratsky A (2009). Astroglia in dementia and Alzheimer’s disease. Cell Death Differ 16, 378–85. [DOI] [PubMed] [Google Scholar]
  116. Romero-Gomez M, Jover M, Diaz-Gomez D, de Teran LC, Rodrigo R, Camacho I, Echevarria M, Felipo V, and Bautista JD (2006). Phosphate-activated glutaminase activity is enhanced in brain, intestine and kidneys of rats following portacaval anastomosis. World J Gastroenterol 12, 2406–11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Rossi D, Brambilla L, Valori CF, Crugnola A, Giaccone G, Capobianco R, Mangieri M, Kingston AE, Bloc A, Bezzi P, and Volterra A (2005). Defective tumor necrosis factor-alpha-dependent control of astrocyte glutamate release in a transgenic mouse model of Alzheimer disease. J Biol Chem 280, 42088–96. [DOI] [PubMed] [Google Scholar]
  118. Rothstein JD, Dykes-Hoberg M, Pardo CA, Bristol LA, Jin L, Kuncl RW, Kanai Y, Hediger MA, Wang Y, Schielke JP, and Welty DF (1996). Knockout of glutamate transporters reveals a major role for astroglial transport in excitotoxicity and clearance of glutamate. Neuron 16, 675–86. [DOI] [PubMed] [Google Scholar]
  119. Rouach N, Avignone E, Meme W, Koulakoff A, Venance L, Blomstrand F, and Giaume C (2002). Gap junctions and connexin expression in the normal and pathological central nervous system. Biol Cell 94, 457–75. [DOI] [PubMed] [Google Scholar]
  120. Rudge JS (1993). Astrocyte-derived neurotrophic factors In Astrocytes: Pharmacology and Function (Murphy S, Ed.), pp. 267–305. San Diego, CA: Academic Press. [Google Scholar]
  121. Rushmore TH, Morton MR, and Pickett CB (1991). The antioxidant responsive element. Activation by oxidative stress and identification of the DNA consensus sequence required for functional activity. J Biol Chem 266, 11632–39. [PubMed] [Google Scholar]
  122. Sagara JI, Miura K, and Bannai S (1993). Maintenance of neuronal glutathione by glial cells. J Neurochem 61, 1672–76. [DOI] [PubMed] [Google Scholar]
  123. Saito R, Graf R, Hubel K, Fujita T, Rosner G, and Heiss WD (1997). Reduction of infarct volume by halothane: Effect on cerebral blood flow or perifocal spreading depression-like depolarizations. J Cereb Blood Flow Metab 17, 857–64. [DOI] [PubMed] [Google Scholar]
  124. Schousboe A (1981). Transport and metabolism of glutamate and GABA in neurons and glial cells. Int Rev Neurobiol 22, 1–45. [DOI] [PubMed] [Google Scholar]
  125. Senitz D, Reichenbach A, and Smith TG (1995). Surface complexity of human neocortical astrocytic cells: Changes with development, aging, and dementia. J Hirnforsch 36, 531–37. [PubMed] [Google Scholar]
  126. Serres S, Bezancon E, Franconi JM, and Merle M (2004). Ex vivo analysis of lactate and glucose metabolism in the rat brain under different states of depressed activity. J Biol Chem 279, 47881–89. [DOI] [PubMed] [Google Scholar]
  127. Serres S, Bezancon E, Franconi JM, and Merle M (2005). Ex vivo NMR study of lactate metabolism in rat brain under various depressed states. J Neurosci Res 79, 19–25. [DOI] [PubMed] [Google Scholar]
  128. Serres S, Bouyer JJ, Bezancon E, Canioni P, and Merle M (2003). Involvement of brain lactate in neuronal metabolism. NMR Biomed 16, 430–39. [DOI] [PubMed] [Google Scholar]
  129. Shank RP, Bennett GS, Freytag SO, and Campbell GL (1985). Pyruvate carboxylase: An astrocyte-specific enzyme implicated in the replenishment of amino acid neurotransmitter pools. Brain Res 329, 364–67. [DOI] [PubMed] [Google Scholar]
  130. Shanker G, Allen JW, Mutkus LA, and Aschner M (2001). The uptake of cysteine in cultured primary astrocytes and neurons. Brain Res 902, 156–63. [DOI] [PubMed] [Google Scholar]
  131. Shawcross DL, Balata S, Olde Damink SW, Hayes PC, Wardlaw J, Marshall I, Deutz NE, Williams R, and Jalan R (2004). Low myoinositol and high glutamine levels in brain are associated with neuropsychological deterioration after induced hyperammonemia. Am J Physiol Gastrointest Liver Physiol 287, G503–9. [DOI] [PubMed] [Google Scholar]
  132. Shawcross D, and Jalan R (2005). The pathophysiologic basis of hepatic encephalopathy: Central role for ammonia and inflammation. Cell Mol Life Sci 62, 2295–2304. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Siushansian R, Bechberger JF, Cechetto DF, Hachinski VC, and Naus CC (2001). Connexin43 null mutation increases infarct size after stroke. J Comp Neurol Nov 26, 387–94. [DOI] [PubMed] [Google Scholar]
  134. Storck T, Schulte S, Hofmann K, and Stoffel W (1992). Structure, expression, and functional analysis of a Na(+)-dependent glutamate/aspartate transporter from rat brain. Proc Natl Acad Sci U S A 89, 10955–59. [DOI] [PMC free article] [PubMed] [Google Scholar]
  135. Takahashi H, Koehler RC, Brusilow SW, and Traystman RJ (1991). Inhibition of brain glutamine accumulation prevents cerebral edema in hyperammonemic rats. Am J Physiol 261, H825–29. [DOI] [PubMed] [Google Scholar]
  136. Thompson RJ, and MacVicar BA (2008). Connexin and pannexin hemi-channels of neurons and astrocytes. Channels 2, 81–86. [DOI] [PubMed] [Google Scholar]
  137. Tian G, Kong Q, Lai L, Ray-Chaudhury A, and Lin CL (2010). Increased expression of cholesterol 24S-hydroxylase results in disruption of glial glutamate transporter EAAT2 association with lipid rafts: A potential role in Alzheimer’s disease. J Neurochem 113, 978–89. [DOI] [PMC free article] [PubMed] [Google Scholar]
  138. Traber PG, Dal Canto M, Ganger DR, and Blei AT (1987). Electron microscopic evaluation of brain edema in rabbits with galactosamine-induced fulminant hepatic failure: Ultrastructure and integrity of the blood-brain barrier. Hepatology 7, 1272–77. [DOI] [PubMed] [Google Scholar]
  139. van den Berg CJ, and Garfinkel D (1971). A stimulation study of brain compartments. Metabolism of glutamate and related substances in mouse brain. Biochem J 123, 211–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Vincent VA, Tilders FJ, and Van Dam AM (1997). Inhibition of endotoxin-induced nitric oxide synthase production in microglial cells by the presence of astroglial cells: A role for transforming growth factor beta. Glia 19, 190–98. [DOI] [PubMed] [Google Scholar]
  141. Walz W (1989). Role of glial cells in the regulation of the brain ion microenvironment. Prog Neurobiol 33, 309–33. [DOI] [PubMed] [Google Scholar]
  142. Wang XF, and Cynader MS (2000). Astrocytes provide cysteine to neurons by releasing glutathione. J Neurochem 74, 1434–42. [DOI] [PubMed] [Google Scholar]
  143. Warren KS, and Schenker S (1964). Effect of an inhibitor of glutamine synthesis (methionine sulfoximine) on ammonia toxicity and metabolism. J Lab Clin Med 64, 442–49. [PubMed] [Google Scholar]
  144. Wegiel J, Wang KC, Tarnawski M, and Lach B (2000). Microglia cells are the driving force in fibrillar plaque formation, whereas astrocytes are a leading factor in plague degradation. Acta Neuropathol 100, 356–64. [DOI] [PubMed] [Google Scholar]
  145. Westergaard N, Drejer J, Schousboe A, and Sonnewald U (1996). Evaluation of the importance of transamination versus deamination in astrocytic metabolism of [U-13C]glutamate. Glia 17, 160–68. [DOI] [PubMed] [Google Scholar]
  146. Wright G, Soper R, Brooks HF, Stadlbauer V, Vairappan B, Davies NA, Andreola F, Hodges S, Moss RF, Davies DC, and Jalan R (2010). Role of aquaporin-4 in the development of brain oedema in liver failure. J Hepatol 53, 91–97. [DOI] [PubMed] [Google Scholar]
  147. Wyss-Coray T, Loike JD, Brionne TC, Lu E, Anankov R, Yan F, Silverstein SC, and Husemann J (2003). Adult mouse astrocytes degrade amyloid-beta in vitro and in situ. Nat Med 9, 453–57. [DOI] [PubMed] [Google Scholar]
  148. Xu L, Emery JF, Ouyang YB, Voloboueva LA, and Giffard RG (2010). Astrocyte targeted overexpression of Hsp72 or SOD2 reduces neuronal vulnerability to forebrain ischemia. Glia 58, 1042–49. [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Yankner BA, Duffy LK, and Kirschner DA (1990). Neurotrophic and neurotoxic effects of amyloid beta protein: Reversal by tachykinin neuropeptides. Science 250, 279–82. [DOI] [PubMed] [Google Scholar]
  150. Yu AC, Drejer J, Hertz L, and Schousboe A (1983). Pyruvate carboxylase activity in primary cultures of astrocytes and neurons. J Neurochem 41, 1484–87. [DOI] [PubMed] [Google Scholar]
  151. Zhang W, Qin L, Wang T, Wei SJ, Gao HM, Liu J, Wilson B, Liu B, Kim HC, and Hong JS (2005). 3-hydroxymorphinan is neurotrophic to dopaminergic neurons and is also neuroprotective against LPS-induced neurotoxicity. FASEB J 19, 395–97. [DOI] [PubMed] [Google Scholar]
  152. Zielinska M, Stafiej A, Law RO, and Albrecht J (2004). Effects of methionine sulfoximine on the glutamine and glutamate content and cell volume in rat cerebral cortical slices: Involvement of mechanisms not related to inhibition of glutamine synthesis. Neurotoxicology 25, 443–49. [DOI] [PubMed] [Google Scholar]
  153. Zwingmann C, Chatauret N, Leibfritz D, and Butterworth RF (2003). Selective increase of brain lactate synthesis in experimental acute liver failure: Results of a [H-C] nuclear magnetic resonance study. Hepatology 37, 420–28. [DOI] [PubMed] [Google Scholar]
  154. Zwingmann C, Chatauret N, Rose C, Leibfritz D, and Butterworth RF (2004). Selective alterations of brain osmolytes in acute liver failure: Protective effect of mild hypothermia. Brain Res 999, 118–23. [DOI] [PubMed] [Google Scholar]

RESOURCES