ArticlePDF Available

Cellular interplay between neurons and glia: toward a comprehensive mechanism for excitotoxic neuronal loss in neurodegeneration

Authors:

Abstract and Figures

Astrocytes perform vital maintenance, functional enhancement, and protective roles for their associated neurons; however these same mechanisms may become deleterious for neurons under some conditions. In this review, we highlight two normally protective pathways, the endoplasmic reticulum (ER) stress response and an endogenous antioxidant response, which may become neurotoxic when activated in astrocytes during the inflammation associated with neurodegeneration. Stimulation of these multifaceted pathways affects a panoply of cellular processes. Of particular importance is the effect these pathways have on the homeostasis of the excitatory amino acid neurotransmitter, glutamate. The endogenous antioxidant response increases extracellular glutamate in the pursuit of making the cellular antioxidant, glutathione, by increasing expression of the xCT subunit of the cystine/glutamate antiporter. Meanwhile, inflammatory mediators such as TNFα reduce levels of membrane-bound glutamate scavenging proteins such as the excitatory amino acid transporters. Together, these cellular activities may result in a net increase in extracellular glutamate that could alter neuronal function and lead to excitotoxicity. Here we discuss the role of N-methyl-D-aspartate receptors, which, when excessively stimulated by glutamate, can cause neuronal dysfunction and loss via activation of calpains. While there are other pathways acting in concert or parallel to those we describe here, this review explores a rationale to explain how two protective mechanisms may result in neuronal loss during neurodegeneration.
Content may be subject to copyright.
Cellular interplay between neurons and glia: toward a
comprehensive mechanism for excitotoxic neuronal loss in
neurodegeneration
Alison J.B. Markowitz1, Michael G. White1, Dennis L. Kolson2, and Kelly L. Jordan-
Sciutto1,*
1 Department of Pathology, School of Dental Medicine, University of Pennsylvania, Philadelphia, PA 19104
2 Department of Neurology, School of Medicine, University of Pennsylvania, Philadelphia, PA, 19104
Abstract
Astrocytes perform vital maintenance, functional enhancement, and protective roles for their
associated neurons; however these same mechanisms may become deleterious for neurons under
some conditions. In this review, we highlight two normally protective pathways, the endoplasmic
reticulum (ER) stress response and an endogenous antioxidant response, which may become
neurotoxic when activated in astrocytes during the inflammation associated with neurodegeneration.
Stimulation of these multifaceted pathways affects a panoply of cellular processes. Of particular
importance is the effect these pathways have on the homeostasis of the excitatory amino acid
neurotransmitter, glutamate. The endogenous antioxidant response increases extracellular glutamate
in the pursuit of making the cellular antioxidant, glutathione, by increasing expression of the xCT
subunit of the cystine/glutamate antiporter. Meanwhile, inflammatory mediators such as TNFα
reduce levels of membrane-bound glutamate scavenging proteins such as the excitatory amino acid
transporters. Together, these cellular activities may result in a net increase in extracellular glutamate
that could alter neuronal function and lead to excitotoxicity. Here we discuss the role of N-methyl-
D-aspartate receptors, which, when excessively stimulated by glutamate, can cause neuronal
dysfunction and loss via activation of calpains. While there are other pathways acting in concert or
parallel to those we describe here, this review explores a rationale to explain how two protective
mechanisms may result in neuronal loss during neurodegeneration.
Keywords
glutamate; endoplasmic reticulum stress; oxidative stress; brain; antioxidant response; Nrf2; PERK;
calpain; NMDA receptor; integrated stress response; xCT; excitatory amino acid transporters
Introduction: Insults and Injury in Neurodegeneration
Slow, chronic, progressive damage and loss of neurons underlie the symptoms observed in
neurodegenerative diseases including Alzheimer disease (AD), Parkinson disease (PD),
amyotrophic lateral sclerosis (ALS) and HIV-associated dementia (HAD). Although the
precise mechanisms of neuronal loss in these diseases have yet to be elucidated, evidence from
human autopsy tissue, animal studies, and in vitro work demonstrate a prominent role for
*To Whom Correspondence Should Be Addressed: Kelly L. Jordan-Sciutto, PhD, 240 S. 40th St, Room 312 Levy Building, Department
of Pathology, University of Pennsylvania, Philadelphia, PA 19104, phone: 215-898-4196, fax: 215-573-2050, e-mail:
Jordan@path.dental.upenn.edu.
NIH Public Access
Author Manuscript
Cellscience. Author manuscript; available in PMC 2009 January 2.
Published in final edited form as:
Cellscience. 2007 July 27; 4(1): 111–146.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
oxidative stress (Lipton et al., 1994; Ehret et al., 1996; Koutsilieri et al., 1997; Kruman et al.,
1998; Malorni et al., 1998; Romero-Alvira and Roche, 1998; Mollace et al., 2001; Viviani et
al., 2001). Endogenous agents that reduce cellular oxidative damage such as glutathione and
catalase are depleted in patients with AD, PD, ALS and HAD (Sofic et al., 1992; Castagna et
al., 1995; Sohal and Weindruch, 1996; Lovell et al., 1998; Choi et al., 2000; Sayre et al.,
2001). By providing exogenous antioxidants or enzymes that destroy free radicals in animal
models or in vitro models of AD, PD, and HAD (Kruman et al., 1998; Malorni et al., 1998;
Floyd et al., 1999; Ramassamy et al., 1999; Wang and Xu, 2005) or as therapeutics in human
trials (Anonymous, 1998; Muller et al., 2000; Shor-Posner et al., 2002; Turchan et al., 2003;
Weinreb et al., 2004), neuronal loss can be attenuated. Though the sufficiency of these
treatments in preventing neurodegeneration still remains unclear, these studies suggest that
reduction in oxidative stress will mitigate disease progression.
Neuroinflammation can trigger oxidative stress and act as a prominent insult in AD, PD, ALS,
and HAD. While there is evidence of microgliosis and astrogliosis in each of these
neurodegenerative conditions, neuroinflammation is most closely associated with disease
etiology and progression in HAD. Pathologic studies of the brains of patients with HAD suggest
an inflammatory mechanism in the progression of this disease (Kaul et al., 2001; Garden et al.,
2002; Ghorpade et al., 2003). Neuronal dysfunction and death in HAD is mediated by the
release of neurotoxic factors from activated macrophages and microglia including, but not
limited to, reactive oxygen species and excitatory amino acids (Pulliam et al., 1991; Yoshioka
et al., 1995; Conant et al., 1998; Bagetta et al., 1999; Nath et al., 1999; Kaul et al., 2005).
Moreover, these factors stimulate astrocytes and neurons to release additional pro-
inflammatory factors, perpetuating the inflammatory state in the CNS, and potentially
exacerbating neuronal damage (Gonzalez-Scarano and Martin-Garcia, 2005). Soluble
macrophage factors associated with neurologic damage in the CNS of AIDS patients include
excitotoxins (i.e. quinolinic acid, glutamate), reactive oxygen species (ROS), cytokines,
chemokines and neurotrophic factors (Masliah et al., 1992; Brouwers et al., 1993; Gelbard et
al., 1994; Pulliam et al., 1994; Crowe, 1995; Achim and Wiley, 1996; Giulian et al., 1996;
Pulliam et al., 1996; Heyes et al., 1998; Sanders et al., 1998). How neurons “interpret” the
conflicting signaling environment in HAD or other neurodegenerative diseases ultimately
determines viability. Factors such as neurotrophic factor and chemokines have been reported
to activate survival pathways in neurons as part of their signaling cascades, while ROS and
cytokines can induce death via apoptosis or necrosis (Nicotera et al., 1997; D’Mello, 1998;
Pulliam et al., 1998; Bibel and Barde, 2000; Boldyrev, 2000; Holmin and Mathiesen, 2000;
Takikita et al., 2001; Annunziato et al., 2003; Barzilai et al., 2003). Understanding the impact
of these factors with disparate effects on neuronal viability is essential to unraveling the
mechanisms that lead to neurodegeneration.
As important as investigating the mechanisms of neuronal loss in neurodegeneration is the role
of neuronal support cells in neuroprotection and function, with particular emphasis on the
astrocytes. Astrocytes not only support neuronal survival, but provide significant support of
synaptic function by modulating neurotransmitter reuptake, neurotransmitter metabolism, and
neurotrophic factor release (reviewed in (Benarroch, 2005)). Astrocytes have been shown to
be sufficient to provide neuronal protection from oxidative stress, an insult associated with
neurodegeneration (Shih et al., 2003; Shih et al., 2005). The cellular pathway by which
astrocytes protect from oxidative stress is part of the integrated stress response (Cullinan et al.,
2003; Shih et al., 2003; Shih et al., 2005; Cullinan and Diehl, 2006), a recently described
mechanisms made up of at least two pathways important for maintenance of cellular
homeostasis. In this review, we will highlight the importance of these protective pathways in
the interplay between neurons and astrocytes under neuroinflammatory conditions and propose
a model by which protective pathways can be subverted to neurotoxic pathways culminating
in neuronal dysfunction and death.
Markowitz et al. Page 2
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Altered Astrocyte Functions in Response to Neurodegenerative Insults
Integrated Stress Response (ISR)
In the past several years, it has become clear that there is a common mechanism by which cells
respond to various stresses. This mechanism has been coined the integrated stress response
(ISR), also referred to as the endoplasmic reticulum (ER) stress response and the unfolded
protein response (UPR). Although this response has been linked to misfolded or misprocessed
proteins (Shen et al., 2004), it is well characterized as part of the cellular response to multiple
stresses including oxidative stress, altered lipid metabolism, viral infection, hypoxia, and
nutrient deprivation, several of which are triggered in AD, PD, ALS and HAD (Yu et al.,
1999; Barbosa-Tessmann et al., 2000; Kudo et al., 2002; Ryu et al., 2002; Ledoux et al.,
2003; Katayama et al., 2004; Isler et al., 2005; Mao et al., 2005; Xue et al., 2005; Blais et al.,
2006; Fels and Koumenis, 2006; Kikuchi et al., 2006; Wei et al., 2006).
This coordinated response to stress is currently characterized by three pathways, each of which
is regulated by its own initiator that is activated upon insult exposure. Specifically, the three
ISR initiators are pancreatic ER kinase (PERK), inositol require enzyme 1α (IRE1α), and
activating transcription factor 6 (ATF6) (Figure 1). In the endoplasmic reticulum, PERK,
IRE1α, and ATF6 are kept in inactive forms through complex formation with the chaperone
protein BiP (binding protein, also known as GRP78) (Bertolotti et al., 2000;Okamura et al.,
2000;Shen et al., 2004). Induction of stress disrupts interaction between BiP and the three ISR
regulators (Hurtley and Helenius, 1989) permitting activation of these key regulators. When
free of BiP, ATF6 migrates to the transgolgi where the 50 kD N-terminal cytosolic fragment
is released by proteolysis, which then translocates to the nucleus and activates expression of
genes containing the ER stress response element (ERSE) in their promoter regions (Haze et
al., 1999;Li et al., 2000;Ye et al., 2000;Yoshida et al., 2001;Lee et al., 2002;Shen et al.,
2004). Disruption of BiP:IRE1α interaction permits dimerization of IRE1α, a serine/threonine
kinase with riboendonuclease activity (Nikawa and Yamashita, 1992;Cox et al., 1993). When
dimerized, IRE1α autophosphorylates, activating a splicing event that removes an inactivating
intron from the unspliced XBP1 (uXBP1) mRNA (Kawahara et al., 1997;Sidrauski and Walter,
1997). Removal of this intron permits efficient translation of the spliced XBP1 (sXBP1)
protein, which also activates the ERSE. Finally, disruption of BiP and PERK interactions
allows multimerization of PERK, triggering its kinase activity, thereby resulting in
autophosphorylation and phosphorylation of eukaryotic initiation factor 2 α (eIF2α) (Shi et al.,
1998;Harding et al., 1999;Harding et al., 2000).
Of specific relevance for the PERK pathway, phosphorylation of eIF2α leads to inhibition of
translation initiation for a majority of cellular proteins (Shi et al., 1998; Harding et al., 1999;
Harding et al., 2000) while enhancing translation of select proteins whose functions are to
facilitate reestablishment of cellular homeostasis (Fawcett et al., 1999; Harding et al., 2000).
eIF2α is a GTP binding protein that delivers the initiator methionine-tRNA to the 40S ribosome
(Reviewed in (Brostrom and Brostrom, 1998)). During translation initiation, the eIF2α-bound
GTP is hydrolyzed to GDP. Phosphorylation of eIF2α stabilizes the interaction between
eIF2α/GDP/eIF2 preventing exchange of the GDP for GTP. Thus, when eIF2α is
phosphorylated, translation re-initiation is favored over new initiation, increasing translation
of mRNAs with multiple micro open reading frames (called μORF) in their 5 coding regions.
Among the mRNAs translated is the activating transcription factor 4 (ATF4) transcript whose
gene product also transactivates expression of the ERSE (Fawcett et al., 1999; Harding et al.,
2000). The ERSE is found in the promoter of chaperone genes including BiP and an apoptotic
regulator, CHOP (CCAAT enhancer binding protein homologous protein; reviewed in
(Kaufman, 1999)). Activation of ERSE genes, including BiP and CHOP, are blunted in
PERK/ cells suggesting that PERK activity contributes significantly to ISR activation
(Harding et al., 2000; Novoa et al., 2001).
Markowitz et al. Page 3
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
In addition to inhibition of translation and activation of the ERSE, PERK can modify gene
expression through phosphorylation of the Nrf2 transcription factor (Nuclear factor E2-related
factor 2), a key regulator of redox homeostasis (Cullinan et al., 2003). In astrocytes, Nrf2 has
been shown to upregulate expression of antioxidant genes in response to toxins and is required
for protection against oxidative stress (Johnson et al., 2002; Lee et al., 2003; Kraft et al.,
2004; Qiang et al., 2004). Nrf2 is normally sequestered in the cytoplasm, but in response to
oxidative stress and/or phosphorylation by PERK, Nrf2 translocates to the nucleus and
activates expression of detoxifying enzymes and antioxidants (Itoh et al., 1999). Thus, PERK
contributes to the activation of the antioxidant pathway (Cullinan and Diehl, 2004).
Antioxidant Pathway of the ISR: Nrf2
When triggered, the antioxidant response can neutralize free radicals, preventing the
deleterious effects these radicals have on cells such as direct oxidation of bases within DNA,
reaction with unsaturated lipids leading to cellular membrane breakdown, or oxidative
modification of amino acids. The enzymes responsible for the endogenous antioxidant response
are regulated predominantly by transcriptional activation via Nrf2 (Itoh et al., 1997; Kwak et
al., 2001; McMahon et al., 2001). The Nrf2 transcription factor is normally bound in the
cytoplasm by the Kelch-like ECH-associated protein 1(Keap1) (Itoh et al., 1999). In response
to oxidative stress, sulfhydryl groups on Keap1 become oxidized releasing Nrf2 permitting its
translocation into the nucleus (Dinkova-Kostova et al., 2002). Release of Nrf2 from Keap1 is
regulated, at least in part, by protein kinase C phosphorylation (Huang et al., 2000, 2002;
Bloom and Jaiswal, 2003; Numazawa et al., 2003). These findings implicate Keap1 as a
“sensor” of cellular stress that can be activated by oxidative stress and/or activation of second
messenger cascades.
Upon translocation to the nucleus, Nrf2 increases transcription of genes containing the cis-
acting regulatory antioxidant response element (ARE; also referred to as the electrophilic
response element, EpRE) in the promoter region with core sequence 5-GTGACnnnGC-3
(Johnson et al., 2002; van Muiswinkel and Kuiperij, 2005). Such ARE-containing genes
include antioxidant response enzymes and phase II detoxifying genes which play an essential
role in protecting cells against oxidative stress by facilitating the inactivation and elimination
of the original insult (Johnson et al., 2002). Among the gene products regulated at the
transcriptional level by Nrf2 are the enzymes for glutathione synthesis, an essential antioxidant
in the brain. Among the enzymes contributing to glutathione synthesis is the xCT subunit of
the glutamate/cystine antiporter, which also contributes to glutamate homeostasis.
Studies have suggested Nrf2 regulation has importance in various neurodegenerative
conditions. In pharmacologically induced Parkinsonian mice, induction of Nrf2 can alleviate
symptoms (Burton et al., 2006). These findings correspond with work that has demonstrated
that downregulation of DJ-1, a PD-associated protein, leads to decreased Nrf2 by affecting its
stabilization with Keap1 (Clements et al., 2006). In AD, brains from patients with the disease
or with the Lewy body variant exhibit decreased nuclear Nrf2 (Ramsey et al., 2007), implicating
an inability of these cells to increase expression of Nrf2 target genes; similar findings have
been found in ALS (Kirby et al., 2005). As Nrf2 regulates the expression of a host of genes,
these findings encourage further exploration into the regulation of Nrf2 and role of genes under
its control.
Maintenance of Glutamate Homeostasis by Astrocytes
Glutamate acts as a major excitatory neurotransmitter in the central nervous system (CNS) and
must therefore be carefully regulated at the synaptic cleft. One regulatory mechanism is the
glutamate-glutamine cycle. In this process, astrocytes rapidly uptake glutamate from the
extracellular space and convert it into glutamine by glutamine synthetase (Norenberg and
Markowitz et al. Page 4
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Martinez-Hernandez, 1979). Glutamine is released from astrocytes and transferred to nerve
terminals where it is taken up into neurons and converted into glutamate by glutaminase.
Interestingly, this glutamate can be decarboxylated by glutamic acid decarboxylase (GAD)
which will generate GABA, an inhibitory transmitter (Battaglioli and Martin, 1991; Wu et al.,
2007). Alternatively, glutamate released into the synaptic cleft can diffuse to post-synaptic
elements where it can act on a variety of glutamatergic receptors. The two main classes of these
receptors are the ionotropic receptors (NMDA and non-NMDA subtypes) and the metabotropic
glutamatergic receptors (mGluR). Ionotropic receptors are predominately involved in rapid
signaling, whereas mGluRs are involved in more sustained responses. As the glutamate
network relies on fast signaling and most receptors are prone to rapid desensitization, excess
extracellular glutamate can have deleterious effects including induction of NMDA-receptor-
mediated excitotoxicity which can cause cell death. Though some agents, such as glutamate
pyruvate transaminase (GPT), can effectively degrade glutamate (Matthews et al., 2000),
studies have indicated that astrocytes act as the primary endogenous regulators of extracellular
glutamate levels in the CNS (Rothstein et al., 1996; Peghini et al., 1997; Tanaka et al., 1997).
Astrocytes scavenge excess glutamate to prevent neuronal excitotoxic death predominantly
through two separate transport systems, the XAG- system comprised of the excitatory amino
acid transporters (EAATs1-5) and system xc-, consisting of the cystine/glutamate antiporter
(xCT antiporter).
System XAG-EAAT1-5, along with the neutral amino acid transporters ASCT1 and 2, are considered
members of the soluble carrier gene (SLC) series within the SLC1 family as assigned by the
Human Genome Organization Nomenclature Committee. EAAT-1 and EAAT-2 mediate
glutamate transport in mature astrocytes, whereas EAAT-3 and EAAT-4 are found
predominantly in mature neurons (Palacin et al., 1998; McBean, 2002). Recent work has
suggested that there may be levels of EAAT-2 expression in macrophages (Rimaniol et al.,
2000). EAAT-5 is primarily located in retina (Arriza et al., 1994; Palacin et al., 1998).
Expression of these transporters is developmentally regulated as EAAT-1 and –2 are expressed
in immature neurons (Meaney et al., 1998; Yamada et al., 1998) and EAAT-3 and 4 are found
in fetal astrocytes (Hu et al., 2003; Maragakis et al., 2003). There is strong evidence to support
regional variation with EAAT-1 expression predominately in the cerebellum (Arriza et al.,
1994) and high EAAT-2 mRNA expression in astrocytes of the cerebral cortex (Su et al.,
2003). EAAT-1 and 3 mRNA is also expressed in non-nervous tissue. EAAT-1 mRNA has
high levels of expression in the heart and skeletal muscle, as well as expression in the placenta
and lung; EAAT-3 mRNA is in abundance in the kidney, has lower levels in the brain, placenta,
lung, and has very low levels in muscle, liver, and heart (Arriza et al., 1994). These expression
patterns in human tissue differ from those in rat tissue even though the rat transporters have
high homology to their human counterparts (Velaz-Faircloth et al., 1996).
EAATs 1–3 are sodium cotransporters and potassium countertransporters that have a high
affinity for L-glutamate (Palacin et al., 1998). Based on mammalian cell assays, EAATs have
an affinity for L-glutamate varying from 48 to 97μM (Arriza et al., 1994), though in HeLa
cells, oocytes, or rat brain, EAAT-2 has demonstrated higher affinity for glutamate with Km
values between 2 and 18μM (Pines et al., 1992; Arriza et al., 1994; Palacin et al., 1998). These
affinity values differ from the millimolar values more typical of amino acid transporters of the
epithelium. During transport, one glutamate molecule is coupled with the cotransport of three
Na+ and one H+ molecules and countertransport of one K+ molecule with the net result of an
inward coupled charged movement of +2 (Zerangue and Kavanaugh, 1996; Levy et al.,
1998). These transporters are trimeric proteins consisting of three identical subunits. Though
still controversial, crystallographic structures of a glutamate transporter from Pyrococcus
horikoshii, which has 37% amino acid identity to human EAAT-2, indicate that each individual
Markowitz et al. Page 5
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
subunit has a glutamate binding site (Yernool et al., 2004), supporting the hypothesis of an
individual anion pore for each subunit. Recently, sodium-driven glutamate uptake currents
have been shown to be independently mediated by each subunit, signifying an active anion
conductance even at low glutamate concentrations (Koch et al., 2007).
Expression and localization of EAATs are independently regulated. Epidermal growth factor
(EGF), transforming growth factor α (TGF-α), dibutyryl cAMP (dBcAMP), and bromo-cAMP
all increased EAAT-2 transcription in human fetal astrocytes (HFA) whereas tumor necrosis
factor α (TNFα) decreased expression in HFA (Fine et al., 1996; Su et al., 2003). Recent
evidence further suggests that EAAT-2 is highly regulated at the translational level by showing
that some EAAT-2 mRNAs have long 5-untranslated regions (UTR) that, in primary astrocyte
cell lines, are translationally regulated by extracellular factors such as corticosterone and retinol
(Tian et al., 2007). EAAT-3 localization is highly regulated by neuronal activity (Swanson et
al., 1997) and intracellular signaling pathways involving α protein kinase C (α PKC) (Guillet
et al., 2005) and phosphatidylinositol-3-kinase (PI3K) (Krizman-Genda et al., 2005) as shown
by studies on the rat homolog. Astrocyte-secreted factors have also been shown to affect
localization of EAAT-3 (Canolle et al., 2004).
Glutamate-induced neuronal death has been implicated as the cause of neurodegeneration in a
host of diseases, with evidence that altered EAAT regulation contributes to the pathology.
Neuronal death characterized in HIV-associated dementia (HAD), an infliction that affects 10–
20% of HIV-1 infected individuals (Gonzalez-Scarano and Martin-Garcia, 2005), has been
linked to glutamate-induced excitotoxicity (Nishida et al., 1996; O’Donnell et al., 2006). When
exposed to HIV-related factors in vitro, astrocytes exhibit decreased EAAT-2 expression and
activity with EAAT-2 mRNA levels decreasing by approximately 66% (Su et al., 2003) and
glutamate uptake decreasing by an average of 76% after 6 hours of HIV-1 exposure(Wang et
al., 2003). Similar effects on EAAT expression and activity were found in astrocytes exposed
to gp120, the HIV-1 envelope glycoprotein, and TNFα, a cytokine released by HIV-infected
macrophages and microglia in the CNS (Pitt et al., 2003; Wang et al., 2003). These data suggest
that changes in transcriptional regulation of EAAT-2 may contribute to the increases of
extracellular glutamate levels seen in HAD. This mechanism is further corroborated by
evidence that interruption of proper glutamate regulation following treatment with gp120
causes memory impairments (Fernandes et al., 2006), a common clinical symptom in HAD.
As the neurodegeneration characterized in HAD is an indirect effect of HIV-infected
macrophages and microglia that have an activated inflammatory response, it is possible that
the changes in glutamate receptors associated with HAD will be translatable to other
neuroinflammatory conditions.
Aberrant EAAT expression has been implicated in other neurodegenerative diseases. Knocking
out the mouse homolog of EAAT-2 leads to lethal spontaneous seizures, neurodegeneration in
the hippocampus, and increased susceptibility to acute cortical injury (Tanaka et al., 1997).
Specifically, Tanaka et al (1997) found that knockout mice had spontaneous seizures that
resembled NMDA-induced seizures and premature death. In concordance with these findings,
the sporadic form of ALS, characterized by loss of motor neurons (Wisman et al., 2003) and
by increased glutamate levels in cerebrospinal fluid (CSF) (Spreux-Varoquaux et al., 2002),
has up to 95% reduction in EAAT-2 expression in motor cortex and spinal cord (Howland et
al., 2002). One hypothesis is that this EAAT reduction is due to RNA processing (Lin et al.,
1998), but more recent evidence implicates translation attenuation (Tian et al., 2007), a state
that can be induced by ISR as well. In AD, the hippocampus and gyrus frontalis medialis have
decreased EAAT-2 expression, especially around the amyloid plaques, whereas the cerebellum
has upregulated EAAT expression (Jacob et al., 2007). However, the transgenic APP/Ld/2
mouse, which overexpresses a 695 amino acid form of human amyloid precursor protein (APP),
Markowitz et al. Page 6
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
exhibits a decrease in glial-specific glutamate transporter activity that corresponds with
decreased protein expression but not mRNA levels (Masliah et al., 2000).
Neuronal EAAT-3 has been similarly studied in seizures as well as in transient ischemia. In
patients with temporal lobe epilepsy with hippocampal sclerosis, though there were fewer total
EAAT-3 positive cells, EAAT-3 expression was increased in individual neurons in CA2 and
in the granule cell layer of the dentate gyrus compared with patients without hippocampal
sclerosis (Mathern et al., 1999; Proper et al., 2002). Studies have also looked at EAAT-3
regulation after short ischemic events that act as a form of preconditioning, resulting in
resistance to damage after subsequent severe ischemic injury. After induction of a short cerebral
ischemic event in rats, there is neuronal upregulation of EAAT-3 with particular intensity at
the plasma membrane; this upregulation appears to be regulated by TNFα and the TNFα
receptor subfamily member 1 (TNFR1) (Pradillo et al., 2006). Though disease or injury to the
CNS seems to affect EAAT-2 and EAAT-3 expression similarly, in contrast to EAAT-2,
EAAT-3 does not demonstrate changes in the frontal cortex of AD patients by
immunoreactivity or at the mRNA level (Li et al., 1997). As damage to the frontal cortex
underlies the symptomatology of AD, in regards to glutamate transport, it is the affects of
EAAT-2 regulation that might be important in the neuropathogenesis of this disease.
As the structure and kinetics of EAATs continue to be elucidated and their role in disease
explored, there has been a great push to discover pharmacological agents that alter function of
these glutamate transporters with a goal of developing therapeutics (reviewed in (Bridges and
Esslinger, 2005)). Most inhibitors to date that show selectivity are limited to EAAT-2 inhibitors
(Vandenberg et al., 1997; Eliasof et al., 2001; Shimamoto et al., 2004; Moussa et al., 2007).
However, as the structural differences between the individual EAATs are further elucidated,
it may be easier to exploit these differences to increase selectivity. More importantly,
understanding the structure of these receptors may aid in the development of functional
enhancers which may be needed to compensate for the lost EAAT activity observed in
neurodegenerative diseases.
System xcThe other major glutamate transport system in astrocytes, the xc- system, also referred to as
the cystine/glutamate transporter and as the xCT antiporter, mediates over 50% of glutamate
transport in glioma cell lines (Ye et al., 1999). It acts as an antiporter by taking up cystine in
exchange for glutamate export. By blocking glutamate release from this antiporter, the rat
striatum has a 60% decrease in extrasynaptic glutamate levels, implicating that the origin of
this glutamate is from non-vesicular release, and that system xc- is a major source (Baker et
al., 2002). Studies on cocaine withdrawal in rats have also implicated system xc- as an important
regulator of extracellular glutamate (Baker et al., 2003). System xc- also has a vital role in
cellular protection against oxidative stress (Lewerenz et al., 2006; Shih et al., 2006), as the
import of cystine is vital to the synthesis of glutathione. As the antiporter plays an essential
for glutathione production in exchange for increasing extracellular glutamate, its regulation
must be carefully controlled. We believe that aberrant expression and/or activity of system
xc- may contribute to excitotoxicity during neurodegeneration
The xc- system is a highly regulated electroneutral system that exchanges cystine for glutamate
in a Na+-independent 1:1 ratio (Palacin et al., 1998; Sato et al., 1999). The antiporter is
composed of two subunits: a heavy chain comprised of a cell surface antigen 4F2hc that is
necessary for antiporter translocation to the plasma membrane (Bassi et al., 2001), and a light
chain known as xCT that has two variants formed by alternative splicing (McBean, 2002),
either of which can confer substrate specificity. In humans, the xCT light chain has been termed
SLC7A11 by the Human Genome Organization Nomenclature Committee. In mice, xCT is a
502 amino acid protein with 12 assumed transmembrane domains (Sato et al., 1999; Gasol et
Markowitz et al. Page 7
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
al., 2004). Though the predicted size of the xCT protein is approximately 55kDa (Sato et al.,
1999), research in mice has found that an xCT antiserum recognizes 40kDa and 80kDa bands
that correspond to the xCT protein (Burdo et al., 2006; La Bella et al., 2007). Northern blot
analysis has revealed xCT mRNA in mouse brain, rat cortical cells, retinal ganglion cells, and
human glioma cells (Sato et al., 2002; Burdo et al., 2006; Dun et al., 2006), but not in mouse
heart, kidney, liver, or lung (Sato et al., 1999). Immunocytochemical studies in rat brain have
found strong xCT expression in all cell types examined, with increasing expression throughout
development (La Bella et al., 2007). In the adult mouse brain, xCT mRNA has robust expression
in areas facing the cerebral ventricles, medial habenuclei, and in the meninges, scattered
expression in the hypothalamic regions, and low level expression in the cortex and cerebellum
(Sato et al., 2002).
Within the 5-flanking region of the xCT gene are several sequences resembling the cis-acting
electrophilic response element (EpRE), also known as the antioxidant response element (ARE),
that is responsible for the induction of the adjacent genes. As such, xCT is transcriptionally
regulated by electrophilic agents such as diethylmaleate, and hydroquione by activation of the
endogenous antioxidant response via the Nrf2 transcription factor (Sasaki et al., 2002). During
oxidative stress or other events that cause the activation of the cellular integrated stress response
(Cullinan et al., 2003), Nrf2 translocates to the nucleus where, in the presence of small Maf
proteins, it binds the ARE and up-regulates transcription of these ARE-containing genes. Nrf2
has been found to regulate the xCT antiporter in astrocytes (Qiang et al., 2004), with
overexpression of Nrf2 in rat cortical cultures leading to increased xCT mRNA expression,
xCT activity, and intracellular glutathione levels (Shih et al., 2003). Though Nrf2 has been
found to control the glutathione pathway (Sun et al., 2005), Nrf2 is not the only regulator of
xCT, though other methods of regulation have yet to be determined (Pacchioni et al., 2007).
One recently suggested mechanism is through the ISR-regulated transcription factor, ATF4,
which seems to be important in controlling xCT expression during amino acid deprivation
(Sato et al., 2004). As neurodegenerative conditions can induce oxidative stress, upregulation
of the antioxidant response, including the xCT antiporter, can act as a protective mechanism
by preventing damage from oxidative stress. In HIV-associated dementia (HAD), studies have
found the exposure of retinal pigment epithelial to the transactivator protein, TAT, induces
xCT expression (Bridges et al., 2004). As the most abundant source of oxidative stress in the
brain is from the respiratory burst of activated microglia (Colton and Gilbert, 1987; Colton et
al., 2000), a major pathological hallmark in HAD, it is not surprising to see changes in the
antioxidant pathway and, therefore, in xCT in this disease.
Though important in glutathione synthesis, upregulation of the xCT antiporter can have
potentially damaging effects, as it is a mechanism for glutamate export and critical in the
regulation of extracellular glutamate (Augustin et al., 2007). The pathology behind a variety
of neurological insults takes advantage of the xCT antiporter to propagate cell damage. In glial-
derived tumors, or gliomas, high levels of glutamate are released, causing excitotoxic neuronal
death, mediated by increased xCT expression (Chung et al., 2005). When pharmacological
inhibitors are used to block system xc- activity, the progression of tumor growth is disrupted
(Chung et al., 2005). Kaposi’s sarcoma-associated herpesvirus (KSHV), a virus linked to
lymphoproliferative syndromes that frequently affects immunocompromised individuals, also
uses the xCT antiporter, as KSHV cell fusion and virion entry correlates closely with expression
of the xCT light chain of the antiporter (Kaleeba and Berger, 2006). In AD, inhibition of system
xc- may also prove advantageous for neurons. In neuronal cultures exposed to microglia,
exposure to amyloid β-peptides (Aβ) increases microglial xCT mRNA expression and
glutamate release which can trigger neuronal death; if system xc- is inhibited, microglia can
increase release of apolipoprotein E rescuing neurons from Aβ toxicity (Qin et al., 2006),
indicating a potential deleterious role of system xc-. Under these situations, attenuating xCT
antiporter activity would be expected to have beneficial outcomes. Thus, regulation of system
Markowitz et al. Page 8
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
xc- requires a delicate balance, ensuring adequate cystine for glutathione synthesis, but
minimizing glutamate export.
Regulation of Glutathione by Astrocytes
Oxidative stress can lead to endoplasmic reticulum (ER) stress as the ER attempts to handle
the accumulation of oxidized proteins. Cystine is one of the required components in the
synthesis of glutathione (GSH), an important antioxidant of the CNS. In exchange for
glutamate, cystine is transported into the cell through the xCT antiporter. Once in the cell,
cystine is reduced to cysteine. A reaction between cysteine and glutamic acid results in γ-
glutamyl-cysteine, which in turns reacts with glycine to form GSH (Peuchen et al., 1997). As
cysteine acts as the rate limiting element in GSH synthesis, the xCT antiporter is crucial for
maintaining GSH homeostasis (Dun et al., 2006), and is critically important in protecting cells
from free radicals, peroxides, and other oxidizing agents (O’Connor et al., 1995).
Glutathione is heterogeneously distributed throughout the nervous system, potentially
affecting region-and cell-specific capacities to respond to oxidative stress (Philbert et al.,
1991). In the central nervous system, GSH is compartmentalized between glia and neurons,
with glial cells exhibiting higher levels of GSH (Knollema et al., 1996), especially microglia
(Chatterjee et al., 1999). GSH released from astrocytes provides the precursors for neuronal
GSH synthesis. GSH acts as an antioxidant by either catalyzing the nucleophilic addition of
the sulfur atom in glutathione to an electrophilic group, or by covalently binding with reactive
metabolites (Ahlgren-Beckendorf et al., 1999), upon which it is oxidized to glutathione
disulfide (GSSG). Therefore, not only is GSH an essential antioxidant, it also participates in a
variety of cell processes such as cell proliferation and regulation of apoptosis, detoxification
of xenobiotics, thiol redox state, and storage and transport of cysteine, among others (reviewed
in ((Cotgreave and Gerdes, 1998; Arrigo, 1999; Dringen, 2000)).
Altered glutathione levels are characterized in a variety of conditions from PD (Sofic et al.,
1992) to schizophrenia (Yao et al., 2006) as well as in normal aging (Sohal and Weindruch,
1996). In HIV patients, glutathione levels are altered by viral proteins (Price et al., 2005) and
act as a key determinant of patient survival, with decreased GSH associated with a markedly
decreased survival rate (Herzenberg et al., 1997); administration of the GSH pro-drug N-
acetylcysteine (NAC) replenishes GSH, identifying itself as an important potential therapeutic
(De Rosa et al., 2000; Muller et al., 2000). Targeting the consequences of glutathione depletion
has proved an effective method to attenuate neuronal loss in models of CNS disorders (Malorni
et al., 1998; Anderson et al., 2004; Hart et al., 2004).
As more glutathione is needed, logic stands that xCT activity and/or expression would need to
be augmented in order to increase glutathione synthesis. However, as upregulation of this
antiporter mediates the regulation of extracellular glutamate (Augustin et al., 2007), careful
regulation of system xc- is needed in order to maintain a balance between protective glutathione
synthesis and toxic increases in extracellular glutamate.
Neuronal Responses to Altered Astrocyte Function
NMDA Receptor Stimulation
Increased release of glutamate from astrocytes mediates their neurotoxic activities through the
NMDA receptors. Functional NMDA receptors are heteromeric assemblies of four subunits:
two NMDA-R1 subunits and two NMDA-R2. Together, the four subunits form a channel in
the plasma membrane permitting influx of Ca2+ when stimulated. The NMDA-R1 subunits
(NR1) are derived from a single gene that can exist as eight different variants based on
alternative splicing (Goebel et al., 2005), while four separate genes encode the NMDA-R2
subunits (NR2A, NR2B, NR2C and NR2D) (Lynch and Guttmann, 2001; Waxman and Lynch,
Markowitz et al. Page 9
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
2005). The NR1 subunit binds glycine, while the NR2 subunit binds glutamate and quinolinic
acid. Although all four NR2 subunits bind glutamate with equal affinity, NR2A and NR2B
trigger greater excitotoxicity than NR2C and NR2D (reviewed in (Lynch and Guttmann,
2001; Waxman and Lynch, 2005). Quinolinic acid activates NR2A- and NR2B-containing
receptors, but not those containing NR2C or NR2D subunits (Carvalho et al., 1996). NR2B-
containing NMDA receptors are often located extrasynaptically (Charton et al., 1999; Ali and
Salter, 2001; Riccio and Ginty, 2002). Interestingly, brain regions enriched in NMDA receptors
composed of NR2A and NR2B subunits, which include the hippocampus, striatum, forebrain,
are commonly injured in excitotoxic insults such as ischemia, epilepsy and also in HIV-1
infection (Conti et al., 1999; Everall et al., 1999; Heyes et al., 2001; Lynch and Guttmann,
2002; Archibald et al., 2004; Sa et al., 2004; Waxman and Lynch, 2005). Furthermore, regions
such as the cerebellum that express relatively high levels of NR2C and low levels of NR2B
are commonly spared in excitotoxic insults (Lynch and Guttmann, 2002).
There are two inter-related levels of regulation of NMDA receptor expression and function
that ultimately influence their roles in neuronal survival and death: NR subunit expression by
developmental stage and phosphorylation state. A major mechanism of regulation of the
NMDA receptor is the phosphorylation/de-phosphorylation by neuronal kinases and
phosphatases. The phosphorylation state of NR1 and NR2 NMDA receptor subunits regulates
their function and can also influence their localization to synaptic and extra-synaptic sites by
regulating associations with cellular proteins. NR1 subunits are typically phosphorylated on
serine and threonine residues (ser896, ser897) while NR2A and NR2B are typically
phosphorylated on tyrosine residues (tyr842, tyr1387, tyr1472), as well as at other residues
(Kennedy and Manzerra, 2001; Li et al., 2001; Chung et al., 2004; Besshoh et al., 2005). NMDA
receptor function is modulated by tyrosine kinases and phosphatases, as well as serine/
threonine kinases, although how this modulation is achieved is not clear (Kennedy and
Manzerra, 2001; Li et al., 2001; Goebel et al., 2005). In NR1, phosphorylation of ser896 and
ser897 is required for PKC-mediated release of NR1-containing NMDA receptors from the
ER and may play a role in NMDA receptor membrane trafficking (Scott et al., 2001). Brain-
derived neurotrophic factor (BDNF) causes a rapid phosphorylation of NR1 that is associated
with potentiation of synaptic transmission (Suen et al., 1997).
In NR2 subunits, the C-terminal tails of both NR2A and NR2B contain three tyrosines located
within YXXØ motifs (NR2A: tyr 842, tyr1325, tyr 1454, and NR2B tyr843, tyr1336, and
tyr1472). Phosphorylation of tyr1472 in NR2B located within the YXXØ motif correlates with
amyloid-β-induced internalization of the NMDA receptor in cortical neurons (Snyder et al.,
2005), while inhibition of tyr842 phosphorylation on NR2A is associated with decreased
membrane expression (Vissel et al., 2001). Notably, ischemic insult to neurons is associated
with increased NR1 (ser896), NR2A (tyr1387) and NR2B phosphorylation, as well as with
enhanced association of these receptors within the postsynaptic density (Cheung et al., 2003;
Besshoh et al., 2005). The net effect is the increased synaptic expression of NR2A– and NR2B-
containing NMDA receptors. However, loss of tyr1472 phosphorylation in NR2B has been
associated with stabilization of NR2B on the cell surface, as well as with increased
internalization through endocytosis via clathrin coated pits (Roche et al., 2001; Snyder et al.,
2005). It thus remains to be determined how tyr1472 phosphorylation affects hippocampal
neuronal sensitivity to excitotoxicity. Recently, cdk5-dependent phosphorylation of NR2A
ser1232 in vitro and in vivo has been shown to regulate NMDA-induced long-term potentiation
(LTP) in CA1 hippocampal neurons (Kennedy and Manzerra, 2001; Li et al., 2001).
Furthermore, enhanced phosphorylation of NR2B tyr1336 might be associated with activation
of neuronal survival pathways (Wu et al. J. Biol. Chem. 13: 20075–20087).
Another important consideration regarding the role of NR2 subunits in excitotoxic death is
their site of localization. Generally, extrasynaptic NMDA receptors primarily containing
Markowitz et al. Page 10
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
NR2B subunits are thought to mediate excitotoxicity, whereas synaptically-associated NMDA
receptors mainly containing NR2A subunits at ages when both subunits are expressed, are
associated with neuroprotective functions, synaptic plasticity, and long-term potentiation
(Rumbaugh and Vicini, 1999; Tovar and Westbrook, 1999; Sattler and Tymianski, 2001;
Hardingham and Bading, 2003). Synaptic and extrasynaptic localization of NMDA receptors
are differentially regulated by their phosphorylation state (Li et al., 2002), and by their
association with post-synaptic density proteins (PSD-95, SAP 102) (Dong et al., 2004). Thus,
specific phosphorylation events within the C-terminal tails of NR subunits provide a major
level of regulation of neuronal entry into both survival and death pathways.
A second major level of regulation of NMDA receptor function is the developmentally-linked
expression of NR subunits (Liu et al. J. Neuroscience 27: 2846–2857). NR2A and NR2B-
containing receptors demonstrate varying temporal expression in the rat hippocampus and the
expression coincides with hippocampal neuronal susceptibility to age-dependent excitotoxicity
in maturing cultures (O’Donnell et al., 2006); mouse cortical cultures demonstrate a similar
age-dependent effect in regards to glutamate toxicity (Fogal et al., 2005). These observations
suggest a common pathway of neuronal death mediated by NR2A and/or NR2B subunits in
classic excitotoxic brain injury, including HIV-1 infection. Accordingly, the development of
NR2A- (Siao and Tsirka, 2002) and NR2B-selective antagonists (Lynch and Guttmann,
2001) has been driven by a widespread interest in producing non-toxic, selective NMDA
receptor antagonists as neuroprotectants against such brain injury (Lipton, 2004). Most NMDA
receptor-mediated toxicity is regulated through calpain activation, hence, selective NR subunit
inhibitors and calpain antagonists can provide neuroprotection.
Activation of Calpains in Neurons
Excitotoxicity due to overstimulation of NMDA receptors results in increased cytosolic
calcium concentrations ([Ca2+]). A family of calcium-activated cysteine proteases, the
calpains, has been implicated in modulating cellular mechanisms in response to changes in
[Ca2+] (Guroff, 1964; Wang et al., 1992). Calpains are upregulated in a variety of
neurodegenerative conditions, both acute, such as traumatic brain injury (McCracken et al.,
1999) and brain ischemia (Yamashima, 2000; Blomgren et al., 2001; Shioda et al., 2006), and
chronic, such as multiple sclerosis (Shields et al., 1999), AD (Grynspan et al., 1997; Tsuji et
al., 1998; Raynaud and Marcilhac, 2006), Huntington’s disease (Lee et al., 2006), and HAD
(Vanderklish and Bahr, 2000; Ray and Banik, 2003; Wang et al., 2007). Calpains are thought
mainly to function intracellularly and have been found in virtually all vertebrate cells, although
calpains or calpain-like genes have been identified in Drosophila and other invertebrates (Goll
et al., 2003). There are 14 identified human genes in the mammalian calpain family encoding
the 80kDa cysteine protease domains and two encoding the small 30kDa regulatory subunit
domain (Suzuki et al., 2004). The mammalian calpain genes share over 90% homology and
there is little evidence of alternative splicing events (Suzuki, 1990). Though calpain substrate
selectivity is not completely understood and may depend on the target peptide’s tertiary
structure, some substrates can be distinguished by a sequence rich in proline, glutamate, and
serine/threonine residues (Rechsteiner and Rogers, 1996; Tompa et al., 2004).
Two calpain isoforms predominate, μ–calpain and m-calpain (Croall and DeMartino, 1991).
Both isoforms share substrate specificities (Molinari and Carafoli, 1997) and are heterodimers
composed of a regulatory 28 kDa subunit that is identical between the two isoforms and an 80-
kDa subunit that shares 55–65% sequence homology (Goll et al., 2003). Although μ–calpains
have been described more prevalently in neurons and m-calpains in glia (Hamakubo et al.,
1986), they are classically distinguished by the in vitro [Ca2+] requirement for their half
maximal activation: μ–calpain (3–50 μM) and m-calpain (400–800 μM) (Suzuki, 1991; Ray
et al., 1999; Goll et al., 2003). Since physiologic [Ca2+] generally range from 50–300nM, and
Markowitz et al. Page 11
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
[Ca2+] necessary to activate m-calpain would result in immediate cell death regardless of
calpain activation, calpain must utilize mechanisms that reduce their required [Ca2+]. In the
presence of Ca2+, μ–calpains and m-calpains both undergo autoproteolysis that reduces their
[Ca2+] activation requirement by over half (Saido et al., 1994; Goll et al., 2003). In addition,
phospholipids, such as phosphatidylinositol 4,5-bisphosphate (PIP2), can also further reduce
the [Ca2+] activation requirement (Saido et al., 1992). Phosphorylation of the calpains may
also play a role in determining their activation (Goll et al., 2003). ERK phosphorylation of
Ser50 of m-calpain can reduce its activation threshold to under 1μM Ca2+ (Glading et al.,
2002). These steps may be necessary for calpain activation under physiologic conditions.
Once activated, calpains are known to regulate multiple cellular mechanisms through cleavage
of their substrates. The limited number of calpain cleavage sites on a given peptide often results
in large, catalytically active fragments and suggests that the role of calpains is often regulatory,
as opposed to the degradational roles of proteases in lysomes or proteasomes (Goll et al.,
2003; Franco and Huttenlocher, 2005). Along these lines, calpains have been implicated in
long-term potentiation in the hippocampus via direct or indirect modulation of kinases,
glutamate receptors, or receptor interacting proteins (Tomimatsu et al., 2002). Further
alteration of cellular activity may occur due to proteolytic modulation of signal transduction
(Kishimoto et al., 1989; Pariat et al., 1997; Pfaff et al., 1999; Pariat et al., 2000), gene expression
and transcription factors, cell cycle machinery (Kubbutat and Vousden, 1997; Zhang et al.,
1997), and apoptosis (Squier et al., 1999; Chua et al., 2000; Lu et al., 2002).
Both apoptotic- and necrotic-like death have been described following excitotoxic stimulation
depending on the stress severity (Lieberthal et al., 1998). Calpains have a variety of roles in
regulating apoptotic mechanisms in cells, including prevention of apoptosome formation. This
action prevents the induction of the intrinsic mitochondrial apoptotic pathway despite potential
mitochondrial cytochrome c release and, therefore, may promote a more necrotic-like
mechanism of death (Lankiewicz et al., 2000). Activation of μ-calpains has also been shown
to lead to lysosomal membrane degradation and, thus, release of cathespin B and L
(Yamashima, 2000). Cathepsin could contribute to breakdown of cellular components and this
has been demonstrated in hippocampal neurons (Kitao et al., 2001). However, calpains can
also have pro-apoptotic effects due to their cleavage of Bcl-2 family proteins (Gil-Parrado et
al., 2002), apoptosis-inducing factor (Polster et al., 2005), Bax (Wood et al., 1998; Gao and
Dou, 2000; Hwang et al., 2003), and Bid (Chen et al., 2001; Chen et al., 2002; Mandic et al.,
2002; Takano et al., 2005; White et al., 2007). Furthermore, calpains can both directly activate
and deactivate caspases through cleavage (Chua et al., 2000; Nakagawa and Yuan, 2000).
Calpains play an important role in disassembling adhesion and cytoskeletal components at the
plasma membrane, which are essential for the mechanisms of cell motility (Glading et al.,
2002; Perrin and Huttenlocher, 2002). Because of this, calpains may play important roles in
neurite outgrowth or remodeling in neurons (Howard et al., 1999). Along these lines, some of
the major calpain substrates are adhesion molecules and cytoskeletal proteins, including
vimentin, spectrin, tau, actin, tubulin, glial fibrilliary acidic protein, neurofilaments, and
microtubule-associated proteins. Activation of calpains during synaptic activity is also crucial
for neurotransmitter release, synaptic plasticity, vesicular trafficking, and structural
stabilization (Wu and Lynch, 2006). Activation of calpains due to excitotoxic insult can result
in synaptic dysfunction or inappropriate disassembly of cellular structures and may contribute
to the classic morphologic impact seen following excitotoxic stress (Buki et al., 2003;
O’Hanlon et al., 2003). As such, calpains have been implicated in Wallerian degeneration
(Zhai et al., 2003), acute axonal degeneration (Kerschensteiner et al., 2005), and in the loss of
soma, neurites, and synapses in aged primary neuronal cultures (Kim et al., 2007).
Markowitz et al. Page 12
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Calpains may further exacerbate excitotoxic conditions by cleaving proteins critical to
cytosolic [Ca2+] regulations, such as the plasmalemmal Ca2+ ATPase (Guerini et al., 2003).
μ-calpains are thought to be closely linked to NMDA receptor function (Hewitt et al., 1998).
In fact, its ability to cleave the NR2 subunits (N-terminal of NR2A, NR2B, and NR2C and C-
terminal of NR2A) of the NMDA receptor may lead to modulation of excitotoxic conditions
(Bi et al., 1998; Guttmann et al., 2001; Simpkins et al., 2003). Electrophysiological properties
of the NMDA receptors are not altered by these cleavages, but these cleavages may regulate
localization, turnover, or signaling. Along these lines, the C-terminus of NR2 subunits in the
cytoplasmic side of the cellular membrane interact with various kinases and synaptic proteins,
such as PSD-95, which is involved in vesicle trafficking. Guttmann et al (2001) have
speculated, and transgenic studies have suggested (Sprengel et al., 1998), that cleavage of the
NR2 subunit results in NMDA receptors that are prevented from interacting with cytosolic
signaling pathways. In addition, it has been suggested that calpain-cleaved NR2B-containing
NMDA receptors may linger on the surface of neurons and thus increase neuronal excitotoxic
susceptibility (O’Donnell et al., 2006).
Altered cell cycle regulation is another means by which excitoxic calpain activation could
induce neuronal dysfunction. Aberrant activation of cell cycle machinery has been linked to a
variety of neurodegenerative diseases (reviewed in (Herrup and Yang, 2007)). As suggested
by Carragher et al, calpain activity may modulate progression through the G1 phase of the cell
cycle, and increased expression of the endogenous calpain inhibitor calpastatin correlates with
decreased cyclin A, D, and cdk2, as well as decreased levels of retinoblastoma protein
phosphorylation (Carragher et al., 2002). Calpains have also been implicated in the regulation
of the E2F1 transcription factor, which is known to regulate proliferation and neuronal death
(Strachan et al., 2005).
Other calpain substrates whose cleavage could alter neuronal function include ion channels
and transporters, growth factor receptors, phospholipases, kinases and phosphatases (Goll et
al., 2003). Since calpains themselves can be phosphorylated, their activation may have
autoregulatory roles. Additionally, calpains may alter neuronal function by modifying kinase
and signal transduction pathways in neurons through cyclin dependent kinase 5 (CDK5)
activation. Calpain cleavage of p35 to its more stable isoform, p25, increases the level, activity,
and subcellular distribution of the p25:CDK5 complex (Musa et al., 1998; O’Hare et al.,
2005). This complex has been associated with CDK5 neurotoxic function (Patrick et al.,
1999), and pharmacologic inhibition of calpains and CDK5 has shown to be protective in some
excitotoxicity models (Wang et al., 2007).
The most stringently specific calpain inhibitor is the endogenously expressed calpastatin,
which is a major mediator of calpains and is, itself, dependent on Ca2+ for its inhibition of
calpains (Tullio et al., 1999; Goll et al., 2003). Synthetic calpain inhibitors, on the other hand,
lack the exact calpain specificity of calpastatin, but do successfully protect neurons from
excitotoxic stress. Calpain inhibitors PD 150606 and MDL-28170 can reduce degeneration in
cultured neurons following excitotoxic or other stress stimuli (Wang et al., 2007). MDL-28170
is neuroprotective in both in vitro and in vivo models of traumatic brain injury (Wang and
Yuen, 1997; Laurer and McIntosh, 2001) and several calpain inhibitors have shown significant
neuroprotection in animal models of CNS injury and disease (Ray and Banik, 2003). However,
clinical trials involving calpain inhibitors have thus far shown little promise (Wang et al.,
2006).
Because calpains have seemingly contradictory roles in preventing and initiating apoptotic
processes, as well as in regulating critical cellular functions, it is hard to define their role as
simply neurodegenerative or neuroprotective. Indeed, the system that may be responsible for
the blebbing and disintegration of neuronal processes might also serve a critical role in
Markowitz et al. Page 13
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
remodeling cytoskeletal proteins important for the neurite regrowth that is seen following
sublethal excitotoxic insults. Most likely, the role of calpains is dependent on both the type
and severity of the stress (Moore et al., 2002). Likewise, the function of calpain activation
could be site-specific within a cell, as non-homogeneous distribution of [Ca2+] can locally
activate calpains (Valeyev et al., 2006). Alternatively, the effect of calpains could be dependent
on the isoform composition of the local calpain pool. The resultant magnitude, duration, and
localization of cytosolic [Ca2+] following an excitotoxic insult could all be critical determinants
of the role calpains will play in excitotoxic responses.
Summary and Model
Taken together, the pathways described in this review have known physiologic roles in CNS
function and response to stress. Certainly, astrocytes need to activate Nrf2 to produce
glutathione and reduce oxidative stress. However, we propose that overstimulation of
protective pathways, such as the antioxidant response and the integrated stress response, in
astrocytes during chronic, progressive exposure to neurodegenerative stimuli will result in
neuronal damage and death by altering the balance of glutamate transport. In our model (Figure
2), activation of the ISR in astrocytes will increase glutamate export by the cystine/glutamate
antiporter to increase import of cystine for glutathione production. At the same time,
inflammatory mediators such as TNFα or the ISR via altered ER and golgi processing will
repress EAAT-1 and EAAT-2 expression and activity removal of excess glutamate from the
extracellular milieu. Such changes will alter astrocytic glutamate homeostasis leading to net
extracellular accumulation of glutamate. Failure of the astrocytes to regulate glutamate will
result in stimulation of neuronal NMDA receptors, perhaps extra-synaptically, leading to
Ca+2 release and activation of calpains. Initially, activation of calpains may have limited local
effects on dendritic morphology, but overtime, its activation may lead to neuronal death. We
propose that the initial local changes in dendritic morphology would cause neuronal
dysfunction, which would manifest clinically, but would be reversible if treated. However,
chronic, progressive activation of calpains would ultimately culminate in neuronal loss making
the clinical deficits fixed. These studies suggest several therapeutic targets for patients with
neurodegenerative diseases, but also highlight the complexity of treating astrocytic
dysfunction, as reducing effects of glutamate toxicity may reduce antioxidant activity, thus
trading one mechanism of cellular toxicity for another. As we expand our understanding of the
regulatory mechanisms governing astrocytic and neuronal responses to stress, comprehensive
treatment programs for patients with neurodegenerative diseases will emerge.
References
Achim CL, Wiley CA. Inflammation in AIDS and the role of the macrophage in brain pathology. Current
Opinion in Neurology 1996;9:221–225. [PubMed: 8839616]
Ahlgren-Beckendorf JA, Reising AM, Schander MA, Herdler JA, Johnson JA. Coordination regulation
of NAD(P)H:Quinone Oxidoreductase and Glutathione-S-Transferases in primary cultures of rat
neurons and glia: Role of the antioxidant/electrophilic response element. GLIA 1999;25:121–142.
Ali DW, Salter MW. NMDA receptor regulation by Src kinase signalling in excitatory synaptic
transmission and plasticity. Curr Opin Neurobiol 2001;11:336–342. [PubMed: 11399432]
Anderson MF, Nilsson M, Eriksson PS, Sims NR. Glutathione monoethyl ester provides neuroprotection
in a rat model of stroke. Neuroscience Letters 2004;354:163–165. [PubMed: 14698463]
Annunziato L, Amoroso S, Pannaccione A, Cataldi M, Pignataro G, D’Alessio A, Sirabella R, Secondo
A, Sibaud L, Di Renzo GF. Apoptosis induced in neuronal cells by oxidative stress: role played by
caspases and intracellular calcium ions. Toxicology Letters 2003;139:125–133. [PubMed: 12628748]
Anonymous. A randomized, double-blind, placebo-controlled trial of deprenyl and thioctic acid in human
immunodeficiency virus-associated cognitive impairment. Dana Consortium on the Therapy of HIV
Markowitz et al. Page 14
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Dementia and Related Cognitive Disorders.[see comment]. Neurology 1998;50:645–651. [PubMed:
9521250]
Archibald SL, Masliah E, Fennema-Notestine C, Marcotte TD, Ellis RJ, McCutchan JA, Heaton RK,
Grant I, Mallory M, Miller A, Jernigan TL. Correlation of in vivo neuroimaging abnormalities with
postmortem human immunodeficiency virus encephalitis and dendritic loss. ArchNeurol
2004;61:369–376.
Arrigo AP. Gene expression and the thiol redox state. Free Radical Biology & Medicine 1999;27:936–
944. [PubMed: 10569626]
Arriza JL, Fairman WA, Wadiche JI, Murdoch GH, Kavanaugh MP, Amara SG. Functional comparisons
of three glutamate transporter subtypes cloned from human motor cortex. J Neurosci 1994;14:5559–
5569. [PubMed: 7521911]
Augustin H, Grosjean Y, Chen K, Sheng Q, Featherstone DE. Nonvesicular Release of Glutamate by
Glial xCT Transporters Suppresses Glutamate Receptor Clustering In Vivo. Journal of Neuroscience
2007;27:111–123. [PubMed: 17202478]
Bagetta G, Corasaniti MT, Berliocchi L, Nistico R, Giammarioli AM, Malorni W, Aloe L, Finazzi-Agro
A. Involvement of interleukin-1[beta] in the mechanism of human immunodeficiency virus type 1
(HIV-1) recombinant protein gp120-induced apoptosis in the neocortex of rat. Neuroscience
1999;89:1051–1066. [PubMed: 10362294]
Baker DA, Xi Z-X, Shen H, Swanson CJ, Kalivas PW. The Origin and Neuronal Function of In Vivo
Nonsynaptic Glutamate. Journal of Neuroscience 2002;22:9134–9141. [PubMed: 12388621]
Baker DA, McFarland K, Lake RW, Shen H, Tang X-C, Toda S, Kalivas PW. Neuroadaptions in cystine-
glutamate exchange underlie cocaine relapse. Nature Neuroscience 2003;6:743–749.
Barbosa-Tessmann IP, Chen C, Zhong C, Siu F, Schuster SM, Nick HS, Kilberg MS. Activation of the
Human Asparagine Synthetase Gene by the Amino Acid Response and the Endoplasmic Reticulum
Stress Response Pathways Occurs by Common Genomic Elements. J Biol Chem 2000;275:26976–
26985. [PubMed: 10856289]
Barzilai A, Daily D, Zilkha-Falb R, Ziv I, Offen D, Melamed E, Shirvan A. The molecular mechanisms
of dopamine toxicity. Advances in Neurology 2003;91:73–82. [PubMed: 12442665]
Bassi M, Gasol E, Manzoni M, Pineda M, Riboni M, Martin R, Zorzano A, Borsani G, Palacin M.
Identification and characterisation of human xCT that co-expresses, with 4F2 heavy chain, the amino
acid transport activity system xc- Pflugers Archiv European Journal of Physiology 2001;442:286.
[PubMed: 11417227]
Battaglioli G, Martin D. GABA synthesis in brain slices is dependent on glutamine produced in astrocytes.
Neurochemical Research 1991;16:151–156. [PubMed: 1881516]
Benarroch EE. Neuron-astrocyte interactions: partnership for normal function and disease in the central
nervous system. Mayo Clinic Proceedings 2005;80:1326–1338. [PubMed: 16212146]
Bertolotti A, Zhang Y, Hendershot LM, Harding HP, Ron D. Dynamic interaction of BiP and ER stress
transducers in the unfolded-protein response. Nature Cell Biology 2000;2:326–332.
Besshoh S, Bawa D, Teves L, Wallace MC, Gurd JW. Increased phosphorylation and redistribution of
NMDA receptors between synaptic lipid rafts and post-synaptic densities following transient global
ischemia in the rat brain. J Neurochem 2005;93:186–194. [PubMed: 15773918]
Bi X, Rong Y, Chen J, Dang S, Wang Z, Baudry M. Calpain-mediated regulation of NMDA receptor
structure and function. Brain Research 1998;790:245–253. [PubMed: 9593918]
Bibel M, Barde YA. Neurotrophins: key regulators of cell fate and cell shape in the vertebrate nervous
system. Genes & Development 2000;14:2919–2937. [PubMed: 11114882]
Blais JD, Addison C, Edge R, Falls T, Zhao H, Wary K, Koumenis C, Harding HP, Ron D, Holcik M,
Bell JC. Perk-dependent translational regulation promotes tumor cell adaptation and angiogenesis in
response to hypoxic stress. Molecular and Cellular Biology 2006;26:9517–9532. [PubMed:
17030613]
Blomgren K, Zhu C, Wang X, Karlsson JO, Leverin AL, Bahr BA, Mallard C, Hagberg H. Synergistic
activation of caspase-3 by m-calpain after neonatal hypoxia-ischemia: a mechanism of “pathological
apoptosis”? Journal of Biological Chemistry 2001;276:10191–10198. [PubMed: 11124942]
Bloom DA, Jaiswal AK. Phosphorylation of Nrf2S40 by PKC in response to antioxidants leads to the
release of Nrf2 from INrf2 but not required for Nrf2 stabilization/accumulation in the nucleus and
Markowitz et al. Page 15
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
transcriptional activation of ARE-mediated NQ01 Gene expression. Journal of Biological Chemistry.
2003In Press
Boldyrev AA. Discrimination between apoptosis and necrosis of neurons under oxidative stress.
Biochemistry (Moscow) 2000;65:834–842. [PubMed: 10951102]
Bridges CC, Hu H, Miyauchi S, Siddaramappa UN, Ganapathy ME, Ignatowicz L, Maddox DM, Smith
SB, Ganapathy V. Induction of Cystine-Glutamate Transporter xc- by Human Immunodeficiency
Virus Type 1 Transactivator Protein Tat in Retinal Pigment Epithelium. Invest Ophthalmol Vis Sci
2004;45:2906–2914. [PubMed: 15326101]
Bridges RJ, Esslinger CS. The excitatory amino acid transporters: Pharmacological insights on substrate
and inhibitor specificity of the EAAT subtypes. Pharmacology & Therapeutics 2005;107:271–285.
[PubMed: 16112332]
Brostrom CO, Brostrom MA. Regulation of translational initiation during cellular responses to stress.
Progress in Nucleic Acid Research & Molecular Biology 1998;58:79–125. [PubMed: 9308364]
Brouwers P, Heyes MP, Moss HA, Wolters PL, Poplack DG, Markey SP, Pizzo PA. Quinolinic acid in
the cerebrospinal fluid of children with symptomatic human immunodeficiency virus type 1 disease:
relationships to clinical status and therapeutic response. Journal of Infectious Diseases
1993;168:1380–1386. [PubMed: 8245522]
Buki A, Farkas O, Doczi T, Povlishock JT. Preinjury administration of the calpain inhibitor MDL-28170
attenuates traumatically induced axonal injury. Journal of Neurotrauma 2003;20:261–268. [PubMed:
12820680]
Burdo J, Dargusch R, Schubert D. Distribution of the Cystine/Glutamate Antiporter System xc- in the
Brain, Kidney, and Duodenum. J Histochem Cytochem 2006;54:549–557. [PubMed: 16399997]
Burton NC, Kensler TW, Guilarte TR. In vivo modulation of the Parkinsonian phenotype by Nrf2.
NeuroToxicology 2006;27:1094–1100. [PubMed: 16959318]
Canolle B, Masmejean F, Melon C, Nieuollon A, Pisano P, Lortet S. Glial soluble factors regulate the
activity and expression of the neuronal glutamate transporter EAAC1: implication of cholesterol.
Journal of Neurochemistry 2004;88:1521–1532. [PubMed: 15009653]
Carragher NO, Westhoff MA, Riley D, Potter DA, Dutt P, Elce JS, Greer PA, Frame MC. v-Src-induced
modulation of the calpain-calpastatin proteolytic system regulates transformation. Molecular &
Cellular Biology 2002;22:257–269. [PubMed: 11739739]
Carvalho DP, Dupuy C, Gorin Y, Legue O, Pommier J, Haye B, Virion A. The Ca2+- and reduced
nicotinamide adenine dinucleotide phosphate-dependent hydrogen peroxide generating system is
induced by thyrotropin in porcine thyroid cells. Endocrinology 1996;137:1007–1012. [PubMed:
8603567]
Castagna A, Grazie CL, Accordini A, Giulidori P, Cavalli G, Bottiglieri T, Lazzarin A. Cerebrospinal
fluid S-adenosylmethionine (SAMe) and glutathione concentrations in HIV infection: Effect of
parenteral treatment with SAMe. Neurology 1995;45:1678–1683. [PubMed: 7675226]
Charton JP, Herkert M, Becker CM, Schroder H. Cellular and subcellular localization of the 2B-subunit
of the NMDA receptor in the adult rat telencephalon. Brain Res 1999;816:609–617. [PubMed:
9878886]
Chatterjee S, Noack H, Possel H, Keilhoff G, Wolf G. Glutathione levels in primary glial cultures:
monochlorobimane provides evidence of cell type-specific distribution. GLIA 1999;27:152–161.
[PubMed: 10417814]
Chen M, He H, Zhan S, Krajewski S, Reed JC, Gottlieb RA. Bid is cleaved by calpain to an active fragment
in vitro and during myocardial ischemia/reperfusion. Journal of Biological Chemistry
2001;276:30724–30728. [PubMed: 11404357]
Chen W, Sulcove J, Frank I, Jaffer S, Ozdener H, Kolson DL. Development of a human neuronal cell
model for human immunodeficiency virus (HIV)-infected macrophage-induced neurotoxicity:
apoptosis induced by HIV type 1 primary Isolates and evidence for involvement of the Bcl2/Bcl-xL-
sensitive intrinsic apoptosis pathway. Journal of Virology 2002;76:9407–9419. [PubMed: 12186923]
Cheung HH, Teves L, Wallace MC, Gurd JW. Inhibition of protein kinase C reduces ischemia-induced
tyrosine phosphorylation of the N-methyl-d-aspartate receptor. J Neurochem 2003;86:1441–1449.
[PubMed: 12950452]
Markowitz et al. Page 16
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Choi J, Liu R-M, Kundu RK, Sangiorgi F, Wu W, Maxson R, Forman HJ. Molecular Mechanism of
Decreased Glutathione Content in Human Immunodeficiency Virus Type 1 Tat-transgenic Mice.
Journal of Biological Chemistry 2000;275:3693–3698. [PubMed: 10652368]
Chua BT, Guo K, Li P. Direct cleavage by the calcium-activated protease calpain can lead to inactivation
of caspases. Journal of Biological Chemistry 2000;275:5131–5135. [PubMed: 10671558]
Chung HJ, Huang YH, Lau LF, Huganir RL. Regulation of the NMDA receptor complex and trafficking
by activity-dependent phosphorylation of the NR2B subunit PDZ ligand. J Neurosci 2004;24:10248–
10259. [PubMed: 15537897]
Chung WJ, Lyons SA, Nelson GM, Hamza H, Gladson CL, Gillespie GY, Sontheimer H. Inhibition of
Cystine Uptake Disrupts the Growth of Primary Brain Tumors. Journal of Neuroscience
2005;25:7101–7110. [PubMed: 16079392]
Clements CM, McNally RS, Conti BJ, Mak TW, Ting JPY. DJ-1, a cancer- and Parkinson’s disease-
associated protein, stabilizes the antioxidant transcriptional master regulator Nrf2. PNAS
2006;103:15091–15096. [PubMed: 17015834]
Colton CA, Gilbert DL. Production of superoxide anions by a CNS macrophage, the microglia. FEBS
Letters 1987;223:284–288. [PubMed: 2822487]
Colton CA, Chernyshev ON, Gilbert DL, Vitek MP. Microglial Contribution to Oxidative Stress in
Alzheimer’s Disease. Ann NY Acad Sci 2000;899:292–307. [PubMed: 10863548]
Conant K, Garzino-Demo A, Nath A, McArthur JC, Halliday W, Power C, Gallo RC, Major EO. Induction
of monocyte chemoattractant protein-1 in HIV-1 Tat-stimulated astrocytes and elevation in AIDS
dementia. PNAS 1998;95:3117–3121. [PubMed: 9501225]
Conti F, Barbaresi P, Melone M, Ducati A. Neuronal and glial localization of NR1 and NR2A/B subunits
of the NMDA receptor in the human cerebral cortex. CerebCortex 1999;9:110–120.
Cotgreave IA, Gerdes RG. Recent trends in glutathione biochemistry--glutathione-protein interactions:
a molecular link between oxidative stress and cell proliferation? Biochemical & Biophysical
Research Communications 1998;242:1–9. [PubMed: 9439600]
Cox JS, Shamu CE, Walter P. Transcriptional induction of genes encoding endoplasmic reticulum
resident proteins requires a transmembrane protein kinase. Cell 1993;73:1197–1206. [PubMed:
8513503]
Croall DE, DeMartino GN. Calcium-activated neutral protease (calpain) system: structure, function, and
regulation. Physiological Reviews 1991;71:813–847. [PubMed: 2057527]
Crowe SM. Role of macrophages in the pathogenesis of human immunodeficiency virus (HIV) infection.
Australian & New Zealand Journal of Medicine 1995;25:777–783. [PubMed: 8770353]
Cullinan SB, Diehl JA. PERK-dependent activation of Nrf2 contributes to redox homeostasis and cell
survival following endoplasmic reticulum stress. Journal of Biological Chemistry 2004;279:20108–
20117. [PubMed: 14978030]
Cullinan SB, Diehl JA. Coordination of ER and oxidative stress signaling: The PERK/Nrf2 signaling
pathway. The International Journal of Biochemistry & Cell Biology 2006;38:317.
Cullinan SB, Zhang D, Hannink M, Arvisais E, Kaufman RJ, Diehl JA. Nrf2 Is a Direct PERK Substrate
and Effector of PERK-Dependent Cell Survival. Mol Cell Biol 2003;23:7198–7209. [PubMed:
14517290]
D’Mello SR. Molecular regulation of neuronal apoptosis. Current Topics in Developmental Biology
1998;39:187–213. [PubMed: 9476001]
De Rosa SC, Zaretsky MD, Dubs JG, Roederer M, Anderson M, Green A, Mitra D, Watanabe N,
Nakamura H, Tjioe I, Deresinski SC, Moore WA, Ela SW, Parks D, Herzenberg LA, Herzenberg
LA. N-acetylcysteine replenishes glutathione in HIV infection. [see comment]. European Journal of
Clinical Investigation 2000;30:915–929. [PubMed: 11029607]
Dinkova-Kostova AT, Holtzclaw WD, Cole RN, Itoh K, Wakabayashi N, Katoh Y, Yamamoto M, Talalay
P. Direct evidence that sulfhydryl groups of Keap1 are the sensors regulating induction of phase 2
enzymes that protect against carcinogens and oxidants. Proceedings of the National Academy of
Sciences of the United States of America 2002;99:11908–11913. [PubMed: 12193649]
Dong YN, Waxman EA, Lynch DR. Interactions of postsynaptic density-95 and the NMDA receptor 2
subunit control calpain-mediated cleavage of the NMDA receptor. J Neurosci 2004;24:11035–11045.
[PubMed: 15590920]
Markowitz et al. Page 17
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Dringen R. Metabolism and functions of glutathione in brain. Progress in Neurobiology 2000;62:649–
671. [PubMed: 10880854]
Dun Y, Mysona B, Van Ells T, Amarnath L, Shamsul Ola M, Ganapathy V, Smith S. Expression of the
cystine-glutamate exchanger (xc-) in retinal ganglion cells and regulation by nitric oxide and
oxidative stress. Cell and Tissue Research 2006;324:189. [PubMed: 16609915]
Ehret A, Westendorp MO, Herr I, Debatin KM, Heeney JL, Frank R, Krammer PH. Resistance of
chimpanzee T cells to human immunodeficiency virus type 1 Tat-enhanced oxidative stress and
apoptosis. Journal of Virology 1996;70:6502–6507. [PubMed: 8709290]
Eliasof S, McIlvain HB, Petroski RE, Foster AC, Dunlop J. Pharmacological characterization of threo-3-
methylglutamic acid with excitatory amino acid transporters in native and recombinant systems.
Journal of Neurochemistry 2001;77:550–557. [PubMed: 11299317]
Everall IP, Heaton RK, Marcotte TD, Ellis RJ, McCutchan JA, Atkinson JH, Grant I, Mallory M, Masliah
E. Cortical synaptic density is reduced in mild to moderate human immunodeficiency virus
neurocognitive disorder. HNRC Group. HIV Neurobehavioral Research Center. Brain Pathol
1999;9:209–217. [PubMed: 10219738]
Fawcett TW, Martindale JL, Guyton KZ, Hai T, Holbrook NJ. Complexes containing activating
transcription factor (ATF)/cAMP-responsive-element-binding protein (CREB) interact with the
CCAAT/enhancer-binding protein (C/EBP)-ATF composite site to regulate Gadd153 expression
during the stress response. Biochemical Journal 1999;339:135–141. [PubMed: 10085237]
Fels DR, Koumenis C. The PERK/eIF2alpha/ATF4 module of the UPR in hypoxia resistance and tumor
growth. Cancer Biology and therapy 2006;5:723–728. [PubMed: 16861899]
Fernandes SP, Edwards TM, Ng KT, Robinson SR. HIV-1 protein gp120 rapidly impairs memory in
chicks by interrupting the glutamate-glutamine cycle. Neurobiology of Learning and Memory.
2006In Press, Corrected Proof
Fine SM, Angel RA, Perry SW, Epstein LG, Rothstein JD, Dewhurst S, Gelbard HA. Tumor Necrosis
Factor alpha Inhibits Glutamate Uptake by Primary Human Astrocytes. Journal of Biological
Chemistry 1996;271:15303–15306. [PubMed: 8663435]
Floyd RA, Hensley K, Jaffery F, Maidt L, Robinson K, Pye Q, Stewart C. Increased oxidative stress
brought on by pro-inflammatory cytokines in neurodegenerative processes and the protective role of
nitrone-based free radical traps. Life Sciences 1999;65:1893–1899. [PubMed: 10576433]
Fogal B, Trettel J, Uliasz TF, Levine ES, Hewett SJ. Changes in secondary glutamate release underlie
the developmental regulation of excitotoxic neuronal cell death. Neuroscience 2005;132:929–942.
[PubMed: 15857699]
Franco SJ, Huttenlocher A. Regulating cell migration: calpains make the cut. Journal of Cell Science
2005;118:3829–3838. [PubMed: 16129881]
Gao G, Dou QP. N-terminal cleavage of bax by calpain generates a potent proapoptotic 18-kDa fragment
that promotes bcl-2-independent cytochrome C release and apoptotic cell death. Journal of Cellular
Biochemistry 2000;80:53–72. [PubMed: 11029754]
Garden GA, Budd SL, Tsai E, Hanson L, Kaul M, D’Emilia DM, Friedlander RM, Yuan J, Masliah E,
Lipton SA. Caspase Cascades in Human Immunodeficiency Virus-Associated Neurodegeneration.
Journal of Neuroscience 2002;22:4015–4024. [PubMed: 12019321]
Gasol E, Jimenez-Vidal M, Chillaron J, Zorzano A, Palacin M. Membrane Topology of System Xc-Light
Subunit Reveals a Re-entrant Loop with Substrate-restricted Accessibility. Journal of Biological
Chemistry 2004;279:31228–31236. [PubMed: 15151999]
Gelbard HA, Nottet HSLM, Swindells S, Jett M, Dzenko KA, Genis P, White R, Wang L, Choi Y-B,
Zhang D, Lipton SA, Tourtellotte WW, Epstein LG, Gendelman HE. Platelet activating factor: a
candidate HIV-1-induced neurotoxin. J Virol 1994;68:4628–4635. [PubMed: 8207837]
Ghorpade A, Holter S, Borgmann K, Persidsky R, Wu L. HIV-1 and IL-1 beta regulate Fas ligand
expression in human astrocytes through the NF-kappa B pathway. Journal of Neuroimmunology
2003;141:141–149. [PubMed: 12965265]
Gil-Parrado S, Fernandez-Montalvan A, Assfalg-Machleidt I, Popp O, Bestvater F, Holloschi A, Knoch
TA, Auerswald EA, Welsh K, Reed JC, Fritz H, Fuentes-Prior P, Spiess E, Salvesen GS, Machleidt
W. Ionomycin-activated calpain triggers apoptosis. A probable role for Bcl-2 family members.
Journal of Biological Chemistry 2002;277:27217–27226. [PubMed: 12000759]
Markowitz et al. Page 18
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Giulian D, Yu J, Li X, Tom D, Li J, Wendt E, Lin S-N, Schwarcz R, Noonan C. Study of Receptor-
Mediated Neurotoxins Released by HIV-1-Infected Mononuclear Phagocytes Found in Human
Brain. Journal of Neuroscience 1996;16:3139–3153. [PubMed: 8627353]
Glading A, Lauffenburger DA, Wells A. Cutting to the chase: calpain proteases in cell motility. Trends
in Cell Biology 2002;12:46–54. [PubMed: 11854009]
Goebel SM, Alvestad RM, Coultrap SJ, Browning MD. Tyrosine phosphorylation of the N-methyl-D-
aspartate receptor is enhanced in synaptic membrane fractions of the adult rat hippocampus. Brain
Res Mol Brain Res 2005;142:65–79. [PubMed: 16257472]
Goll DE, Thompson VF, Li H, Wei W, Cong J. The calpain system. Physiological Reviews 2003;83:731–
801. [PubMed: 12843408]
Gonzalez-Scarano F, Martin-Garcia J. The Neuropathogenesis of AIDS. Nature Reviews Immunology
2005;5:69–81.
Grynspan F, Griffin WR, Cataldo A, Katayama S, Nixon RA. Active site-directed antibodies identify
calpain II as an early-appearing and pervasive component of neurofibrillary pathology in Alzheimer’s
disease. Brain Research 1997;763:145–158. [PubMed: 9296555]
Guerini D, Pan B, Carafoli E. Expression, purification, and characterization of isoform 1 of the plasma
membrane Ca2+ pump: focus on calpain sensitivity. Journal of Biological Chemistry
2003;278:38141–38148. [PubMed: 12851406]
Guillet BA, Velly LJ, Canolle B, Masmejean FM, Nieoullon AL, Pisano P. Differential regulation by
protein kinases of activity and cell surface expression of glutamate transporters in neuron-enriched
cultures. Neurochemistry International 2005;46:337–346. [PubMed: 15707698]
Guroff G. A neutral, calcium-activated proteinase from the soluble fraction of rat brain. Journal of
Biological Chemistry 1964;239:149–155. [PubMed: 14114836]
Guttmann RP, Baker DL, Seifert KM, Cohen AS, Coulter DA, Lynch DR. Specific proteolysis of the
NR2 subunit at multiple sites by calpain. Journal of Neurochemistry 2001;78:1083–1093. [PubMed:
11553682]
Hamakubo T, Kannagi R, Murachi T, Matus A. Distribution of calpains I and II in rat brain. Journal of
Neuroscience 1986;6:3103–3111. [PubMed: 3021924]
Harding HP, Zhang Y, Ron D. Protein translation and folding are coupled by an endoplasmic-reticulum-
resident kinase.[see comment][erratum appears in Nature 1999 Mar 4;398(6722):90]. Nature
1999;397:271–274. [PubMed: 9930704]
Harding HP, Zhang Y, Bertolotti A, Zeng H, Ron D. Perk is essential for translational regulation and cell
survival during the unfolded protein response. Molecular Cell 2000;5:897–904. [PubMed: 10882126]
Hardingham GE, Bading H. The Yin and Yang of NMDA receptor signalling. Trends Neurosci
2003;26:81–89. [PubMed: 12536131]
Hart AM, Terenghi G, Kellerth JO, Wiberg M. Sensory neuroprotection, mitochondrial preservation, and
therapeutic potential of N-acetyl-cysteine after nerve injury. Neuroscience 2004;125:91–101.
[PubMed: 15051148]
Haze K, Yoshida H, Yanagi H, Yura T, Mori K. Mammalian transcription factor ATF6 is synthesized as
a transmembrane protein and activated by proteolysis in response to endoplasmic reticulum stress.
Molecular Biology of the Cell 1999;10:3787–3799. [PubMed: 10564271]
Herrup K, Yang Y. Cell cycle regulation in the postmitotic neuron: oxymoron or new biology? Nature
Reviews Neuroscience 2007;8:368–378.
Herzenberg LA, De Rosa SC, Dubs JG, Roederer M, Anderson MT, Ela SW, Deresinski SC, Herzenberg
LA. Glutathione deficiency is associated with impaired survival in HIV†disease. PNAS
1997;94:1967–1972. [PubMed: 9050888]
Hewitt KE, Lesiuk HJ, Tauskela JS, Morley P, Durkin JP. Selective coupling of mu-calpain activation
with the NMDA receptor is independent of translocation and autolysis in primary cortical neurons.
Journal of Neuroscience Research 1998;54:223–232. [PubMed: 9788281]
Heyes MP, Saito K, Lackner A, Wiley CA, Achim CL, Markey SP. Sources of the neurotoxin quinolinic
acid in the brain of HIV-1 infected patients and retrovirus-infected macaques. FASEB J
1998;12:881–896. [PubMed: 9657528]
Markowitz et al. Page 19
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Heyes MP, Ellis RJ, Ryan L, Childers ME, Grant I, Wolfson T, Archibald S, Jernigan TL, Center HGHNR.
Elevated cerebrospinal fluid quinolinic acid levels are associated with region-specific cerebral
volume loss in HIV infection. Brain 2001;124:1033–1042. [PubMed: 11335705]
Holmin S, Mathiesen T. Intracerebral administration of interleukin-1beta and induction of inflammation,
apoptosis, and vasogenic edema. Journal of Neurosurgery 2000;92:108–120. [PubMed: 10616089]
Howard MJ, David G, Barrett JN. Resealing of transected myelinated mammalian axons in vivo: evidence
for involvement of calpain. Neuroscience 1999;93:807–815. [PubMed: 10465464]
Howland DS, Liu J, She Y, Goad B, Maragakis NJ, Kim B, Erickson J, Kulik J, DeVito L, Psaltis G,
DeGennaro LJ, Cleveland DW, Rothstein JD. Focal loss of the glutamate transporter EAAT2 in a
transgenic rat model of SOD1 mutant-mediated amyotrophic lateral sclerosis (ALS). PNAS
2002;99:1604–1609. [PubMed: 11818550]
Hu W, Walters W, Xia X, Karmally S, Bethea J. Neuronal glutamate transporter EAAT4 is expressed in
astrocytes. GLIA 2003;44:13–25. [PubMed: 12951653]
Huang HC, Nguyen T, Pickett CB. Regulation of the antioxidant response element by protein kinase C-
mediated phosphorylation of NF-E2-related factor 2. [erratum appears in Proc Natl Acad Sci U S
A 2001 Jan 2;98(1):379]. Proceedings of the National Academy of Sciences of the United States
of America 2000;97:12475–12480. [PubMed: 11035812]
Huang HC, Nguyen T, Pickett CB. Phosphorylation of Nrf2 at Ser-40 by protein kinase C regulates
antioxidant response element-mediated transcription. Journal of Biological Chemistry
2002;277:42769–42774. [PubMed: 12198130]
Hurtley SM, Helenius A. Protein oligomerization in the endoplasmic reticulum. Annual Review of Cell
Biology 1989;5:277–307.
Hwang SY, Paik S, Park SH, Kim HS, Lee IS, Kim SP, Baek WK, Suh MH, Kwon TK, Park JW, Park
JB, Lee JJ, Suh SI. N-phenethyl-2-phenylacetamide isolated from Xenorhabdus nematophilus
induces apoptosis through caspase activation and calpain-mediated Bax cleavage in U937 cells.
International Journal of Oncology 2003;22:151–157. [PubMed: 12469198]
Isler JA, Skalet AH, Alwine JC. Human Cytomegalovirus infection activates and regulates the unfolded
protein response. Journal of Virology 2005;79:6890–6899. [PubMed: 15890928]
Itoh K, Wakabayashi N, Katoh Y, Ishii T, Igarashi K, Engel JD, Yamamoto M. Keap1 represses nuclear
activation of antioxidant responsive elements by Nrf2 through binding to the amino-terminal Neh2
domain. Genes & Development 1999;13:76–86. [PubMed: 9887101]
Itoh K, Chiba T, Takahashi S, Ishii T, Igarashi K, Katoh Y, Oyake T, Hayashi N, Satoh K, Hatayama I,
Yamamoto M, Nabeshima Y. An Nrf2/small Maf heterodimer mediates the induction of phase II
detoxifying enzyme genes through antioxidant response elements. Biochemical & Biophysical
Research Communications 1997;236:313–322. [PubMed: 9240432]
Jacob CP, Koutsilieri E, Bartl J, Neuen-Jacob E, Arzberger T, Zander N, Ravid R, Roggendorf W,
Riederer P, Grnblatt E. Alterations in Expression of Glutamatergic Transporters and Receptors
in Sporadic Alzheimer’s Disease. Journal of Alzheimer’s Disease 2007;11:97–116.
Johnson DA, Andrews GK, Xu W, Johnson JA. Activation of the antioxidant response element in primary
cortical neuronal cultures derived from transgenic reporter mice. Journal of Neurochemistry
2002;81:1233–1241. [PubMed: 12068071]
Kaleeba JAR, Berger EA. Kaposi’s Sarcoma-Associated Herpesvirus Fusion-Entry Receptor: Cystine
Transporter xCT. Science 2006;311:1921–1924. [PubMed: 16574866]
Katayama T, Imaizumi K, Manabe T, Hitomi J, Kudo T, Tohyama M. Induction of neuronal death by
ER stress in Alzheimer’s disease. Journal of Chemical Neuroanatomy 2004;28:67–78. [PubMed:
15363492]
Kaufman RJ. Stress signaling from the lumen of the endoplasmic reticulum: coordination of gene
transcriptional and translational controls. Genes & Development 1999;13:1211–1233. [PubMed:
10346810]
Kaul M, Garden GA, Lipton SA. Pathways to neuronal injury and apoptosis in HIV-associated dementia.
Nature 2001;410:988. [PubMed: 11309629]
Kaul M, Zheng J, Okamoto S, Lipton SA. HIV-1 infection and AIDS: consequences for the central
nervous system. Cell Death and Differentiation 2005;12:878–892. [PubMed: 15832177]
Markowitz et al. Page 20
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Kawahara T, Yanagi H, Yura T, Mori K. Endoplasmic Reticulum Stress-induced mRNA Splicing Permits
Synthesis of Transcription Factor Hac1p/Ern4p That Activates the Unfolded Protein Response. Mol
Biol Cell 1997;8:1845–1862. [PubMed: 9348528]
Kennedy MB, Manzerra P. Telling tails. Proc Natl Acad Sci U S A 2001;98:12323–12324. [PubMed:
11675481]
Kerschensteiner M, Schwab ME, Lichtman JW, Misgeld T. In vivo imaging of axonal degeneration and
regeneration in the injured spinal cord. Nature Medicine 2005;11:572–577.
Kikuchi H, Almer G, Yamashita S, Guegan C, Nagai M, Xu Z, Sosunov AA, McKhann GM 2nd,
Przedborski S. Spinal cord endoplasmic reticulum stress associated with a microsomal
accumulation of mutant superoxide dismutase-1 in an ALS model. Proceedings of the National
Academy of Sciences of the United States of America 2006;103:6025–6030. [PubMed: 16595634]
Kim MJ, Oh SJ, Park SH, Kang HJ, Won MH, Kang TC, Park JB, Kim JI, Kim J, Lee JY. Neuronal loss
in primary long-term cortical culture involves neurodegeneration-like cell death via calpain and p35
processing, but not developmental apoptosis or aging. Experimental & Molecular Medicine
2007;39:14–26. [PubMed: 17334225]
Kirby J, Halligan E, Baptista MJ, Allen S, Heath PR, Holden H, Barber SC, Loynes CA, Wood-Allum
CA, Lunec J, Shaw PJ. Mutant SOD1 alters the motor neuronal transcriptome: implications for
familial ALS. Brain 2005;128:1686–1706. [PubMed: 15872021]
Kishimoto A, Mikawa K, Hashimoto K, Yasuda I, Tanaka S, Tominaga M, Kuroda T, Nishizuka Y.
Limited proteolysis of protein kinase C subspecies by calcium-dependent neutral protease (calpain).
Journal of Biological Chemistry 1989;264:4088–4092. [PubMed: 2537303]
Kitao Y, Ozawa K, Miyazaki M, Tamatani M, Kobayashi T, Yanagi H, Okabe M, Ikawa M, Yamashima
T, Stern DM, Hori O, Ogawa S. Expression of the endoplasmic reticulum molecular chaperone
(ORP150) rescues hippocampal neurons from glutamate toxicity. Journal of Clinical Investigation
2001;108:1439–1450. [PubMed: 11714735]
Knollema S, Hom HW, Schirmer H, Korf J, Ter Horst GJ. Immunolocalization of glutathione reductase
in the murine brain. Journal of Comparative Neurology 1996;373:157–172. [PubMed: 8889919]
Koch HP, Lane Brown R, Larsson HP. The Glutamate-Activated Anion Conductance in Excitatory
Amino Acid Transporters Is Gated Independently by the Individual Subunits. Journal of
Neuroscience 2007;27:2943–2947. [PubMed: 17360917]
Koutsilieri E, Gotz ME, Sopper S, Sauer U, Demuth M, ter Meulen V, Riederer P. Regulation of
glutathione and cell toxicity following exposure to neurotropic substances and human
immunodeficiency virus-1 in vitro. Journal of Neurovirology 1997;3:342–349. [PubMed: 9372455]
Kraft AD, Johnson DA, Johnson JA. Nuclear Factor E2-Related Factor 2-Dependent Antioxidant
Response Element Activation by tert-Butylhydroquinone and Sulforaphane Occurring
Preferentially in Astrocytes Conditions Neurons against Oxidative Insult. J Neurosci
2004;24:1101–1112. [PubMed: 14762128]
Krizman-Genda E, Gonzalez M, Zelenaia O, Robinson MB. Evidence that Akt mediates platelet-derived
growth factor-dependent increases in activity and surface expression of the neuronal glutamate
transporter, EAAC1. Neuropharmacology 2005;49:872–882. [PubMed: 16182322]
Kruman II, Nath A, Mattson MP. HIV-1 protein Tat induces apoptosis of hippocampal neurons by a
mechanism involving caspase activation, calcium overload, and oxidative stress. Experimental
Neurology 1998;154:276–288. [PubMed: 9878167]
Kubbutat MH, Vousden KH. Proteolytic cleavage of human p53 by calpain: a potential regulator of
protein stability. Molecular & Cellular Biology 1997;17:460–468. [PubMed: 8972227]
Kudo T, Katayama T, Imaizumi K, Yasuda Y, Yatera M, Okochi M, Tohyama M, Takeda M. The
Unfolded Protein Response Is Involved in the Pathology of Alzheimer’s Disease. Ann NY Acad
Sci 2002;977:349–355. [PubMed: 12480772]
Kwak MK, Itoh K, Yamamoto M, Sutter TR, Kensler TW. Role of transcription factor Nrf2 in the
induction of hepatic phase 2 and antioxidative enzymes in vivo by the cancer chemoprotective agent,
3H-1, 2-dimethiole-3-thione. Molecular Medicine 2001;7:135–145. [PubMed: 11471548]
La Bella V, Valentino F, Piccoli T, Piccoli F. Expression and developmental regulation of the cystine/
glutamate exchanger (xc-) in the rat. Neurochemical Research 2007;32:1081–1090. [PubMed:
17401668]
Markowitz et al. Page 21
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Lankiewicz S, Marc Luetjens C, Truc Bui N, Krohn AJ, Poppe M, Cole GM, Saido TC, Prehn JH.
Activation of calpain I converts excitotoxic neuron death into a caspase-independent cell death.
Journal of Biological Chemistry 2000;275:17064–17071. [PubMed: 10828077]
Laurer HL, McIntosh TK. Pharmacologic therapy in traumatic brain injury: update on experimental
treatment strategies. Current Pharmaceutical Design 2001;7:1505–1516. [PubMed: 11562295]
Ledoux S, Yang R, Friedlander G, Laouari D. Glucose depletion enhances P-glycoprotein expression in
hepatoma cells: role of endoplasmic reticulum stress response. Cancer Research 2003;63:7284–
7290. [PubMed: 14612525]
Lee JM, Calkins MJ, Chan K, Kan YW, Johnson JA. Identification of the NF-E2-related factor-2-
dependent genes conferring protection against oxidative stress in primary cortical astrocytes using
oligonucleotide microarray analysis. Journal of Biological Chemistry 2003;278:12029–12038.
[PubMed: 12556532]
Lee K, Tirasophon W, Shen X, Michalak M, Prywes R, Okada T, Yoshida H, Mori K, Kaufman RJ.
IRE1-mediated unconventional mRNA splicing and S2P-mediated ATF6 cleavage merge to
regulate XBP1 in signaling the unfolded protein response. Genes & Development 2002;16:452–
466. [PubMed: 11850408]
Lee ST, Chu K, Park JE, Kang L, Ko SY, Jung KH, Kim M. Memantine reduces striatal cell death with
decreasing calpain level in 3-nitropropionic model of Huntington’s disease. Brain Research
2006;1118:199–207. [PubMed: 16959224]
Levy LM, Warr O, Attwell D. Stoichiometry of the Glial Glutamate Transporter GLT-1 Expressed
Inducibly in a Chinese Hamster Ovary Cell Line Selected for Low Endogenous Na+-Dependent
Glutamate Uptake. Journal of Neuroscience 1998;18:9620–9628. [PubMed: 9822723]
Lewerenz J, Klein M, Methner A. Cooperative action of glutamate transporters and cystine/glutamate
antiporter system Xc protects from oxidative glutamate toxicity. Journal of Neurochemistry
2006;1:1–10.
Li B, Chen N, Luo T, Otsu Y, Murphy TH, Raymond LA. Differential regulation of synaptic and extra-
synaptic NMDA receptors. Nat Neurosci 2002;5:833–834. [PubMed: 12195433]
Li BS, Sun MK, Zhang L, Takahashi S, Ma W, Vinade L, Kulkarni AB, Brady RO, Pant HC. Regulation
of NMDA receptors by cyclin-dependent kinase-5. Proc Natl Acad Sci U S A 2001;98:12742–
12747. [PubMed: 11675505]
Li M, Baumeister P, Roy B, Phan T, Foti D, Luo S, Lee AS. ATF6 as a transcription activator of the
endoplasmic reticulum stress element: thapsigargin stress-induced changes and synergistic
interactions with NF-Y and YY1. Molecular & Cellular Biology 2000;20:5096–5106. [PubMed:
10866666]
Li S, Mallory ME, Alford M, Tanaka S, Masliah E. Glutamate Transporter Alterations in Alzheimer
Disease are Possibly Associated with Abnormal APP Expression. Journal of Neuropathology and
Experimental Neurology 1997;56:910–911.
Lieberthal W, Menza SA, Levine JS. Graded ATP depletion can cause necrosis or apoptosis of cultured
mouse proximal tubular cells. American Journal of Physiology 1998;274:F315–327. [PubMed:
9486226]
Lin C-LG, Bristol LA, Jin L, Dykes-Hoberg M, Crawford T, Clawson L, Rothstein JD. Aberrant RNA
Processing in a Neurodegenerative Disease: the Cause for Absent EAAT2, a Glutamate Transporter,
in Amyotrophic Lateral Sclerosis. Neuron 1998;20:589–602. [PubMed: 9539131]
Lipton SA. Paradigm shift in NMDA receptor antagonist drug development: molecular mechanism of
uncompetitive inhibition by memantine in the treatment of Alzheimer’s disease and other neurologic
disorders. Journal of Alzheimers Disease 2004;6:S61–S74.
Lipton SA, Yeh M, Dreyer EB. Update on current models of HIV-related neuronal injury: platelet-
activating factor, arachidonic acid and nitric oxide. Advances in Neuroimmunology 1994;4:181–
188. [PubMed: 7874385]
Lovell MA, Xie C, Markesbery WR. Decreased glutathione transferase activity in brain and ventricular
fluid in Alzheimer’s disease. Neurology 1998;51:1562–1566. [PubMed: 9855502]
Lu T, Xu Y, Mericle MT, Mellgren RL. Participation of the conventional calpains in apoptosis.
Biochimica et Biophysica Acta 2002;1590:16–26. [PubMed: 12063165]
Markowitz et al. Page 22
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Lynch DR, Guttmann RP. NMDA receptor pharmacology: perspectives from molecular biology. Curr
Drug Targets 2001;2:215–231. [PubMed: 11554549]
Lynch DR, Guttmann RP. Excitotoxicity: perspectives based on N-Methyl-D-Aspartate receptor
subtypes. Journal of Pharmacology and Experimental Therapeutics 2002;300:717–723. [PubMed:
11861773]
Malorni W, Rivabene R, Lucia BM, Ferrara R, Mazzone AM, Cauda R, Paganelli R. The role of oxidative
imbalance in progression to AIDS: effect of the thiol supplier N-acetylcysteine. AIDS Research &
Human Retroviruses 1998;14:1589–1596. [PubMed: 9840292]
Mandic A, Viktorsson K, Strandberg L, Heiden T, Hansson J, Linder S, Shoshan MC. Calpain-mediated
Bid cleavage and calpain-independent Bak modulation: two separate pathways in cisplatin-induced
apoptosis. Molecular & Cellular Biology 2002;22:3003–3013. [PubMed: 11940658]
Mao W, Iwai C, Qin F, Liang C-s. Norepinephrine induces endoplasmic reticulum stress and
downregulation of norepinephrine transporter density in PC12 cells via oxidative stress. Am J
Physiol Heart Circ Physiol 2005;288:H2381–2389. [PubMed: 15626688]
Maragakis NJ, Dietrich J, Wong V, Xue H, Mayer-Proschel M, Rao MS, Rothstein JD. Glutamate
transporter expression and function in human glial progenitors. GLIA 2003;45:133–143. [PubMed:
14730707]
Masliah E, Ge N, Achim CL, Hansen LA, Wiley CA. Selective neuronal vulnerability in HIV encephalitis.
Journal of Neuropathology and Experimental Neurology 1992;51:585–593. [PubMed: 1484289]
Masliah E, Alford M, Mallory ME, Rockenstein E, Moechars D, Van Leuven F. Abnormal glutamate
transport function in mutant amyloid precursor protein transgenic mice. Experimental Neurology
2000;163:381–387. [PubMed: 10833311]
Mathern GW, Mendoza D, Lozada A, Pretorius JK, Dehnes Y, Danbolt NC, Nelson N, Leite JP, Chimelli
L, Born DE, Sakamoto AC, Assirati JA, Fried I, Peacock WJ, Ojemann GA, Adelson PD.
Hippocampal GABA and glutamate transporter immunoreactivity in patients with temporal lobe
epilepsy. Neurology 1999;52:453–473. [PubMed: 10025773]
Matthews CC, Zielke HR, Wollack JB, Fishman PS. Enzymatic Degradation Protects Neurons from
Glutamate Excitotoxicity. Journal of Neurochemistry 2000;75:1045–1052. [PubMed: 10936185]
McBean GJ. Cerebral cystine uptake: a tale of two transporters. Trends in Pharmacological Sciences
2002;23:299–302. [PubMed: 12119142]
McCracken E, Hunter AJ, Patel S, Graham DI, Dewar D. Calpain activation and cytoskeletal protein
breakdown in the corpus callosum of head-injured patients. Journal of Neurotrauma 1999;16:749–
761. [PubMed: 10521135]
McMahon M, Itoh K, Yamamoto M, Chanas SA, Henderson CJ, McLellan LI, Wolf CR, Cavin C, Hayes
JD. The Cap‘n’Collar basic leucine zipper transcription factor Nrf2 (NF-E2 p45-related factor 2)
controls both constitutive and inducible expression of intestinal detoxification and glutathione
biosynthetic enzymes. Cancer Research 2001;61:3299–3307. [PubMed: 11309284]
Meaney J, Balcar V, Rothstein JD, Jeffrey P. Glutamate transport in cultures from developing avian
cerebellum: presence of GLT-1 immunoreactivity in Purkinje neurons. Journal of Neuroscience
Research 1998;54:595–603. [PubMed: 9843150]
Molinari M, Carafoli E. Calpain: a cytosolic proteinase active at the membranes. Journal of Membrane
Biology 1997;156:1–8. [PubMed: 9070458]
Mollace V, Nottet HS, Clayette P, Turco MC, Muscoli C, Salvemini D, Perno CF. Oxidative stress and
neuroAIDS: triggers, modulators and novel antioxidants. Trends in Neurosciences 2001;24:411–
416. [PubMed: 11410272]
Moore JD, Rothwell NJ, Gibson RM. Involvement of caspases and calpains in cerebrocortical neuronal
cell death is stimulus-dependent. British Journal of Pharmacology 2002;135:1069–1077. [PubMed:
11861336]
Moussa CE-H, Rae C, Bubb WA, Griffin JL, Deters NA, Balcar VJ. Inhibitors of glutamate transport
modulate distinct patterns in brain metabolism. Journal of Neuroscience Research 2007;85:342–
350. [PubMed: 17086545]
Muller F, Svardal AM, Nordoy I, Berge RK, Aukrust P, Froland SS. Virological and immunological
effects of antioxidant treatment in patients with HIV infection. European Journal of Clinical
Investigation 2000;30:905–914. [PubMed: 11029606]
Markowitz et al. Page 23
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Musa FR, Tokuda M, Kuwata Y, Ogawa T, Tomizawa K, Konishi R, Takenaka I, Hatase O. Expression
of cyclin-dependent kinase 5 and associated cyclins in Leydig and Sertoli cells of the testis. Journal
of Andrology 1998;19:657–666. [PubMed: 9876017]
Nakagawa T, Yuan J. Cross-talk between two cysteine protease families. Activation of caspase-12 by
calpain in apoptosis. Journal of Cell Biology 2000;150:887–894. [PubMed: 10953012]
Nath A, Conant K, Chen P, Scott C, Major EO. Transient Exposure to HIV-1 Tat Protein Results in
Cytokine Production in Macrophages and Astrocytes. A hit and run phenomenon. J Biol Chem
1999;274:17098–17102. [PubMed: 10358063]
Nicotera P, Ankarcrona M, Bonfoco E, Orrenius S, Lipton SA. Neuronal necrosis and apoptosis: two
distinct events induced by exposure to glutamate or oxidative stress. Advances in Neurology
1997;72:95–101. [PubMed: 8993688]
Nikawa J, Yamashita S. IRE1 encodes a putative protein kinase containing a membrane-spanning domain
and is required for inositol phototrophy in Saccharomyces cerevisiae. Molecular Microbiology
1992;6:1441–1446. [PubMed: 1625574]
Nishida K, Markey SP, Kustova Y, Morse HC, Skolnick P, Basile AS, Sei Y. Increased Brain Levels of
Platelet-Activating Factor in a Murine Acquired Immune Deficiency Syndrome Are NMDA
Receptor-Mediated. Journal of Neurochemistry 1996;66:433–435. [PubMed: 8522986]
Norenberg MD, Martinez-Hernandez A. Fine structural localization of glutamine synthetase in astrocytes
of rat brain. Brain Research 1979;161:303–310. [PubMed: 31966]
Novoa I, Zeng H, Harding HP, Ron D. Feedback inhibition of the unfolded protein response by GADD34-
mediated dephosphorylation of eIF2alpha. Journal of Cell Biology 2001;153:1011–1022. [PubMed:
11381086]
Numazawa S, Ishikawa M, Yoshida A, Tanaka S, Yoshida T. Atypical protein kinase C mediates
activation of NF-E2-related factor 2 in response to oxidative stress. American Journal of Physiology
-Cell Physiology 2003;285:C334–342. [PubMed: 12700136]
O’Connor E, Devesa A, Garcia C, Puertes IR, Pellin A, Vina JR. Biosynthesis and maintenance of GSH
in primary astrocyte cultures: role of L-cystine and ascorbate. Brain Research 1995;680:157–163.
[PubMed: 7663972]
O’Donnell LA, Agrawal A, Jordan-Sciutto KL, Dichter MA, Lynch DR, Kolson DL. Human
immunodeficiency virus (HIV)-induced neurotoxicity: Roles for the NMDA receptor subtypes. J
Neurosci 2006;26:981–990. [PubMed: 16421318]
O’Hanlon GM, Humphreys PD, Goldman RS, Halstead SK, Bullens RW, Plomp JJ, Ushkaryov Y,
Willison HJ. Calpain inhibitors protect against axonal degeneration in a model of anti-ganglioside
antibody-mediated motor nerve terminal injury. Brain 2003;126:2497–2509. [PubMed: 12937083]
O’Hare MJ, Kushwaha N, Zhang Y, Aleyasin H, Callaghan SM, Slack RS, Albert PR, Vincent I, Park
DS. Differential roles of nuclear and cytoplasmic cyclin-dependent kinase 5 in apoptotic and
excitotoxic neuronal death. Journal of Neuroscience 2005;25:8954–8966. [PubMed: 16192386]
Okamura K, Kimata Y, Higashio H, Tsuru A, Kohno K. Dissociation of Kar2p/BiP from an ER Sensory
Molecule, Ire1p, Triggers the Unfolded Protein Response in Yeast. Biochemical and Biophysical
Research Communications 2000;279:445–450. [PubMed: 11118306]
Pacchioni AM, Vallone J, Melendez RI, Shih AY, Murphy TH, Kaliva PW. Nrf2 gene deletion fails to
alter psychostimulant-induced behavior or neurotoxicity. Brain Research 2007;1127:26–35.
[PubMed: 17113054]
Palacin M, Estevez R, Bertan J, Zorzano A. Molecular biology of mammalian plasma membrane amino
acid transporters. Physiological Reviews 1998;78:969–1054. [PubMed: 9790568]
Pariat M, Carillo S, Molinari M, Salvat C, Debussche L, Bracco L, Milner J, Piechaczyk M. Proteolysis
by calpains: a possible contribution to degradation of p53. Molecular & Cellular Biology
1997;17:2806–2815. [PubMed: 9111352]
Pariat M, Salvat C, Bebien M, Brockly F, Altieri E, Carillo S, Jariel-Encontre I, Piechaczyk M. The
sensitivity of c-Jun and c-Fos proteins to calpains depends on conformational determinants of the
monomers and not on formation of dimers. Biochemical Journal 345 Pt 2000;1:129–138.
Patrick GN, Zukerberg L, Nikolic M, de la Monte S, Dikkes P, Tsai LH. Conversion of p35 to p25
deregulates Cdk5 activity and promotes neurodegeneration.[see comment]. Nature 1999;402:615–
622. [PubMed: 10604467]
Markowitz et al. Page 24
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Peghini P, Janzen J, Stoffel W. Glutamate transporter EAAC-1-deficient mice develop dicarboxylic
aminoaciduria and behavioral abnormalities but no neurodegeneration. The EMBO Journal
1997;16:3822–3832. [PubMed: 9233792]
Perrin BJ, Huttenlocher A. Calpain. International Journal of Biochemistry & Cell Biology 2002;34:722–
725. [PubMed: 11950589]
Peuchen S, Bolanos JP, Heales SJR, Almeida A, Duchen MR, Clark JB. Interrelationships between
astrocyte function, oxidative stress and antioxidant status within the central nervous system.
Progress in Neurobiology 1997;32:261–281. [PubMed: 9247965]
Pfaff M, Du X, Ginsberg MH. Calpain cleavage of integrin beta cytoplasmic domains. FEBS Letters
1999;460:17–22. [PubMed: 10571053]
Philbert MA, Beiswanger CM, Waters DK, Reuhl KR, Lowndes HE. Cellular and regional distribution
of reduced glutathione in the nervous system of the rat: Histochemical localization by mercury
orange and o-phthaldialdehyde-induced histofluorescence. Toxicology and Applied Pharmacology
1991;107:215–227. [PubMed: 1994508]
Pines G, Danbolt NC, Bjoras M, Zhang Y, Bendahan A, Eide L, Koepsell H, Storm-Mathisen J, Seeberg
E, Kanner BI. Cloning and expression of a rat brain L-glutamate transporter. Nature 1992;360:464–
467. [PubMed: 1448170]
Pitt D, Nagelmeier IE, Wilson HC, Raine CS. Glutamate uptake by oligodendrocytes: Implications for
excitotoxicity in multiple sclerosis. Neurology 2003;61:1113–1120. [PubMed: 14581674]
Polster BM, Basanez G, Etxebarria A, Hardwick JM, Nicholls DG. Calpain I induces cleavage and release
of apoptosis-inducing factor from isolated mitochondria. Journal of Biological Chemistry
2005;280:6447–6454. [PubMed: 15590628]
Pradillo JM, Hurtado O, Romera C, Cardenas A, Fernandez-Tome P, Alonso-Escolano D, Lorenzo P,
Moro MA, Lizasoain I. TNFR1 mediates increased neuronal membrane EAAT3 expression after
in vivo cerebral ischemic preconditioning. Neuroscience 2006;138:1171–1178. [PubMed:
16442237]
Price TO, Ercal N, Nakaoke R, Banks WA. HIV-1 viral proteins gp120 and Tat induce oxidative stress
in brain endothelial cells. Brain Research 2005;1045:57. [PubMed: 15910762]
Proper EA, Hoogland G, Kappen SM, Jansen GH, Rensen MGA, Schrama LH, van Veelen CWM, van
Rijen PC, van Nieuwenhuizen O, Gispen WH, de Graan PNE. Distribution of glutamate transporters
in the hippocampus of patients with pharmaco-resistant temporal lobe epilepsy. Brain 2002;125:32–
43. [PubMed: 11834591]
Pulliam L, Herndier BG, Tang NM, McGrath MS. Human immunodeficiency virus-infected macrophages
produce soluble factors that cause histological and neurochemical alteration in cultures human
brains. J Clin Invest 1991;87:503–512. [PubMed: 1671392]
Pulliam L, Clarke JA, McGuire D, McGrath MS. Investigation of HIV-infected macrophage neurotoxin
production from patients with AIDS dementia. Advances in Neuroimmunology 1994;4:195–198.
[PubMed: 7874387]
Pulliam L, Zhou M, Stubblebine M, Bitler CM. Differential modulation of cell death proteins in human
brain cells by tumor necrosis factor alpha and platelet activating factor. Journal of Neuroscience
Research 1998;54:530–538. [PubMed: 9822163]
Pulliam L, Clarke JA, McGrath MS, Moore D, McGuire D. Monokine products as predictors of AIDS
dementia. AIDS 1996;10:1495–1500. [PubMed: 8931783]
Qiang W, Cahill JM, Liu J, Kuang X, Liu N, Scofield VL, Voorhees JR, Reid AJ, Yan M, Lynn WS,
Wong PK. Activation of transcription factor Nrf-2 and its downstream targets in response to
moloney murine leukemia virus ts1-induced thiol depletion and oxidative stress in astrocytes. J
Virol 2004;78:11926–11938. [PubMed: 15479833]
Qin S, Colin C, Hinners I, Gervais A, Cheret C, Mallat M. System Xc- and Apolipoprotein E Expressed
by Microglia Have Opposite Effects on the Neurotoxicity of Amyloid-beta Peptide 1–40. J Neurosci
2006;26:3345–3356. [PubMed: 16554485]
Ramassamy C, Averill D, Beffert U, Bastianetto S, Theroux L, Lussier-Cacan S, Cohn JS, Christen Y,
Davignon J, Quirion R, Poirier J. Oxidative damage and protection by antioxidants in the frontal
cortex of Alzheimer’s disease is related to the apolipoprotein E genotype. Free Radical Biology and
Medicine 1999;27:544–553. [PubMed: 10490274]
Markowitz et al. Page 25
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Ramsey CP, Glass CA, Montgomery MB, Lindl KA, Ritson GP, Chia LA, Hamilton RL, Chu CT, Jordan-
Sciutto KL. Expression of Nrf2 in Neurodegenerative Diseases. Journal of Neuropathology and
Experimental Neurology 2007;66:75–85. [PubMed: 17204939]
Ray SK, Banik NL. Calpain and its involvement in the pathophysiology of CNS injuries and diseases:
therapeutic potential of calpain inhibitors for prevention of neurodegeneration. Current Drug
Targets -CNS & Neurological Disorders 2003;2:173–189. [PubMed: 12769798]
Ray SK, Shields DC, Saido TC, Matzelle DC, Wilford GG, Hogan EL, Banik NL. Calpain activity and
translational expression increased in spinal cord injury. Brain Research 1999;816:375–380.
[PubMed: 9878837]
Raynaud F, Marcilhac A. Implication of calpain in neuronal apoptosis. A possible regulation of
Alzheimer’s disease. FEBS Journal 2006;273:3437–3443. [PubMed: 16884489]
Rechsteiner M, Rogers SW. PEST sequences and regulation by proteolysis. Trends in Biochemical
Sciences 1996;21:267–271. [PubMed: 8755249]
Riccio A, Ginty DD. What a privilege to reside at the synapse: NMDA receptor signaling to CREB. Nat
Neurosci 2002;5:389–390. [PubMed: 11976696]
Rimaniol A-C, Haik S, Martin M, Le Grand R, Boussin FD, Dereuddre-Bosquet N, Gras G, Dormont D.
Na+-Dependent High Affinity Glutamate Transport in Macrophages. Journal of Immunology
2000;164:5430–5438.
Roche KW, Standley S, McCallum J, Dune Ly C, Ehlers MD, Wenthold RJ. Molecular determinants of
NMDA receptor internalization. Nat Neurosci 2001;4:794–802. [PubMed: 11477425]
Romero-Alvira D, Roche E. The keys of oxidative stress in acquired immune deficiency syndrome
apoptosis. Medical Hypotheses 1998;51:169–173. [PubMed: 9881826]
Rothstein JD, Dykes-Hoberg M, Pardo CA, Bristol LA, Jin L, Kuncl RW, Kanai Y, Hediger MA, Wang
Y, Schielke JP, Welty DF. Knockout of Glutamate Transporters Reveals a Major Role for Astroglial
Transport in Excitotoxicity and Clearance of Glutamate. Neuron 1996;16:675–686. [PubMed:
8785064]
Rumbaugh G, Vicini S. Distinct synaptic and extrasynaptic NMDA receptors in developing cerebellar
granule neurons. J Neurosci 1999;19:10603–10610. [PubMed: 10594044]
Ryu EJ, Harding HP, Angelastro JM, Vitolo OV, Ron D, Greene LA. Endoplasmic reticulum stress and
the unfolded protein response in cellular models of Parkinson’s disease. Journal of Neuroscience
2002;22:10690–10698. [PubMed: 12486162]
Sa MJ, Madeira MD, Ruela C, Volk B, Mota-Miranda A, Paula-Barbosa MM. Dendritic changes in the
hippocampal formation of AIDS patients: a quantitative Golgi study. Acta Neuropathol(Berl)
2004;107:97–110. [PubMed: 14605830]
Saido TC, Shibata M, Takenawa T, Murofushi H, Suzuki K. Positive regulation of mu-calpain action by
polyphosphoinositides. Journal of Biological Chemistry 1992;267:24585–24590. [PubMed:
1332961]
Saido TC, Nagao S, Shiramine M, Tsukaguchi M, Yoshizawa T, Sorimachi H, Ito H, Tsuchiya T,
Kawashima S, Suzuki K. Distinct kinetics of subunit autolysis in mammalian m-calpain activation.
FEBS Letters 1994;346:263–267. [PubMed: 8013644]
Sanders V, Pittman C, White M, Wang G, Wiley C, Achim C. Chemkines and receptors in HIV
encephalitis. AIDS 1998;12:1021–1026. [PubMed: 9662198]
Sasaki H, Sato H, Kuriyama-Matsumura K, Sato K, Maebara K, Wang H, Tamba M, Itoh K, Yamamoto
M, Bannai S. Electrophile response element-mediated induction of the cystine/glutamate exchange
transporter gene expression. Journal of Biological Chemistry 2002;277:44765–44771. [PubMed:
12235164]
Sato H, Tamba M, Ishii T, Bannai S. Cloning and Expression of a Plasma Membrane Cystine/Glutamate
Exchange Transporter Composed of Two Distinct Proteins. J Biol Chem 1999;274:11455–11458.
[PubMed: 10206947]
Sato H, Nomura S, Maebara K, Sato K, Tamba M, Bannai S. Transcriptional control of cystine/glutamate
transporter gene by amino acid deprivation. Biochemical and Biophysical Research
Communications 2004;325:109. [PubMed: 15522208]
Markowitz et al. Page 26
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Sato H, Tamba M, Okuno S, Sato K, Keino-Masu K, Masu M, Bannai S. Distribution of Cystine/
Glutamate Exchange Transporter, System Xc-, in the Mouse Brain. J Neurosci 2002;22:8028–8033.
[PubMed: 12223556]
Sattler R, Tymianski M. Molecular mechanisms of glutamate receptor-mediated excitotoxic neuronal
cell death. Mol Neurobiol 2001;24:107–129. [PubMed: 11831548]
Sayre LM, Smith MA, Perry G. Chemistry and biochemistry of oxidative stress in neurodegenerative
disease. Current Medicinal Chemistry 2001;8:721–738. [PubMed: 11375746]
Scott DB, Blanpied TA, Swanson GT, Zhang C, Ehlers MD. An NMDA receptor ER retention signal
regulated by phosphorylation and alternative splicing. J Neurosci 2001;21:3063–3072. [PubMed:
11312291]
Shen X, Zhang K, Kaufman RJ. The Unfolded Protein Response - a Stress Signaling Pathway of the
Endoplasmic Reticulum. Journal of Chemical Neuroanatomy 2004;28:79–92. [PubMed: 15363493]
Shi Y, Vattem KM, Sood R, An J, Liang J, Stramm L, Wek RC. Identification and characterization of
pancreatic eukaryotic initiation factor 2 alpha-subunit kinase, PEK, involved in translational control.
Molecular & Cellular Biology 1998;18:7499–7509. [PubMed: 9819435]
Shields DC, Schaecher KE, Saido TC, Banik NL. A putative mechanism of demyelination in multiple
sclerosis by a proteolytic enzyme, calpain. Proceedings of the National Academy of Sciences of the
United States of America 1999;96:11486–11491. [PubMed: 10500203]
Shih AY, Li P, Murphy TH. A small-molecule-inducible Nrf2-mediated antioxidant response provides
effective prophylaxis against cerebral ischemia in vivo. Journal of Neuroscience 2005;25:10321–
10335. [PubMed: 16267240]
Shih AY, Erb H, Sun X, Toda S, Kalivas PW, Murphy TH. Cystine/Glutamate Exchange Modulates
Glutathione Supply for Neuroprotection from Oxidative Stress and Cell Proliferation. J Neurosci
2006;26:10514–10523. [PubMed: 17035536]
Shih AY, Johnson DA, Wong G, Kraft AD, Jiang L, Erb H, Johnson JA, Murphy TH. Coordinate
Regulation of Glutathione Biosynthesis and Release by Nrf2-Expressing Glia Potently Protects
Neurons from Oxidative Stress. J Neurosci 2003;23:3394–3406. [PubMed: 12716947]
Shimamoto K, Sakai R, Takaoka K, Yumoto N, Nakajima T, Amara SG, Shigeri Y. Characterization of
Novel L-threo-beta-Benzyloxyaspartate Derivatives, Potent Blockers of the Glutamate
Transporters. Molecular Pharmacology 2004;65:1008–1015. [PubMed: 15044631]
Shioda N, Moriguchi S, Shirasaki Y, Fukunaga K. Generation of constitutively active calcineurin by
calpain contributes to delayed neuronal death following mouse brain ischemia. Journal of
Neurochemistry 2006;98:310–320. [PubMed: 16805817]
Shor-Posner G, Lecusay R, Morales G, Campa A, Miguez-Burbano M-J. Neuroprotection in HIV-positive
drug users: Implications for antioxidant therapy. JAIDS 2002;31:S84–S88. [PubMed: 12394787]
Siao C-J, Tsirka SE. Tissue plasminogen activator mediates microglial activation via its finger domain
through annexin II. JNeurosci 2002;22:3352–3358. [PubMed: 11978811]
Sidrauski C, Walter P. The transmembrane kinase Ire1p is a site-specific endonuclease that initiates
mRNA splicing in the unfolded protein response. Cell 1997;90:1031–1039. [PubMed: 9323131]
Simpkins KL, Guttmann RP, Dong Y, Chen Z, Sokol S, Neumar RW, Lynch DR. Selective activation
induced cleavage of the NR2B subunit by calpain. Journal of Neuroscience 2003;23:11322–11331.
[PubMed: 14672996]
Snyder EM, Nong Y, Almeida CG, Paul S, Moran T, Choi EY, Nairn AC, Salter MW, Lombroso PJ,
Gouras GK, Greengard P. Regulation of NMDA receptor trafficking by amyloid-beta. Nat Neurosci
2005;8:1051–1058. [PubMed: 16025111]
Sofic E, Lange KW, Jellinger K, Riederer P. Reduced and oxidized glutathione in the substantia nigra of
patients with Parkinson’s disease. Neuroscience Letters 1992;142:128–130. [PubMed: 1454205]
Sohal RS, Weindruch R. Oxidative Stress, Caloric Restriction, and Aging. Science 1996;273:59–63.
[PubMed: 8658196]
Sprengel R, Suchanek B, Amico C, Brusa R, Burnashev N, Rozov A, Hvalby O, Jensen V, Paulsen O,
Andersen P, Kim JJ, Thompson RF, Sun W, Webster LC, Grant SG, Eilers J, Konnerth A, Li J,
McNamara JO, Seeburg PH. Importance of the intracellular domain of NR2 subunits for NMDA
receptor function in vivo. Cell 1998;92:279–289. [PubMed: 9458051]
Markowitz et al. Page 27
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Spreux-Varoquaux O, Bensimon G, Lacomblez L, Salachas F, Pradat PF, Le Forestier N, Marouan A,
Dib M, Meininger V. Glutamate levels in cerebrospinal fluid in amyotrophic lateral sclerosis: a
reappraisal using a new HPLC method with coulometric detection in a large cohort of patients.
Journal of the Neurological Sciences 2002;193:73–78. [PubMed: 11790386]
Squier MK, Sehnert AJ, Sellins KS, Malkinson AM, Takano E, Cohen JJ. Calpain and calpastatin regulate
neutrophil apoptosis. Journal of Cellular Physiology 1999;178:311–319. [PubMed: 9989777]
Strachan GD, Koike MA, Siman R, Hall DJ, Jordan-Sciutto KL. E2F1 induces cell death, calpain
activation, and MDMX degradation in a transcription independent manner implicating a novel role
for E2F1 in neuronal loss in SIV encephalitis. Journal of Cellular Biochemistry 2005;96:728–740.
[PubMed: 16088944]
Su, Z-z; Leszczyniecka, M.; Kang, D-c; Sarkar, D.; Chao, W.; Volsky, DJ.; Fisher, PB. Insights into
glutamate transport regulation in human astrocytes: Cloning of the promoter for excitatory amino
acid transporter 2 (EAAT2). PNAS 2003;100:1955–1960. [PubMed: 12578975]
Suen PC, Wu K, Levine ES, Mount HT, Xu JL, Lin SY, Black IB. Brain-derived neurotrophic factor
rapidly enhances phosphorylation of the postsynaptic N-methyl-D-aspartate receptor subunit 1.
Proc Natl Acad Sci U S A 1997;94:8191–8195. [PubMed: 9223337]
Sun X, Erb H, Murphy TH. Coordinate regulation of glutathione metabolism in astrocytes by Nrf2.
Biochemical and Biophysical Research Communications 2005;326:371–377. [PubMed: 15582588]
Suzuki, K. The structure of the calpains and the calpain gene. In: Mellgren, RL.; Murachi, T., editors.
Intracellular Calcium-Dependent Proteolysis. Boca Raton, FL: CRC; 1990. p. 25-35.
Suzuki K. Nomenclature of calcium dependent proteinase. Biomedica Biochimica Acta 1991;50:483–
484. [PubMed: 1801713]
Suzuki K, Hata S, Kawabata Y, Sorimachi H. Structure, activation, and biology of calpain. Diabetes
2004;53(Suppl 1):S12–18. [PubMed: 14749260]
Swanson RA, Liu J, Miller JW, Rothstein JD, Farrell KA, Stein B, Longuemare MC. Neuronal Regulation
of Glutamate Transporter Subtype Expression in Astrocytes. J Neurosci 1997;17:932–940.
[PubMed: 8994048]
Takano J, Tomioka M, Tsubuki S, Higuchi M, Iwata N, Itohara S, Maki M, Saido TC. Calpain mediates
excitotoxic DNA fragmentation via mitochondrial pathways in adult brains: evidence from
calpastatin mutant mice. Journal of Biological Chemistry 2005;280:16175–16184. [PubMed:
15691848]
Takikita S, Takano T, Narita T, Takikita M, Ohno M, Shimada M. Neuronal apoptosis mediated by IL-1
beta expression in viral encephalitis caused by a neuroadapted strain of the mumps virus (Kilham
Strain) in hamsters. Experimental Neurology 2001;172:47–59. [PubMed: 11681839]
Tanaka K, Watase K, Manabe T, Yamada K, Watanabe M, Takahashi K, Iwama H, Nishikawa T, Ichihara
N, Kikuchi T, Okuyama S, Kawashima N, Hori S, Takimoto M, Wada K. Epilepsy and Exacerbation
of Brain Injury in Mice Lacking the Glutamate Transporter GLT-1. Science 1997;276:1699–1702.
[PubMed: 9180080]
Tian G, Lai L, Guo H, Lin Y, Butchbach MER, Chang Y, Lin C-lG. Translational Control of Glial
Glutamate Transporter EAAT2 Expression. J Biol Chem 2007;282:1727–1737. [PubMed:
17138558]
Tomimatsu Y, Idemoto S, Moriguchi S, Watanabe S, Nakanishi H. Proteases involved in long-term
potentiation. Life Sciences 2002;72:355–361. [PubMed: 12467876]
Tompa P, Buzder-Lantos P, Tantos A, Farkas A, Szilagyi A, Banoczi Z, Hudecz F, Friedrich P. On the
sequential determinants of calpain cleavage. Journal of Biological Chemistry 2004;279:20775–
20785. [PubMed: 14988399]
Tovar KR, Westbrook GL. The incorporation of NMDA receptors with a distinct subunit composition at
nascent hippocampal synapses in vitro. J Neurosci 1999;19:4180–4188. [PubMed: 10234045]
Tsuji T, Shimohama S, Kimura J, Shimizu K. m-Calpain (calcium-activated neutral proteinase) in
Alzheimer’s disease brains. Neuroscience Letters 1998;248:109–112. [PubMed: 9654354]
Tullio RD, Passalacqua M, Averna M, Salamino F, Melloni E, Pontremoli S. Changes in intracellular
localization of calpastatin during calpain activation. Biochemical Journal 343 Pt 1999;2:467–472.
Markowitz et al. Page 28
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Turchan J, Pocernich CB, Gairola C, Chauhan A, Schifitto G, Butterfield DA, Buch S, Narayan O, Sinai
A, Geiger J, Berger J, Elford H, Nath A. Oxidative stress in HIV demented patients and protection
ex vivo with novel antioxidants. Neurology 2003;60:307–314. [PubMed: 12552050]
Valeyev NV, Downing AK, Skorinkin AI, Campbell ID, Kotov NV. A calcium dependent de-adhesion
mechanism regulates the direction and rate of cell migration: a mathematical model. In Silico
Biology 2006;6:545–572. [PubMed: 17518764]
van Muiswinkel FL, Kuiperij HB. The Nrf2-ARE Signalling Pathway: Promising Drug Target to Combat
Oxidative Stress in Neurodegenerative Disorders. Current Drug Targets - CNS & Neurological
Disorders 2005;4:267–281. [PubMed: 15975029]
Vandenberg RJ, Mitrovic AD, Chebib M, Balcar VJ, Johnston GAR. Contrasting Modes of Action of
Methylglutamate Derivatives on the Excitatory Amino Acid Transporters, EAAT1 and EAAT2.
Molecular Pharmacology 1997;51:809–815. [PubMed: 9145919]
Vanderklish PW, Bahr BA. The pathogenic activation of calpain: a marker and mediator of cellular
toxicity and disease states. International Journal of Experimental Pathology 2000;81:323–339.
[PubMed: 11168679]
Velaz-Faircloth M, McGraw TS, alandro MS, Fremeau RT Jr, Kilberg MS, Anderson KJ.
Characterization and distribution of the neuronal glutamate transporter EAAC1 in rat brain. Am J
Physiol Cell Physiol 1996;270:C67–75.
Vissel B, Krupp JJ, Heinemann SF, Westbrook GL. A use-dependent tyrosine dephosphorylation of
NMDA receptors is independent of ion flux. Nat Neurosci 2001;4:587–596. [PubMed: 11369939]
Viviani B, Corsini E, Binaglia M, Galli CL, Marinovich M. Reactive oxygen species generated by glia
are responsible for neuron death induced by human immunodeficiency virus-glycoprotein 120 in
vitro. Neuroscience 2001;107:51–58. [PubMed: 11744246]
Wang KK, Yuen PW. Development and therapeutic potential of calpain inhibitors. Advances in
Pharmacology 1997;37:117–152. [PubMed: 8891101]
Wang KK, Larner SF, Robinson G, Hayes RL. Neuroprotection targets after traumatic brain injury.
Current Opinion in Neurology 2006;19:514–519. [PubMed: 17102687]
Wang S, Lees GJ, Rosengren LE, Karlsson JE, Hamberger A, Haglid KG. Proteolysis of filament proteins
in glial and neuronal cells after in vivo stimulation of hippocampal NMDA receptors.
Neurochemical Research 1992;17:1005–1009. [PubMed: 1387196]
Wang XJ, Xu JX. Salvianic acid A protects human neuroblastoma SH-SY5Y cells against MPP+-induced
cytotoxicity. Neuroscience Research 2005;51:129–138. [PubMed: 15681030]
Wang Y, White MG, Akay C, Chodroff R, Robinson J, Lindl KA, Dichter MA, Qian Y, Mao Z, Kolson
D, Jordan-Sciutto K. Activation of cyclin-dependent kinase 5 by calpain contributes to human
immunodeficiency virus-induced neurotoxicity. Journal of Neurochemistry. 2007In Press
Wang Z, Pekarskaya O, Bencheikh M, Chao W, Gelbard HA, Ghorpade A, Rothstein JD, Volsky DJ.
Reduced expression of glutamate transporter EAAT2 and impaired glutamate transport in human
primary astrocytes exposed to HIV-1 or gp120. Virology 2003;312:60–73. [PubMed: 12890621]
Waxman EA, Lynch DR. N-methyl-D-aspartate receptor subtypes: multiple roles in excitotoxicity and
neurological disease. Neuroscientist 2005;11:37–49. [PubMed: 15632277]
Wei Y, Wang D, Topczewski F, Pagliassotti MJ. Saturated fatty acids induce endoplasmic reticulum
stress and apoptosis independently of ceramide in liver cells. American Journal of Physiology -
Endocrinology & Metabolism 2006;291:E275–281. [PubMed: 16492686]
Weinreb O, Mandel S, Amit T, Youdim MBH. Neurological mechanisms of green tea polyphenols in
Alzheimer’s and Parkinson’s diseases. The Journal of Nutritional Biochemistry 2004;15:506.
[PubMed: 15350981]
White MG, Luca LE, Nonner D, Saleh O, Hu B, Barrett EF, Barrett JN. Cellular mechanisms of neuronal
damage from hyperthermia. Progress In Brain Research 2007:162.
Wisman LAB, van Muiswinkel FL, de Graan PNE, Hol EM, Bar PR. Cells over-expressing EAAT2
protect motoneurons from excitotoxic death in vitro. NeuroReport 2003;14:1967–1970. [PubMed:
14561930]
Wood DE, Thomas A, Devi LA, Berman Y, Beavis RC, Reed JC, Newcomb EW. Bax cleavage is
mediated by calpain during drug-induced apoptosis. Oncogene 1998;17:1069–1078. [PubMed:
9764817]
Markowitz et al. Page 29
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Wu H, Jin Y, Buddhala C, Osterhaus G, Cohen E, Jin H, Wei J, Davis K, Obata K, Wu J-Y. Role of
glutamate decarboxylase (GAD) isoform, GAD65, in GABA synthesis and transport into synaptic
vesicles--Evidence from GAD65-knockout mice studies. Brain Research 2007;1154:80. [PubMed:
17482148]
Wu HY, Lynch DR. Calpain and synaptic function. Molecular Neurobiology 2006;33:215–236.
[PubMed: 16954597]
Xue X, Piao J-H, Nakajima A, Sakon-Komazawa S, Kojima Y, Mori K, Yagita H, Okumura K, Harding
H, Nakano H. Tumor necrosis factor alpha (TNF-alpha) induces the unfolded protein response
(UPR) in a reactive oxygen species (ROS)-dependent fashion, and the UPR conteracts ROS
accumulation by TNF-alpha. Journal of Biological Chemistry 2005;280:33917–33925. [PubMed:
16107336]
Yamada K, Watanabe M, Shibata T, Nagashima M, Tanaka K, Inoue Y. Glutamate Transporter GLT-1
Is Transiently Localized on Growing Axons of the Mouse Spinal Cord before Establishing
Astrocytic Expression. Journal of Neuroscience 1998;18:5706–5713. [PubMed: 9671661]
Yamashima T. Implication of cysteine proteases calpain, cathepsin and caspase in ischemic neuronal
death of primates. Progress in Neurobiology 2000;62:273–295. [PubMed: 10840150]
Yao JK, Leonard S, Reddy R. Altered glutathione redox state in schizophrenia. Disease Markers
2006;22:83–93. [PubMed: 16410648]
Ye J, Rawson RB, Komuro R, Chen X, Dave UP, Prywes R, Brown MS, Goldstein JL. ER stress induces
cleavage of membrane-bound ATF6 by the same proteases that process SREBPs. Molecular Cell
2000;6:1355–1364. [PubMed: 11163209]
Ye Z-C, Rothstein JD, Sontheimer H. Compromised Glutamate Transport in Human Glioma Cells:
Reduction-Mislocalization of Sodium-Dependent Glutamate Transporters and Enhanced Activity
of Cystine-Glutamate Exchange. J Neurosci 1999;19:10767–10777. [PubMed: 10594060]
Yernool D, Boudker O, Jin Y, Gouaux E. Structure of a glutamate transporter homologue from
Pyrococcus horikoshii. Nature 2004;431:811–818. [PubMed: 15483603]
Yoshida H, Matsui T, Yamamoto A, Okada T, Mori K. XBP1 mRNA is induced by ATF6 and spliced
by IRE1 in response to ER stress to produce a highly active transcription factor. Cell 2001;107:881–
891. [PubMed: 11779464]
Yoshioka M, Bradley WG, Shapshak P, Nagano I, Stewart RV, Xin KQ, Srivastava AK, Nakamura S.
Role of immune activation and cytokine expression in HIV-1-associated neurologic diseases.
Advances in Neuroimmunology 1995;5:335–358. [PubMed: 8748077]
Yu Z, Luo H, Fu W, Mattson MP. The endoplasmic reticulum stress-responsive protein GRP78 protects
neurons against excitotoxicity and apoptosis: suppression of oxidative stress and stabilization of
calcium homeostasis. Experimental Neurology 1999;155:302–314. [PubMed: 10072306]
Zerangue N, Kavanaugh MP. Flux coupling in a neuronal glutamate transporter. Nature 1996;383:634–
637. [PubMed: 8857541]
Zhai Q, Wang J, Kim A, Liu Q, Watts R, Hoopfer E, Mitchison T, Luo L, He Z. Involvement of the
ubiquitin-proteasome system in the early stages of wallerian degeneration. Neuron 2003;39:217–
225. [PubMed: 12873380]
Zhang W, Lu Q, Xie ZJ, Mellgren RL. Inhibition of the growth of WI-38 fibroblasts by
benzyloxycarbonyl-Leu-Leu-Tyr diazomethyl ketone: evidence that cleavage of p53 by a calpain-
like protease is necessary for G1 to S-phase transition. Oncogene 1997;14:255–263. [PubMed:
9018111]
Markowitz et al. Page 30
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 1. The integrated stress response (ISR)
In response to oxidative stress, unfolded proteins, inflammation and other cellular stresses, a
stress response is activated to reduce cellular metabolism until cellular homeostasis is restored.
In response to stress, the endoplasmic reticulum chaperone protein BiP utilization by unfolded
proteins frees the 3 initiators of the ISR: ATF6, IRE1α, and PERK. ATF6 is translocated to
the golgi and cleaved releasing ATF6-p50, a transcription factor. Free IRE1α splices uXBP1
to sXBP1 which can now be translated. PERK phosphorylates 2 known substrates, eIF2α and
Nrf2. eIF2α depresses general translation initiation, but favors translation of ATF4. ATF6-p53,
sXBP1, and ATF4 translocate to the nucleus and activate transcription of chaperone proteins
like BiP to re-establish homeostasis. Concomitantly, phosphorylation of Nrf2 by PERK
disrupts interaction between Keap1 and Nrf2 leading to translocation of Nrf2 to the nucleus
and transactivation of antioxidant genes including GST and xCT. Together, these responses
reduce oxidative stress and establish protein refolding.
Markowitz et al. Page 31
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 2. Model for ISR activation in altering astrocyte neuroprotection
In neurodegeneration, ISR activation occurs due to chronic exposure to reactive oxygen
species, unfolded proteins and/or inflammatory mediators. Chronic activation of ISR leads to
constant elevation of the glutamate/cystine antiporter (xCT) via transcriptional activation by
Nrf2. Increased xCT exports glutamate out of the astrocyte in order to import cystine for
glutathione generation to reduce reactive oxygen species. Inflammatory mediators and,
potentially, ISR activation results in decreased EAAT1/2 activity blocking glutamate uptake
from the extracellular milieu, leading to an increase in extracellular glutamate. Increased
extracellular glutamate is free to stimulate NMDA receptors on neurons causing influx of
Ca2+ ions. Increased Ca2+ levels activate calpain proteases causing damage to dendritic spines
and dendrites. Such damage to neurons will lead to neuronal dysfunction. Finally, chronic or
extreme activation of calpain will lead to cell death.
Markowitz et al. Page 32
Cellscience. Author manuscript; available in PMC 2009 January 2.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
... Oxidative stress is most notably caused by the production and accumulation of reactive oxygen species (ROS) and has been recognized as a key contributor to many neurodegenerative diseases, including Alzheimer's disease (Martins et al., 2018). ROS are highly reactive, chemically unstable, and cause oxidative damage to cells resulting in lipid peroxidation, nucleic acid oxidation and mitochondrial dysfunction, but levels are kept relatively low by cellular antioxidants (Markowitz et al., 2007). Direct antioxidants can scavenge ROS and are redox active, while indirect antioxidants can induce cytoprotective enzymes and intracellular antioxidants (Markowitz et al., 2007). ...
... ROS are highly reactive, chemically unstable, and cause oxidative damage to cells resulting in lipid peroxidation, nucleic acid oxidation and mitochondrial dysfunction, but levels are kept relatively low by cellular antioxidants (Markowitz et al., 2007). Direct antioxidants can scavenge ROS and are redox active, while indirect antioxidants can induce cytoprotective enzymes and intracellular antioxidants (Markowitz et al., 2007). Conversely, when antioxidant defenses exceed oxidant production, a type of 'reductive stress' can occur; this may occur in the early stages of neurodegenerative disease before symptomology and may be a preceding stage towards subsequent oxidative stress which compounds further neurotoxicity (Lloret et al., 2016). ...
Article
Full-text available
Phytochemicals are often promoted generally as antioxidants and demonstrate variable levels of reactive oxygen species (ROS) sequestration in vitro, which attributes to their neuroprotective bioactivity. Sesquiterpenes from cannabis and essential oils may demonstrate bifunctional properties towards cellular oxidative stress, possessing pro-oxidant activities by generating ROS or scavenging ROS directly. Sesquiterpenes can also oxidize forming sesquiterpene oxides, however the relative contribution they make to the bioactivity or cytotoxicity of complex botanical extracts more generally is unclear, while selected cannabis-prevalent terpenes such as β-caryophyllene may also activate cannabinoid receptors as part of their biological activity. In the present study, we investigated selected sesquiterpenes β-caryophyllene and humulene and their oxidized forms (β-caryophyllene oxide and zerumbone, respectively) against established antioxidants (ascorbic acid, α-tocopherol, and glutathione) and in the presence of cannabinoid receptor 1 and cannabinoid receptor 2 antagonists, to gain a better understanding of the molecular and cellular mechanisms of neuroprotection versus neurotoxicity in semi-differentiated rat neuronal phaeochromocytoma (PC12) cells. Our results demonstrate that the sesquiterpenes β-caryophyllene, humulene and zerumbone possess concentration-dependent neurotoxic effects in PC12 cells. Both β-caryophyllene- and humulene-evoked toxicity was unaffected by CB1 or CB2 receptor antagonism, demonstrating this occurred independently of cannabinoid receptors. Both glutathione and α-tocopherol were variably able to alleviate the concentration-dependent loss of PC12 cell viability from exposure to β-caryophyllene, humulene and zerumbone. During 4-hour exposure to sesquiterpenes only modest increases in ROS levels were noted in PC12 cells, with glutathione co-incubation significantly inhibiting intracellular ROS production. However, significant increases in ROS levels in PC12 cells were demonstrated during 24-hour incubation with either antioxidants or sesquiterpenes individually, and with additive toxicity exhibited in combination. Overall, the results highlight a concentration-dependent profile of sesquiterpene neurotoxicity independent of cannabinoid receptors and dissociated from the formation of reactive oxygen species as a marker or correlate to the loss of cell viability.
... Second, lysed or dead cells release cytosolic glutamate (39). Third, the reverse cystine/glutamate antiporter function of neuronal-glial cells surrounds the neurons (40). Fourth, the presence of impaired regulation of glutamate receptors (40). ...
... Third, the reverse cystine/glutamate antiporter function of neuronal-glial cells surrounds the neurons (40). Fourth, the presence of impaired regulation of glutamate receptors (40). These mechanisms lead to a positive feedback loop of cell death and accumulation of glutamate neurotransmitters promoting the development of excitatory neurotoxicity. ...
Article
Full-text available
Objectives To evaluate the diagnostic and prognostic values of glutamate chemical exchange saturation transfer (GluCEST) magnetic resonance imaging as a quantitative method for pathogenetic research and clinical application of carbon monoxide (CO) poisoning-induced encephalopathy combined with the proton magnetic resonance spectroscopy (¹H-MRS) and the related histopathological and behavioral changes. Methods A total of 63 Sprague–Dawley rats were randomly divided into four groups. Group A (n = 12) was used for animal modeling verification; Group B (n = 15) was used for magnetic resonance molecular imaging, Group C (n = 15) was used for animal behavior experiments, and Group D (n = 21) was used for histopathological examination. All the above quantitative results were analyzed by statistics. Results The peak value of carboxyhemoglobin saturation in the blood after modeling was 7.3-fold higher than before and lasted at least 2.5 h. The GluCEST values of the parietal lobe, hippocampus, and thalamus were significantly higher than the base values in CO poisoning rats (p < 0.05) and the ¹H-MRS showed significant differences in the parietal lobe and hippocampus. In the Morris water maze tests, the average latency and distance were significantly prolonged in poisoned rats (p < 0.05), and the cumulative time was shorter and negatively correlated with GluCEST. Conclusion The GluCEST imaging non-invasively reflects the changes of glutamate in the brain in vivo with higher sensitivity and spatial resolution than ¹H-MRS. Our study implies that GluCEST imaging may be used as a new imaging method for providing a pathogenetic and prognostic assessment of CO-associated encephalopathy.
... Excessive extracellular glutamate concentrations reverse glutamate antiporter, so glial cells no longer have enough cystine to synthesize gluthione (GSH), an antioxidant [26]. Lack of GSH leads to more reactive oxygen species (ROSs) that damage and kill the glial cell, which is unable to uptake and process extracellular glutamate (Nicholls, 2009). ...
Article
Full-text available
Monosodium glutamate (MSG) is basically used to develop the flavour of variety food items all over the world. The interest of using MSG has been growing rapidly with respect to food processing industry since last 30 years. Many researchers have identified MSG as the most controversial food ingredients for its negative impact towards the animal and human. The present review is nothing but a detail study on MSG, its historical background, chemical characters, commercial importance, use as food additives and its direct and indirect negative impacts on health. This specific study has pointed out how some common diseases like epidemic obesity, diabetes and cardiovascular disease are increasing rapidly day by day in the Indian continental area, linked with excess intake of processed food among the people. Thus, MSG intake also creates several lifestyles related disorder indirectly. The aim of this review is to aware people about the role of MSG on human health.
... These results in combination with the data presented in this report suggest that increased ROS in combination with decreased cellular capacity to manage oxidative stress may occur simultaneously in HIV+ patient brains to impair neuronal function.In addition to disrupting glutamate balance through macrophage infection directly, both HIV and ARVs can alter astrocytic function.Excitatory amino acid transporter 2 (EAAT-2) expression is decreased in astrocytes following treatment with either lopinavir or amprenavir, resulting in increased extracellular glutamate(Vivithanaporn et al. 2016). Additionally, HIV-induced oxidative stress in astrocytes can lead to glutamate release through increased glutathione production as well as other mechanisms(Vázquez-Santiago et al. 2014;Markowitz et al. 2007). Lending further support to the proposal that HIV and ARVs act in concert to disrupt neuronal function, neurons exposed to ARVs may have heightened responses to elevated glutamate concentrations.Indeed, Robertson et al, 2012 found that certain prescribed combinations of ARVs caused neurons to have exaggerated calcium influx in response to glutamate. ...
Article
HIV-associated neurocognitive disorder (HAND) describes a wide range of cognitive impairments experienced by up to 55% of HIV+ individuals despite viral suppression by combined antiretroviral therapy. Reasons for the persistence of this disease are unknown, but may be related to both the presence of HIV-infected macrophages in the central nervous system as well as neurotoxicity of antiretroviral drugs. In this thesis, we identified two independent mechanisms of HIV-associated and antiretroviral-associated toxicity that may each contribute distinctly to HAND neuropathogenesis. First, we showed that β-site amyloid precursor protein cleaving enzyme 1 (BACE1), which may play a role in the onset and progression of Alzheimer’s Disease, was both increased in HIV+ patient brains and required for HIV-associated neurotoxicity in vitro. The BACE1 cleavage target amyloid precursor protein (APP) also mediated toxicity and was required for neuroprotective effects of BACE1 inhibition. Second, we showed that two frontline treatment antiretroviral drugs have neurotoxic potential in vitro and that neurotoxicity of antiretrovirals is highly variable both across and within drug classes. Neurotoxicity of one drug, lopinavir, was mediated by oxidative stress. Taken together, these data indicate that HIV and antiretrovirals may contribute to HAND persistence and that both BACE1 inhibitors and drugs targeting oxidative stress may be effective as adjunctive therapeutics in HIV+ patients.
... The activated microglial cells can induce oxidative stress when cultured with neuronal cells. 42,43 Although the response is natural, oxidative stress can be detrimental to ...
Article
Full-text available
Spinal cord injury (SCI) commonly leads to lifelong disability due to the limited regenerative capacity of the adult central nervous system. Nanomicelles can be used as therapeutic systems to provide effective treatments for SCI. In this study, a novel triblock monomethyl poly(ethylene glycol)-poly(l-lactide)-poly(trimethylene carbonate) copolymer was successfully synthesized. Next, polymeric nanomicelles loaded with zonisamide (ZNS), a Food and Drug Administration-approved antiepileptic drug, were prepared and characterized. The ZNS-loaded micelles (ZNS-M) were further utilized for the treatment of SCI in vitro and in vivo. The obtained ZNS-M were ~50 nm in diameter with good solubility and dispersibility. Additionally, these controlled-release micelles showed significant antioxidative and neuron-protective effects in vitro. Finally, our results indicated that ZNS-M treatment could promote motor function recovery and could increase neuron and axon density in a hemisection SCI model. In summary, these results may provide an experimental basis for the use of ZNS-M as a clinically applicable therapeutic drug for the treatment of SCI in the future.
... In our studies, the histopathological changes in brain of fetuses could also be explained through induction of oxidative stress. Lack of GSH leads to more reactive oxygen species (ROSs) that damage and kill the glial cell, which then cannot reuptake and process extracellular glutamate as stated [45]. ...
Article
Full-text available
Abstract Aims: Oxcarbazepine (OXC) is a newer antiepileptic agent that has recently become increasingly used either as monotherapy or as an adjunct to other AEDs in adults, adolescents, and children with partial epilepsy. Our aim was to define the the potential risks of the anti-epileptic drug (OXC) orally administered daily to the pregnant rats. Methodology: The pregnant rats administered from 7th till 20th day of gestation with 108 mg/kg oxcarbazepine (Human equivalent dose (HED)). All pregnant rats of the two groups were sacrificed and the growth parameters, skeletal malformation and the histopathology of liver, kidney and brain of the fetus were examined. Results: Our results showed that Oxcarbazepine induced a reduction in the fetal weight and length, delayed, weak and incomplete ossification, wavy ribs and the fetal liver revealed histopathological changes, degenerated hepatocytes possessed karyorrhexed or karyolysed nuclei, the congestion of blood vessels and sinusoids. Kidney revealed alternation changes, shrinked glomeruli, widened capsular space of the Bowman's capsule, hydrophobic degeneration of the tubules and cytoplasmic vacuole. Brain (cerebral cortex) showed neurodegenerative changes, marked neuronal cell degeneration, disorganization of the brain tissue, numerous pyknotised cells and vacuolization of the neuropil. Biochemical studies showed that OXC induced a reduction in the level GSH and catalase compared to control group. Conclusion: These support and proof the potential risks of the OXC administration on fetus. Keywords : Rats; pregnancy; oxcarbazepine; anti-epileptic drug.
Article
One of the changes found in the brain in Alzheimer’s disease (AD) is increased calpain, derived from calcium dysregulation, oxidative stress, and/or neuroinflammation, which are all assumed to be basic pillars in neurodegenerative diseases. The role of calpain in synaptic plasticity, neuronal death, and AD has been discussed in some reviews. However, astrocytic calpain changes sometimes appear to be secondary and consequent to neuronal damage in AD. Herein, we explore the possibility of calpain-mediated astroglial reactivity in AD, both preceding and during the amyloid phase. We discuss the types of brain calpains but focus the review on calpains 1 and 2 and some important targets in astrocytes. We address the signaling involved in controlling calpain expression, mainly involving p38/mitogen-activated protein kinase and calcineurin, as well as how calpain regulates the expression of proteins involved in astroglial reactivity through calcineurin and cyclin-dependent kinase 5. Throughout the text, we have tried to provide evidence of the connection between the alterations caused by calpain and the metabolic changes associated with AD. In addition, we discuss the possibility that calpain mediates amyloid-β clearance in astrocytes, as opposed to amyloid-β accumulation in neurons.
Article
Diabetic retinopathy (DR), the most frequent complication of diabetes, is one of the leading causes of irreversible blindness in working-age adults and has traditionally been regarded as a microvascular disease. However, increasing evidence has revealed that synaptic neurodegeneration of retinal ganglion cells (RGCs) and activation of glial cells may represent some of the earliest events in the pathogenesis of DR. Upon diabetes-induced metabolic stress, abnormal glycogen synthase kinase-3β (GSK-3β) activation drives tau hyperphosphorylation and β-catenin downregulation, leading to mitochondrial impairment and synaptic neurodegeneration prior to RGC apoptosis. Moreover, glial cell activation triggers enhanced inflammation and oxidative stress, which may accelerate the deterioration of diabetic RGCs neurodegeneration. These findings have opened up opportunities for therapies, such as inhibition of GSK-3β, glial cell activation, glutamate excitotoxicity and the use of neuroprotective drugs targeting early neurodegenerative processes in the retina and halting the progression of DR before the manifestation of microvascular abnormalities. Such interventions could potentially remedy early neurodegeneration and help prevent vision loss in people suffering from DR.
Article
Various studies reported the possibility of deterioration of blood–brain barrier (BBB) integrity owing to the aging process. The current work was performed to investigate the ability of Monosodium glutamate (MSG) to cross BBB in aged rats, the damage affecting the anterior horn cells of the spinal cord due to excitotoxicity, and the mechanisms by which quercetin (Que) administration might suppress such damage. Forty male rats aged 18 months were assigned equally to 4 groups: control group, Que group (received Que, 20 mg/kg/d intraperitonealy for 10 days), MSG group (received MSG, 4.0 g/kg/d subcutaneously for 10 days), MSG + Que group (received both Que and MSG as done in the Que and MSG groups respectively). Cervical spinal cord specimens were obtained and prepared for routine histological study, immunohistochemical staining by caspase-3 and glial fibrillary acidic protein (GFAP), assessment of oxidative stress, measurement of cytokines, assessment of caspase-3 activity and GFAP levels as well as for western blotting of phosphorylated activating transcription factor 2 (ATF2pp) as an indicator for the activity of p38 mitogen-activated protein kinase (MAPK). The MSG group revealed variable degenerative and apoptotic changes in the motoneurons and neuroglia, a marked rise in the cytoplasmic caspase-3 expression in motoneurons and a significant reduction (p < 0.001) in the astrocyte surface area percentage. In addition, the spinal cord tissue exhibited a significant elevation (p < 0.001) in the levels of malondialdehyde (MDA), IL-1, IL-6, TNFα, INFɣ, caspase-3 activity and ATF2 pp expression as well as a significant reduction (p < 0.001) in SOD, IL-10 and GFAP levels compared with the control group. On combining Que with MSG, most of the degenerative changes were reversed and all the impaired parameters were nearly normalized except for IL-6 and GFAP levels which were still significantly (p < 0.05) different from those of the control group. Our study suggests that MSG can break through the BBB of the aged rats and induce excitotoxicity dependent changes in spinal cord motoneurons. Most of these changes were reversed by Que probably via targeting the p38 MAPK-ATF2 pathway, antagonizing oxidative stress, anti-inflammatory effect, and promoting GFAP expression.
Article
Full-text available
Human immunodeficiency virus (HIV) has been associated with inflammatory effects that may potentially result in neurodegenerative changes and a number of newer chemotherapeutic agents are being tested to ameliorate these effects. In this study, we investigated the anti-neuroinflammatory activity of a novel resveratrol analog 4-(E)-{(p-tolylimino)-methylbenzene-1,2-diol} (TIMBD) against HIV1-gp120 induced neuroinflammation in SVG astrocytes. SVG astrocytic cells were pretreated with TIMBD or resveratrol (RES) and then transfected with a plasmid encoding HIV1-gp120. The mRNA and protein expression levels of proinflammatory cytokines IL6, IL8 and CCL5 were determined. Protein expression levels of NF-κB, AP1, p-STAT3, p-AKT, p-IKKs and p-p38 MAPK were also determined. TIMBD inhibited gp120-induced RNA and protein expression levels of IL6 and IL8, but not that of CCL5 in SVG astrocytes. Moreover, TIMBD attenuated gp120-induced phosphorylation of cJUN, cFOS, STAT3, p38-MAPK, AKT and IKKs, and the nuclear translocation of NF-κB p-65 subunit whereas RES mostly affected NF-κB protein expression levels. Our results suggest that TIMBD exerts anti-inflammatory effects better than that of RES in SVG astrocytes in vitro. These effects seem to be regulated by AP1, STAT-3 and NF-κB signaling pathways. TIMBD may thus have a potential of being a novel agent for treating HIV1-gp120-mediated neuroinflammatory diseases.
Article
Full-text available
Reuptake plays an important role in regulating synaptic and extracellular concentrations of glutamate. Three glutamate transporters expressed in human motor cortex, termed EAAT1, EAAT2, and EAAT3 (for excitatory amino acid transporter), have been characterized by their molecular cloning and functional expression. Each EAAT subtype mRNA was found in all human brain regions analyzed. The most prominent regional variation in message content was in cerebellum where EAAT1 expression predominated. EAAT1 and EAAT3 mRNAs were also expressed in various non-nervous tissues, whereas expression of EAAT2 was largely restricted to brain. The kinetic parameters and pharmacological characteristics of transport mediated by each EAAT subtype were determined in transfected mammalian cells by radio-label uptake and in microinjected oocytes by voltage-clamp measurements. The affinities of the EAAT subtypes for L-glutamate were similar, with Km determinations varying from 48 to 97 microM in the mammalian cell assay and from 18 to 28 microM in oocytes. Glutamate uptake inhibitors were used to compare the pharmacologies of the EAAT subtypes. The EAAT2 subtype was distinguishable from the EAAT1/EAAT3 subtypes by the potency of several inhibitors, but most notably by sensitivity to kainic acid (KA) and dihydrokainic acid (DHK). KA and DHK potently inhibited EAAT2 transport, but did not significantly affect transport by EAAT1/EAAT3. Using voltage-clamp measurements, most inhibitors were found to be substrates that elicited transport currents. In contrast, KA and DHK did not evoke currents and they were found to block EAAT2-mediated transport competitively. This selective interaction with the EAAT2 subtype could be a significant factor in KA neurotoxicity. These studies provide a foundation for understanding the role of glutamate transporters in human excitatory neurotransmission and in neuropathology.
Article
Full-text available
Molecular biology entered the field of mammalian amino acid transporters in 1990-1991 with the cloning of the first GABA and cationic amino acid transporters. Since then, cDNA have been isolated for more than 20 mammalian amino acid transporters. All of them belong to four protein families. Here we describe the tissue expression, transport characteristics, structure-function relationship, and the putative physiological roles of these transporters. Wherever possible, the ascription of these transporters to known amino acid transport systems is suggested. Significant contributions have been made to the molecular biology of amino acid transport in mammals in the last 3 years, such as the construction of knockouts for the CAT-1 cationic amino acid transporter and the EAAT2 and EAAT3 glutamate transporters, as well as a growing number of studies aimed to elucidate the structure-function relationship of the amino acid transporter. In addition, the first gene (rBAT) responsible for an inherited disease of amino acid transport (cystinuria) has been identified. Identifying the molecular structure of amino acid transport systems of high physiological relevance (e.g., system A, L, N, and x(c)(-)) and of the genes responsible for other aminoacidurias as well as revealing the key molecular mechanisms of the amino acid transporters are the main challenges of the future in this field.
Article
Full-text available
Gadd153, also known as chop, encodes a member of the CCAAT/enhancer-binding protein (C/EBP) transcription factor family and is transcriptionally activated by cellular stress signals. We recently demonstrated that arsenite treatment of rat pheochromocytoma PC12 cells results in the biphasic induction of Gadd153 mRNA expression, controlled in part through binding of C/EBPβ and two uncharacterized protein complexes to the C/EBP–ATF (activating transcription factor) composite site in the Gadd153 promoter. In this report, we identified components of these additional complexes as two ATF/CREB (cAMP-responsive-element-binding protein) transcription factors having differential binding activities dependent upon the time of arsenite exposure. During arsenite treatment of PC12 cells, we observed enhanced binding of ATF4 to the C/EBP–ATF site at 2 h as Gadd153 mRNA levels increased, and enhanced binding of ATF3 complexes at 6 h as Gadd153 expression declined. We further demonstrated that ATF4 activates, while ATF3 represses, Gadd153 promoter activity through the C/EBP–ATF site. ATF3 also repressed ATF4-mediated transactivation and arsenite-induced activation of the Gadd153 promoter. Our results suggest that numerous members of the ATF/CREB family are involved in the cellular stress response, and that regulation of stress-induced biphasic Gadd153 expression in PC12 cells involves the ordered, sequential binding of multiple transcription factor complexes to the C/EBP–ATF composite site.
Article
Basal extracellular glutamate sampled in vivo is present in micromolar concentrations in the extracellular space outside the synaptic cleft, and neither the origin nor the function of this glutamate is known. This report reveals that blockade of glutamate release from the cystine-glutamate antiporter produced a significant decrease (60%) in extrasynaptic glutamate levels in the rat striatum, whereas blockade of voltage-dependent Na+ and Ca2+ channels produced relatively minimal changes (0-30%). This indicates that the primary origin of in vivo extrasynaptic glutamate in the striatum arises from nonvesicular glutamate release by the cystine-glutamate antiporter. By measuring [35S]cystine uptake, it was shown that similar to vesicular release, the activity of the cystine-glutamate antiporter is negatively regulated by group II metabotropic glutamate receptors (mGluR2/3) via a cAMP-dependent protein kinase mechanism. Extracellular glutamate derived from the antiporter was shown to regulate extracellular levels of glutamate and dopamine. Infusion of the mGluR2/3 antagonist (RS)-1-amino-5-phosphonoindan-1-carboxylic acid (APICA) increased extracellular glutamate levels, and previous blockade of the antiporter prevented the APICA-induced rise in extracellular glutamate. This suggests that glutamate released from the antiporter is a source of endogenous tone on mGluR2/3. Blockade of the antiporter also produced an increase in extracellular dopamine that was reversed by infusing the mGluR2/3 agonist (2R,4R)-4-aminopyrrolidine-2,4-dicarboxlylate, indicating that antiporter-derived glutamate can modulate dopamine transmission via mGluR2/3 heteroreceptors. These results suggest that nonvesicular release from the cystine-glutamate antiporter is the primary source of in vivo extracellular glutamate and that this glutamate can modulate both glutamate and dopamine transmission.
Article
Hydrogen peroxide (H2O2) is an essential electron acceptor for thyroid peroxidase-catalyzed iodination and coupling reactions. In the presence of iodide, its production is a limiting step in thyroid hormone biosynthesis. Several studies have demonstrated that the thyroid particulate fraction contains a Ca2+- and NADPH- dependent H@O@ generator (NADPH-O2:oxidoreductase), the so- called thyroid NADPH-oxidase. It has recently been demonstrated that cellular H2O2 release is under the tonic control of TSH in primary cultures of dog thyrocytes. The present study evaluates the effect of TSH on the thyroid NADPH-oxidase and cytochrome c reductase activities, two enzymes believed to be involved on H2O2 generation in the thyroid gland. There was almost no detectable NADPH-dependent H2O2 generator in the membranes of cells grown for 18 h without TSH. But cells grown in the presence of TSH (0.1 mU/ml) had a CA2+- and NADPH-dependent H2O2-generating activity that increased up to the third day in culture, as did the ce...
Article
Cognitive impairment is a frequent manifestation of advanced human immunodeficiency virus (HIV) infection. The response to antiretroviral medication is often partial and poorly sustained. Recent studies suggest that free radical production within the CNS and neuronal apoptosis may play important roles in the pathogenesis of HN dementia. We conducted a randomized double-blind, placebo-controlled trial using a parallel group, 2 x 2 factorial design evaluating deprenyl, a monoamine oxidase B inhibitor and putative anti-apoptotic agent, and thioctic acid, an antioxidant, in 36 patients with HIV-associated cognitive impairment. Both deprenyl and thioctic acid were well tolerated with few adverse events. Deprenyl recipients showed significant improvement on tests of verbal memory compared with patients not taking deprenyl. Thioctic acid treatment did not improve cognitive function. These results suggest that deprenyl treatment is associated with cognitive improvement in subjects with mild HN-associated cognitive impairment, whereas thioctic acid has no benefit. A larger efficacy trial is needed to assess the long-term effect of deprenyl on cognitive performance in patients with HIV infection.
Article
6-Hydroxydopamine, 1-methyl-4-phenyl-pyridinium (MPP+), and rotenone cause the death of dopaminergic neurons in vitro and in vivo and are widely used to model Parkinson's disease. To identify regulated genes in such models, we performed serial analysis of gene expression on neuronal PC12 cells exposed to 6-hydroxydopamine. This revealed a striking increase in transcripts associated with the unfolded protein response. Immunoblotting confirmed phosphorylation of the key endoplasmic reticulum stress kinases IRE1alpha and PERK (PKR-like ER kinase) and induction of their downstream targets. There was a similar response to MPP+ and rotenone, but not to other apoptotic initiators. As evidence that endoplasmic reticulum stress contributes to neuronal death, sympathetic neurons from PERK null mice in which the capacity to respond to endoplasmic reticulum stress is compromised were more sensitive to 6-hydroxydopamine. Our findings, coupled with evidence from familial forms of Parkinson's disease, raise the possibility of widespread involvement of endoplasmic reticulum stress and the unfolded protein response in the pathophysiology of this disease.
Article
The intracellular redox status is a tightly regulated parameter which provides the cell with an optimal ability to counteract the highly oxidizing extracellular environment. Intracellular redox homeostasis is regulated by thiol-containing molecules, such as glutathione and thioredoxin. Essential cellular functions, such as gene expression, are influenced by the balance between pro- and antioxidant conditions. The mechanism by which the transcription of specific eukaryotic genes is redox regulated is complex, however, recent findings suggest that redox-sensitive transcription factors play an essential role in this process. This review is focused on the recent knowledge concerning some eukaryotic transcription factors, whose activation and DNA binding is controlled by the thiol redox status of the cell.
Article
Protein synthesis and the folding of the newly synthesized proteins into the correct three-dimensional structure are coupled in cellular compartments of the exocytosis pathway by a process that modulates the phosphorylation level of eukaryotic initiation factor-2α (eIF2α) in response to a stress signal from the endoplasmic reticulum (ER). Activation of this process leads to reduced rates of initiation of protein translation during ER stress. Here we describe the cloning of perk, a gene encoding a type I transmembrane ER- resident protein. PERK has a lumenal domain that is similar to the ER- stress-sensing lumenal domain of the ER-resident kinase Ire1, and a cytoplasmic portion that contains a protein-kinase domain most similar to that of the known eIF2α kinases, PKR and HRI. ER stress increases PERK's protein-kinase activity and PERK phosphorylates eIF2α on serine residue 51, inhibiting translation of messenger RNA into protein. These properties implicate PERK in a signalling pathway that attenuates protein translation in response to ER stress.