Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2014 Mar 17.
Published in final edited form as: Ann N Y Acad Sci. 2010 Feb;1187:101–121. doi: 10.1111/j.1749-6632.2009.05141.x

Amphetamine toxicities Classical and emerging mechanisms

Bryan K Yamamoto 1, Anna Moszczynska 1, Gary A Gudelsky 2
PMCID: PMC3955986  NIHMSID: NIHMS571323  PMID: 20201848

Abstract

The drugs of abuse, methamphetamine and MDMA, produce long-term decreases in markers of biogenic amine neurotransmission. These decreases have been traditionally linked to nerve terminals and are evident in a variety of species, including rodents, nonhuman primates, and humans. Recent studies indicate that the damage produced by these drugs may be more widespread than originally believed. Changes indicative of damage to cell bodies of biogenic and nonbiogenic amine–containing neurons in several brain areas and endothelial cells that make up the blood–brain barrier have been reported. The processes that mediate this damage involve not only oxidative stress but also include excitotoxic mechanisms, neuroinflammation, the ubiquitin proteasome system, as well as mitochondrial and neurotrophic factor dysfunction. These mechanisms also underlie the toxicity associated with chronic stress and human immunodeficiency virus (HIV) infection, both of which have been shown to augment the toxicity to methamphetamine. Overall, multiple mechanisms are involved and interact to promote neurotoxicity to methamphetamine and MDMA. Moreover, the high coincidence of substituted amphetamine abuse by humans with HIV and/or chronic stress exposure suggests a potential enhanced vulnerability of these individuals to the neurotoxic actions of the amphetamines.

Keywords: amphetamine, methamphetamine, MDMA, neurotoxicity, apoptosis, excitotoxicity, neuroinflammation, proteasome, ubiquitination, neurodegeneration, drug abuse

Introduction

Methamphetamine (METH) and its derivative, 3,4-methylenedioxymethamphetamine (MDMA), are widely abused psychostimulant drugs. The acute effects of these drugs include euphoria, alertness, decreased appetite, increased locomotor activity, and hyperthermia. Long-term abuse of METH and MDMA may result in psychosis, aggressiveness, and neurotoxicity. METH in particular has a very high abuse potential owing primarily to its strong euphoric properties. According to the recent National Institute on Drug Abuse (NIDA) reports1-3 the abuse of METH and MDMA is an extremely serious and growing problem in the U.S. and worldwide. METH and MDMA use among significantly diverse populations has been documented. For instance, young adults who attend “raves” or private clubs are increasingly using amphetamines. METH use is also high among persons infected with human immunodeficiency virus (HIV).4 Although the acute effects of these drugs are relatively well known, the long-term consequences and possible neurotoxicities associated with the administration of these drugs are unclear.

Amphetamines are substrates for transporters associated with the uptake of the biogenic amines dopamine (DA), norepinephrine (NE), and serotonin (5-HT). They either diffuse into or are taken up by nerve terminals via these transporters and subsequently cause a reverse transport of monoamines from the cytoplasm into the synaptic cleft. Amphetamines also promote DA and 5-HT release from storage vesicles and prevent the uptake into vesicles, thus increasing the cytoplasmic concentrations of the neurotransmitter and making them more readily available for reverse transport. In addition, the amphetamines also increase synaptic levels of monoamines by inhibiting their reuptake.5-8 The net result of the acute action of the amphetamines is an increased neurotransmission of DA, 5-HT, and NE. METH and MDMA differ in their affinities for monoamine transporters. MDMA has a greater affinity for the 5-HT transporter (SERT) versus the DA transporter (DAT) than amphetamine or METH.9 Consequently, MDMA causes a greater release of 5-HT than DA. In addition, the substituted amphetamines also increase the release of glutamate (GLU),10-12 which probably contributes to the neurotoxicity profiles of these drugs.

In rodents and nonhuman primates, administration of either a large single dose or repeated high doses of METH or MDMA produces long-lasting deficits in markers of DA and 5-HT nerve terminals (i.e., the levels of a neurotransmitter, its metabolites, biosynthetic enzymes, receptors, and transporters)13-20 while sparing NE terminals.17,21 Amphetamines also produce astrogliosis,22-24 and METH18 but not MDMA21 displays morphological signs of axonal degeneration. Early studies have shown that METH most severely affects DA terminals in the striatum,13,16,18,25,26 whereas DA terminals in the nucleus accumbens, olfactory bulb, frontal cortex, and hypothalamus are minimally affected or unaffected.15, 16 The reasons for this difference are unclear but could be related to the varied densities of DAT in these regions. In contrast to DA terminals, 5-HT terminals in various brain regions including hippocampus, prefrontal cortex, amygdala, and striatum are equally sensitive to the toxic effects of METH.15, 16,20,27 MDMA differs from METH in that it is selectively neurotoxic to 5-HT terminals in multiple brain areas in rodents and nonhuman primates19,28-32; however, it can produce DA deficits in mice.24

A persistent reduction in most DA markers33-37 and SERT38,39 also has been observed in human chronic METH users. Similarly, decreases in SERT have been observed in multiple brain regions in chronic MDMA users.40 Because of many animal studies and more recent studies on humans suggesting that the amphetamines have long-term consequences, efforts have been directed toward the understanding of the mechanisms that contribute to the neurotoxicity of the amphetamines. This review will examine the new characteristics and emerging mechanisms purported to contribute to the neurotoxic profiles of the substituted amphetamines, METH and MDMA.

Classical aspects of METH and MDMA toxicity

Studies on the toxicity of METH and MDMA to monoaminergic terminals indicate that amphetamine toxicity involves the occurrence of at least three acute events: an increase in extracellular and intracellular DA, an increase in extracellular GLU, and hyperthermia. The major classical molecular mechanisms by which these events subsequently produce long-term effects include oxidative stress, excitotoxicity, and mitochondrial dysfunction. These mechanisms interact and result in the augmentation of their consequences. Those studies have been reviewed previously41-48 and therefore will not be discussed in detail. This review, however, will focus on new and emerging aspects that in combination with the more classic mechanisms summarized in the following, broaden the scope of the pharmacological action of the amphetamines and contribute to their long-term toxicity.

Oxidative stress

Several studies using animal models have supported the involvement of oxidative stress in METH and MDMA neurotoxicity (reviewed in Refs. 44 and 48). For instance, METH and MDMA produce reactive oxygen species (ROS) and reactive nitrogen species (RNS) and lipid peroxidation products. Neurotoxic effects of amphetamines can be attenuated by free radical scavengers and antioxidants or overexpression of antioxidant enzymes. Both METH and MDMA decrease the levels of antioxidants in DAergic and/or 5-HTergic terminals. The presence of oxidative stress has also been documented in human METH users.49,50 Research on mechanisms leading to amphetamine-mediated oxidative stress indicate that an early event in METH toxicity is an increase in intracellular DA levels resulting from amphetamine-mediated disruption of vesicular proton gradient and vesicular monoamine transporter function.46 This is followed by an overproduction of toxic metabolites of DA oxidation, including free radicals and quinones.51,52 For MDMA, which is neurotoxic only to 5-HT terminals, it is believed that DA-derived ROS are generated in 5-HT terminals either after SERT-mediated uptake of released DA53 or by the synthesis of DA from tyrosine.54 Alternatively, toxic metabolites of 5-HT oxidation or MDMA itself can also mediate MDMA toxicity.55-57

Excitotoxicity

Excitotoxicity includes a succession of several events: excessive GLU release, activation of GLU receptors, increase in intracellular calcium levels, activation of a variety of calcium-dependent enzymes, generation of free radicals and nitric oxide (NO), and activation of apoptotic pathways, culminating in failure of cellular organelles, such as mitochondria and endoplasmic reticulum (ER), breakdown of cytoskeletal proteins, and DNA damage.58-60 Several studies support a role for excitotoxicity in mediating METH neurotoxicity to striatal terminals. For example, high-dose METH causes a release of GLU in rat striatum10,61 via activation of the striatonigral pathway.62 Inhibition of this release protects against METH toxicity to terminals.62 Agonists of metabotropic GLU receptor 5 (mGluR5)63 and inhibitors of NO synthase (NOS)64 attenuate METH toxicity to striatal DA terminals independently from hyperthermia. Increases in striatal levels of nitrate65 and 3-nitrotyrosine66 suggest that high-dose METH increases the levels of NO. METH increases breakdown of microtubule-associated protein, tau,67 and another cytoskeletal protein, spectrin,68 in rat striatum in vivo and cortical neurons in vitro.69 A role for excitotoxicity in mediating MDMA toxicity is less clear.48 Nevertheless, the mechanism by which excitotoxicity mediates the toxicity of the amphetamines appears to be NO-mediated nitration of proteins associated with DA and 5-HT terminals.48

Mitochondrial function

Administration of both METH and MDMA impairs mitochondrial function. More specifically, toxic doses of METH inhibit mitochondrial electron transport chain enzyme complexes, complex I,70 complex II–III,71 and complex IV,72 in the striatum and other DA-containing brain areas. High-dose MDMA has been shown to decrease mitochondrial complex I–II in rat striatum44 and complex IV in rat striatum, nucleus accumbens, and substantia nigra.72 In addition, MDMA causes oxidative stress in mitochondria and deletions in mitochondrial DNA coding for complex I and IV in several brain areas.73 A correlation between impairment of mitochondria and amphetamine toxicity to monoaminergic terminals has been provided by several studies. For example, coadministration of METH74 or MDMA75 with an inhibitor of energy metabolism synergistically depleted striatal DA or 5-HT, respectively. Conversely, coadministration of amphetamines with energy substrates attenuated the neurotoxicity to DA and 5-HT nerve endings.75,76 The underlying mechanism of the impairment of mitochondrial function appears to involve increases in ROS and RNS64 and/or increases intracellular calcium,43,44,48 which may be mediated by GLU.

Hyperthermia

Hyperthermia occurs after the administration of high doses of both METH and MDMA,77-79 and its occurrence is important for development of amphetamine neurotoxicity to DA and 5-HT terminals. For example, multiple injections of high-dose METH at room temperature produced a significant depletion of DA in the striatum; however, equivalent doses of METH administered in a cold environment blocked striatal DA and 5-HT depletions in mice.78 Similarly, MDMA toxicity to 5-HT terminals during hyperthermic and hypothermic conditions also can be enhanced and attenuated, respectively.79 Hyperthermia by itself does not decrease striatal DA levels in rodents.80 Instead, it is envisioned to enhance the enzymatic and/or nonenzymatic reactions initiated by high-dose METH or MDMA treatment. Hyperthermia might interact with other known mediators of METH neurotoxicity, such as increased GLU neurotransmission and oxidative stress. In fact, GLU receptor antagonists, such as MK-801, have been shown to reduce body temperature and provide neuroprotection.81-83 Similarly, inhibition of METH-induced hyperthermia decreases the formation of ROS in the striatum that, in turn, attenuates the damage to DA terminals.84

New and emerging aspects of the toxicity of amphetamines

As noted in the preceding, a classic mechanism underlying the toxicity of the amphetamines involves oxidative stress to DA and 5-HT terminals. However, a more current and emerging focus has been on the toxic effects of the amphetamines to nonmonoaminergic cell bodies, as originally suggested and demonstrated by several groups in the 1980s and 1990s.19,21,85-90

Emerging mechanisms that may be related to both terminal and cell body damage produced by the amphetamines are processes linked to excitotoxicity, inflammation, proteolytic/proteasomal dysfunction, apoptosis, alterations in trophic support, HIV infection, and the influence of environmental stress. The review that follows will cover this current literature while incorporating these mechanisms into our understanding of the classic processes involved in damage to DA and 5-HT terminals.

Most studies of the mechanisms of METH and MDMA neurotoxicity have, until recently, investigated the toxic effects on DA and 5-HT terminals. Despite significant damage to these terminals, METH and MDMA appear to spare the monoamine-containing cell bodies from which these terminals arise.18,91 Some studies, however, have reported that amphetamines could produce neurodegeneration of nonmonoaminergic cell bodies in several brain areas. For instance, high binge doses of METH87 and MDMA19 produce a loss of DA cells in the substantia nigra of mice and a loss of 5-HT cells in dorsal raphe nucleus in nonhuman primates, respectively. In addition, METH, MDMA and d-amphetamine damage a population of non-monoaminergic neurons and their processes in rat parietal cortex (somatosensory cortex).21,85,88,90,92 In mice, high-dose METH leads to cell death in a variety of brain areas including the striatum, cortex (frontal, parietal, and piriform), indusium griseum, medial habenular nucleus, hippocampus, tenia tecta, and fasciola cinerea.93,94 More recently, a low dose of METH has been shown to damage cell bodies in rat prefrontal cortex of behaviorally sensitized rats,95 whereas an escalating binge dose of METH damages pyramidal neurons in the frontal cortex, CA3 and dentate gyrus regions of the hippocampus, and calbindin interneurons of the striatum.96 Finally, there are several more recent reports of amphetamine toxicity to DA-containing neurons and their terminals in mouse olfactory bulb97,98 and rat retina.99

The mechanisms underlying the damage to cell bodies have yet to be elucidated. Nevertheless, inflammatory cytokines, the ubiquitin proteasome system (UPS), environmental stress, HIV, neurotrophic factors, and apoptotic proteins have recently emerged as mediators of the toxicity of amphetamines that may explain both the terminal and somatic degeneration observed after exposure to these drugs.

Excitotoxicity to nonmonoaminergic cell bodies

Studies of mechanisms underlying METH toxicity to neuronal cell bodies are relatively recent and indicate that an early event in METH toxicity to non-monoaminergic striatal and somatosensory cortical neurons might be a release of GLU that initiates a chain of events culminating in apoptosis.

Striatal GABA neurons and interneurons

Approximately 90% of the neurons in the striatum are GABAergic medium spiny projection neurons, which contain either substance P and dynorphin or enkephalin. The remaining 10% are interneurons, of which the GABA-parvalbumin, somatostatin (SST)/NOS, and cholinergic interneurons are the most prevalent.100 It is the GABA neurons that express enkephalin and parvalbumin in the rat and mouse striatum that are damaged by METH.101-103

Excitotoxicity mediated by GLU was suggested by several studies as a mechanism for cell death produced by METH. Along these lines, striatal neurons express GLU receptors,104,105 and METH causes a release of GLU in rat striatum10,61 via activation of the striatonigral pathway.62 Indirect evidence suggests that METH produces an increase in NO in the striatum65,66 and induces toxicity to GABAergic neurons via mitochondrial dysfunction and ER stress,106 both of which are mediated by GLUergic and calcium-dependent mechanisms. Specifically, ER stress involves the rapid activation of calcium-dependent calpain and its substrate caspase-12, as well as an increase in the expression of other proteins indicative of ER dysfunction, namely, GRP78, BiP, and CHOP.106 In parallel, METH causes a release of cytochrome c, smac/DIABLO, and apoptosis-inducing factor (AIF)107 from mitochondria to the cytosol, presumably the result of damage to mitochondria. In fact, mitochondrial dysfunction has been shown to mediate METH-induced apoptosis in an immortalized rat striatal cell line.108 METH also activates the calcium-dependent protease, calpain, to cause spectrin proteolysis.68 These events are in conjunction with the activation of several effector caspases and prodeath transcription factors, including the NFAT-family transcription factors103 that lead to apoptosis. Along similar lines, GLU excitotoxicity produces caspase-dependent and caspase-independent (AIF mediated) apoptosis in neuronal cells in vitro.109,110 Thus, the convergence of GLU and calcium-dependent and -independent mechanisms that also involve the mitochondria can mediate the observed death to striatal cells after METH exposure.

METH-induced apoptosis of striatal GABA neurons also depends on DA. Administration of DA receptor antagonists prevents DA terminal damage and apoptosis in mouse striatum,111 whereas administration of a D1 receptor antagonist decreases the number of terminal deoxynucleotidyltransferase-mediated dUTP-biotin nick end labeling–positive cells and inhibits upregulation of NFAT transcription factors in rat striatum.103 DA may contribute indirectly to excitotoxicity in GABA neurons via regulation of (i) extracellular GLU,61,112 (ii) N-methyl-d-aspartate (NMDA) receptors,113,114 and/or (iii) substance P signaling115 as well as via ROS formation in the extracellular space.116,117 The activation of NMDA receptors by GLU to induce NO in SST/NOS interneurons could further increase the release of GLU and DA in the striatum,118 resulting in a feed-forward mechanism that promotes METH toxicity.

Substance P also may contribute to METH-induced apoptosis of striatal GABA neurons. Pharmacological blockade of the substance P receptor, neurokinin 1 (NK-1R), attenuates METH-induced damage to DA terminals119 and neuronal apoptosis in the striatum.94 Deletion of the NK-1R–expressing interneurons (SST/NOS and cholinergic) from the striatum prevents METH-induced apoptosis but does not prevent DA terminal damage.120 Most NK-1R–expressing terminals form asymmetric synapses with dendrites and dendritic spines,121 suggesting that substance P modulates excitatory GLUergic neurotransmission. Therefore, substance P may mediate neuronal apoptosis via regulation of GLU release from its afferents and/or via activation of NOS. Collectively, the available data suggest that damage to striatal GABA neurons is mediated by excitotoxicity.

Somatosensory cortex

Administration of amphetamines can cause degeneration of a population of nonmonoaminergic cortical neurons and their processes in layers II/III and IV of rat primary somatosensory cortex.21,22,85,88,90 The damaged neurons have been identified as pyramidal or stellate cells88,90,92 confined to the cytochrome oxidase–rich areas.92 The morphology, localization, absence of monoaminergic markers,85 and substantial decrease in GLU immunoreactivity in affected areas22 suggest that these neurons are GLUergic. The cortical damage produced by METH occurs via an excitotoxic mechanism, as evidenced by the findings that METH induces a rapid increase in NMDA receptor binding122 and that NMDA receptor antagonism89 or removal of excitatory sensory input from rat whiskers to somatosensory cortex123 decreases the rapid Fluoro-Jade staining in this cortical area.

Overall, the toxicity induced by amphetamines appears to be more widespread than originally believed and includes damage to cell bodies as well as terminals. An increase in extracellular GLU may mediate the damage to both targets, but the terminals may be more susceptible because of the occurrence of DA-mediated intracellular oxidative stress and other factors, such as proinflammatory mediators that converge upon the terminals.

Inflammation

METH-induced GLU release may also serve to activate inflammatory mediators of METH toxicity to monoaminergic as well as nonmonoaminergic neurons. For instance, GLU receptor antagonism decreases,124,125 whereas GLU receptor stimulation increases, microglial activation. Thus, activation of GLU receptors increases the production of proinflammatory cytokines interleukin 1β (IL-1β), tumor necrosis factor α (TNF-α), and IL-6.126-129 In turn, cytokines can increase extracellular GLU levels by either inhibition of GLU uptake130 or an increase in GLU release from activated microglia.131 Thus, the interactions between the cytokines and GLU may form a feed-forward cycle to promote neurotoxicity.

METH and MDMA trigger inflammation in brain areas that exhibit DA and 5-HT terminal degeneration. METH elicits microglial activation in rat and mouse striatum132-136 rat cortex (including somatosensory and frontal cortices)96,136-138 and hippocampus137,139 but not in areas where DA levels are unaffected by METH, such as substantia nigra.132,134 METH-induced microglial activation occurs in the rat somatosensory cortex138 but not mouse somatosensory cortex.140 Microglial activation has also been detected in the brains of human METH users141 and nonhuman primates administered METH.142

Microglia might be involved in the toxic effects of METH to DA terminals and GABA neurons via a release of proinflammatory and prooxidative stress molecules into the extracellular space. In mouse striatum, a single dose of METH increased mRNA levels of IL-6, TNF-α, and IL-1α.143,144 Interestingly, METH-induced microglial activation appears to depend on newly synthesized and released DA. Thus, a decrease in DA synthesis or an increase in cytosolic DA can decrease and increase, respectively, microglial activation in mouse striatum.140 These results are in conflict with the finding that DA itself inhibits the activation of microglia in vitro.145 In contrast, DA quinones are powerful activators of microglia.146 Therefore, it can be envisioned that nonenzymatic degradation of DA that is released after METH results in production of DA quinones147 and, in turn, activates striatal microglia to provide a proinflammatory stimulus for neurodegeneration of both DA terminals and striatal cell bodies.

The ability of MDMA to induce microglial activation is more equivocal. For example, MDMA-induced microgliosis was detected in male133 but not female136 mouse striatum and was absent in rat striatum or cortex.137 In contrast, MDMA increased production of IL-1β in rat frontal cortex,148,149 whereas intracerebroventricular administration of IL-1β potentiated MDMA-induced 5-HT toxicity in the cortex.146 An explanation for the varied results and the limited potential of MDMA to induce microglial activation might stem from the fact that MDMA also has an immunosuppressive action that involves suppression of proinflammatory cytokines via an increase in IL-10 production.150 However, central injections of proinflammatory factors interferon γ151 and lipopolysaccharide152 before or immediately after METH administration can attenuate METH toxicity to striatal DA terminals through a decrease in extracellular GLU concentrations,153,154 or a decrease in extracellular and intracellular DA levels.155 These data suggest that the initial and acute upregulation of inflammatory cytokines might be protective by upregulating the buffering capacity of either neurons or glia to counter the excessive and prolonged increases in GLU or DA. In contrast, the neurotoxic effects of the cytokines may be related to the magnitude of their increase after the induction of the GLU excitotoxicity cascade.

Astrocytes can also play a role in substituted amphetamine toxicity through modulation of GLU-mediated excitotoxicity and inflammation. Astrocytes regulate extracellular concentrations of GLU, mainly by uptake of the neurotransmitter. They can also release GLU upon activation through an increase in intracellular calcium.156 For METH, the activation of cortical astrocytes appears to be caused by GLU release and protein kinase C activation and is inhibited by GLU receptor antagonists.157 Under normal physiologic conditions, however, astrocytes suppress microglial activation through the release of anti-inflammatory cytokines and neurotrophic factors.124 For example, astrocytes suppress microglial activation by releasing the anti-inflammatory cytokines transforming growth factor β (TGF-β) or IL-10.158,159 On the other hand, IL-1 and TNF-α are known to be involved in the development of central nervous system inflammation through, among other factors, the induction of chemokines from astrocytes.160 Therefore, astrocytes can mediate either an increase or decrease in inflammation depending on the cytokine that is released. More information is needed to identify the specific conditions under which astrocytes may be pro- or anti-inflammatory.

Oxidative stress plays a key role in substituted amphetamine toxicity, as noted in the preceding. Moreover, oxidative stress and inflammation are intimately linked,48,161 but the exact relationship between the two in mediating amphetamine toxicity is unclear. However, edaravone, a free radical scavenger, blocked METH toxicity to DAergic terminals, the increase in protein oxidation as evidenced by 3-nitrotyrosine immunoreactivity, and the activation of astrocytes, but it did not affect the activation of microglia,162 suggesting that METH-induced activation of microglia and inflammation is independent of oxidative stress. In fact, a variety of intracellular signaling molecules that have been identified to be involved in METH toxicity, such as GLU, DA-quinones, matrix metalloproteinases (MMPs), substance P, and α-synuclein, can induce microglial activation124,146,161 independent of the formation of free radicals. However, oxidative stress can activate microglia to release MMP-3 and α-synuclein,161 thus providing another means by which microglia are activated. The self-perpetuating cycle of oxidative stress and inflammation is further promoted by the diminished capacity of microglia under prooxidant conditions to store iron,163 thereby potentially exacerbating Fenton reaction and iron-dependent oxidative stress that mediates METH toxicity.164 Taken together, activated microglia can initiate, exacerbate, and perpetuate METH neurotoxicity.

The time courses of microglial activation and increases in inflammatory markers vary relative to indices of neurotoxicity. For example, microglial activation in the striatum occurs 1–3 days after METH132-136 and precedes degeneration of DAergic terminals.135,138 On the other hand, rat striatal GABA-enkephalin neurons exhibit an upregulation of FasL, a member of the TNF superfamily of cytokines, that appears as soon as 2–4 h after one high dose of METH.103 Interestingly, Bowyer et al.135 reported the relatively early appearance of phagocytic microglia with Fluoro-Jade C–labeled striatal neurons in mice 12–24 h after one high dose of METH. These findings suggest that damage to striatal cell bodies appears before the neurodegeneration of DA terminals, but it is unknown whether damage to GABA neurons plays a causal role in DA terminal degeneration or is simply an independent event. Regardless of the temporal relationship between the activation of microglia and the appearance of neurodegeneration, microglia are emerging as new players in the toxicity of the amphetamines that, at the minimum, perpetuate excitotoxic events that eventually lead to neurodegeneration. Although factors that promote and perpetuate toxicity have historically been the focus of studies on the neurotoxic amphetamines, more recent efforts have been directed toward endogenous protective systems, such as the UPS, and neurotrophic factors that are emerging as targets whose functions may be compromised by these drugs.

Ubiquitin proteasomal system

Recent studies have shown that the substituted amphetamines promote the dysregulation of the UPS, which may further contribute to neurotoxic and apoptotic events. A decrease in the activity of the UPS can lead to the accumulation of unwanted proteins and has been implicated in the etiology of various neurodegenerative disorders.165 Furthermore, identified mediators of amphetamine neurotoxicity described in the foregoing, such as GLU-induced NOS activity, mitochondrial dysfunction, and oxidative stress, are known to affect or be affected by the UPS. Inhibition of the proteasome can block inducible NOS degradation166 and potentially increase NO production, NO-mediated nitrosative stress, damage to the ubiquitin ligase, parkin,167 and protein misfolding,168 all of which can potentiate the inhibition of the proteasome.60,169,170 Conversely, proteasomal inhibition can produce an impairment of the mitochondria and a release of proapoptotic proteins.171 Therefore, on the basis of the overlap between mediators of amphetamine toxicity and events associated with the UPS, these studies suggest the view that amphetamines can lead to unwanted accumulation of protein through a dysregulation of the UPS.

Administration of high METH or MDMA doses causes formation of intracellular inclusions in the nucleus of medium-sized GABA neurons and cytoplasm of neurons of the substantia nigra pars compacta of mice.117,172-176 The inclusions in GABA neurons stain for ubiquitin and enzymatic components of the UPS (including E3 ligase parkin) but usually not for α-synuclein, whereas inclusions found in substantia nigra neurons stain for α-synuclein, a hallmark of Lewy bodies frequently observed in Parkinson’s disease and other degenerative disorders. Occurrence of ubiquitinated inclusions was also reported in the substantia nigra of 37 subjects who abused METH.177 The specific cause of the inclusions is unknown, but neuronal inclusions can occur when the UPS is inhibited pharmacologically.178,179 Moreover, oxidative stress commonly leads to inclusion formation, and the inclusions produced by METH, MDMA, and MPTP180 are ultrastructurally similar to those produced by DA-mediated oxidative stress.117,172,173 In addition, inclusion formation is decreased upon administration of antioxidant/iron-chelating agent, S-apomorphine.175

It is hypothesized that striatal neuronal inclusions are a consequence of amphetamine-mediated increases in DA release followed by overstimulation of DA D1 receptors.117,181 The underlying mechanism is thought to involve β-arrestin that is present together with ubiquitin in inclusions after exposure of PC12 cells to METH.182 Because β-arrestin is involved in the internalization of DA and mGlu5 receptors,183-185 it suggests the possibility that activation of these receptors contributes to the formation of inclusions in striatal GABA neurons. In addition, DA and non–DA-derived ROS might diffuse to GABA neurons and inhibit the function of proteasome.117

α-Synuclein, a presynaptic protein involved in various degenerative disorders including Parkinson’s disease, might also contribute to DA-dependent inclusion formation in nigral cells after toxic amphetamine administration. Increases in α-synuclein levels are known to be toxic to DA neurons in vitro186 and in vivo.187 Administration of METH and MDMA increases expression of α-synuclein in DA neurons in the substantia nigra of mice.176 It is possible that covalent modification of α-synuclein by DA-derived quinone188,189 after amphetamine administration promotes the formation of toxic α-synuclein aggregates.190

Misfolded protein aggregates or damaged organelles that accumulate cannot be degraded by the UPS. This function is reserved for the lysosomal system and the process of microautophagy. Autophagic vacuole formation by the lysosomal system will remove oxidized and damaged organelles (such as mitochondria) and misfolded protein aggregates produced by METH. Conversely, inhibition of autophagy is deleterious to cells because of a diminished ability to clear α-synuclein aggregates after METH exposure, eventually resulting in caspase-dependent cell death.191

Now it is unclear whether a dysfunction of the UPS system is a consequence or a cause of the toxicity to the amphetamines. It remains to be determined if the excitotoxic, oxidative, and inflammatory mediators discussed earlier directly target the UPS and thus disrupt the normal, ongoing removal of unwanted proteins to ultimately produce the demise of cell bodies and terminals. A likely scenario, however, is that the damage produced by the amphetamines is ultimately dependent upon the balance of factors that promote toxicity (e.g., excitotoxic glutamatergic events, prooxidant processes, inflammation) and endogenous protective systems (such as the UPS), antioxidants, and growth-promoting molecules (such as neurotrophic factors) that can be targeted by toxic insults.

Neurotrophic factors

Several neurotrophic factors can act as survival-promoting proteins. These factors include neurotrophins, glial cell line–derived neurotrophic factor (GDNF) family, and TGF-α.192 Neurotrophins comprise a family consisting of four members: nerve growth factor (NGF), brain-derived neurotrophic factor (BDNF), neurotrophin-3 (NT-3), and NT-4.193 The GDNF family includes GDNF, neurturin, artemin, and persephin.194 Of these, GDNF was the first neurotrophic factor demonstrated to protect DA terminals against METH neurotoxicity in animal models.195-197 Recently, neurturin and artemin are two other GDNF family members that have been shown to protect against METH toxicity in the rat.198 The protective action of GDNF might involve regulation of DA release196 and/or attenuation of METH-mediated oxidative stress: GDNF has been shown to upregulate striatal antioxidant enzymes in vivo199 and reduces levels of free radicals in cultured mesencephalic neurons.200 Conversely, a study by Boger et al.201 demonstrated that a partial GDNF gene deletion increased the susceptibility of mice to METH neurotoxicity during young adulthood and increased age-related deterioration of motor behavior and DA function.

In contrast to the protective effects of several of the growth factors, ciliary neurotrophic factor provides no protection against METH toxicity to DA neurons.198 In non-DAergic primary rat cortical neurons, METH-triggered apoptosis was attenuated by BDNF through the PI3K–Akt but not MAPK–Erk pathway.202 Overall, these results indicate that GDNF may play a greater role in protecting DA terminals against METH toxicity, whereas BDNF may be more potent in the protection of non-DA neurons.

Blood–brain barrier dysfunction

Recent studies have begun to demonstrate another emerging consequence of exposure to high doses of the amphetamines. Administration of MDMA or METH has been shown to increase blood–brain barrier (BBB) permeability in rodents. MDMA-induced damage to the BBB was observed in the striatum and hippocampus.43 Moderate to high doses of METH disrupt the BBB in several brain regions, including the cortex, hippocampus, thalamus, hypothalamus, cerebellum, amygdala, and striatum135,203-206 that, in turn, are augmented by hyperthermia and seizures.135,205,206 Although it is unclear whether there is a relationship between the damage to the BBB and the damage to neurotransmitter systems, the damage to the BBB appears to contribute to striatal neuron degeneration rather than DA terminal damage.135

The mechanisms underlying the damage to the BBB produced by the amphetamines have not been elucidated. However, the amphetamines can cause hyperthermia77-79 and produce ROS,164 both of which trigger BBB breakdown.207 Consistent with these findings, administration of antioxidants attenuates the effects of amphetamines on the BBB205 and further implicates oxidative stress in the effects of amphetamine at the BBB.

Another possible mediator of the damage to the BBB could be the MMPs, whose functions are to degrade tight junction proteins208 present in the extracellular matrix that supports the endothelial cells of the BBB.209 METH has been shown to increase the release of MMP-1 and the MMP activator, urokinase plasminogen activator, in neuron–astrocyte cocultures.210 METH also alters the expression of several tight junction proteins and increases the permeability of brain-derived primary microvascular endothelial cells.211 High doses of METH also increase the levels of MMP-9 in the hippocampus.212 The activation of the MMPs is thought to occur through several mechanisms, including oxidative stress213 and cytokine production.214,215 Collectively, these findings suggest that amphetamine-mediated oxidative stress followed by activation of MMPs and breakdown of tight junctions mediate BBB disruption. Because both activation of MMPs216 and oxidative stress161 can induce inflammation, these events in conjunction with the MMPs could be accompanied by an increase in cytokine production within microglia217 to perpetuate damage and increase the permeability of the BBB. The consequences of the breakdown in the BBB are widespread and may enhance the vulnerability of the brain to toxins and infection, such as those produced by HIV. This and the fact that the BBB breakdown can be mediated by other toxic mechanisms, such as oxidative stress, neuroinflammation, and hyperthermia, suggests it as a new and important contributing factor to the toxicity of amphetamines.

Interactions of amphetamines and HIV

The comorbidity of drug abuse and HIV infection is well known. Early findings of decreases in postmortem levels of DA and homovanillic acid in the caudate nucleus and substantia nigra neuron degeneration in HIV patients suggested that HIV infection might damage nigrostriatal DA neurons.218,219 It was subsequently found that HIV injured not only these two regions but also other brain areas, such as prefrontal cortex, parietal cortex, nucleus accumbens, and hippocampus, thus increasing the vulnerability of these areas to METH toxicity in HIV-infected METH users.220 Along these lines, intrastriatal injections of the HIV protein, Tat, damage both efferent and afferent projections of the rat striatum and/or substantia nigra neurons,221-224 common targets of the toxic effects of METH.

Similar mechanisms mediate the toxicity to the amphetamines and HIV. Oxidative stress, mitochondrial dysfunction, inflammation, and caspase-dependent neuronal apoptosis220,225 all contribute. Similar to METH, Tat potentiates GLU toxicity via interaction with the NMDA receptor,226 causes neuronal cell death via activation of the D1 receptor,227 and decreases DAT function.228

The combined effects of HIV and chronic METH exposure converge to produce neuronal damage and inflammation. N-Acetylaspartate, myoinositol, and brain metabolites are increased more in HIV-positive METH abusers than in HIV patients with no METH abuse.229 Langford et al.230 found decreased blood flow; an increased microglial response; and more pronounced losses of synaptic vesicle–associated protein, synaptophysin, and the interneuron-associated protein, calbindin, in HIV-infected METH abusers relative to HIV-infected non-METH abusers. Similarly, Chana et al.231 reported that HIV-positive METH users have greater losses of frontal cortex calbindin and parvalbumin interneurons than do HIV non-METH abusers and that these effects are associated with cognitive impairment. In addition, METH has been shown to enhance HIV infection of macrophages, the primary target of the virus, and decrease IFN-α in these cells in vitro.232

The mediators of the damage produced by the combination of HIV and METH are being actively investigated. Tat and METH synergistically impair mitochondria in a variety of cellular targets, including DAergic neurons233; a non-DAergic, calbindin-positive hippocampal cell line234; and human fetal neurons.235 This effect on mitochondria is accompanied by oxidative stress and can be blocked by antioxidants.234,235 In HIV-positive rodent striatum, METH produces a synergistic increase in oxidative stress markers, expression of several inflammatory cytokines (e.g., IL-1α, IL-1β, and TNF-α), augmented activity of redox-responsive transcription factors,236,237 and toxicity to striatal DA terminals.224,235,238 These findings indicate that HIV infection increases susceptibility of DAergic and non-DAergic neurons to METH neurotoxicity. Moreover, oxidative stress, inflammation, and possibly excitotoxicity might interact to exacerbate toxicity in HIV-infected METH users.

Both METH and HIV increase permeability of the BBB via damage to tight junction proteins.211 In HIV-positive METH abusers, METH-induced increases in BBB permeability might facilitate an increased transport of HIV-infected leukocytes across the BBB. In fact, both METH and HIV protein gp120, alone and in combination, significantly increase transendothelial migration of immunocompetent cells across the BBB.211 Conversely, HIV-induced increases in BBB permeability might facilitate an increased transport of METH. Finally, METH may contribute to HIV-induced BBB breakdown by stimulating release and/or activation of MMPs. Levels of MMP-2, -7, and -9 are higher in cerebrospinal fluid of HIV-infected individuals,239,240 and both METH and Tat increase the release of MMPs in vitro.210 Overall, the common mechanisms underlying the toxic effects of METH and HIV appear to accurately predict an additive if not a synergistic damage to neurons and endothelial cells of the BBB. Therefore, the dangerous consequences of the comorbidity of amphetamine abuse with HIV infection can be extended to include potentiated and exacerbated damage to multiple cells in the central nervous system.

Interactions of amphetamines and environmental stress

The stress response involves a release of glucocorticoid hormones via activation of the hypothalamic–pituitary–adrenal axis as well as a release proinflammatory cytokines via activation of the immune system.241,242 In experimental animals, chronic stress potentiates the toxicity of neurotoxins243-247 and can cause neurodegeneration by itself.248 Chronic stress also exerts neurotoxic effects in humans.249,250 Several neurochemical effects are common to the amphetamines and stress and include oxidative stress, excitotoxicity, mitochondrial dysfunction, depletion of energy stores, increase in glucose utilization, inflammation, and hyperthermia.48,241,242 In fact, stress can potentiate METH-induced excitotoxicity247,251 and hyperthermia.252

Stress may also contribute to the toxic effects of the amphetamines through the mechanisms summarized in previous sections, such as trophic factor expression, UPS function, and HIV infections. For example, exposure to a variety of stressors decreases the levels of NGF in rat hippocampus.253 In an astroglial cell line, corticosterone reduces basal levels of NGF secretion and stimulated NGF secretion triggered by IL-1β and TGF-β1.254 In addition, corticosterone-induced cell death can be prevented by administration of BDNF255 or insulin-like growth factor256 in cultured rat hippocampal neurons. For UPS activity, the proteasome regulates glucocorticoid receptor activity via regulation of the trafficking of the receptor. Inhibition of the proteasome blocks glucocorticoid receptor translocation to the nucleus,257 which would increase expression and signaling of the receptor at the plasma membrane. Conway-Cambell et al.258 have demonstrated that the proteasome also regulates glucocorticoid receptor activity via the rapid degradation of the activated glucocorticoid receptor. In regard to its interaction with HIV, HIV-positive patients have increased hypothalamic–pituitary–adrenal axis activity259 that, in turn, can potentially increase the toxic effects of stress and the amphetamines as well as their combined exposures. Overall, there are multiple overlapping mechanisms between stress and the amphetamines that predict an augmentation of neurotoxicity produced by their combined exposure, such as that observed in individuals with posttraumatic stress disorder that have a high comorbidity with substance abuse.260

Concluding remarks

There is mounting evidence that the characteristics of amphetamine-induced toxicity extend beyond the selective damage to DA and 5-HT terminals to include neuronal and endothelial cell bodies. The underlying mechanisms have yet to be elucidated, and the consequences of this extended damage remain to be determined. However, the causes of the newly identified consequences to cell bodies most likely involve a convergence of excitotoxic, pro-teolytic, inflammatory, and bioenergetic processes that interact with and contribute to the previously established role of oxidative stress. Although basic experimental studies have provided clear, interpretable roles for each of these causative processes, we now know that each process does not occur in isolation. Moreover, the frequent comorbidities of the abuse of the amphetamines with other exposures, such as environmental stress, hyperthermia, and HIV infection, add to the complexity and severity of the toxicity. More studies are needed that take into account and model the more realistic scenario involving their concurrent exposures, comorbidities, and how they interact before effective therapeutic interventions can developed.

Acknowledgments

The work was supported by NIH DA07427, DA07606, DA16866, DA19486, and DA23085.

Footnotes

Conflicts of interest

The authors declare no conflicts of interest.

References

  • 1.NIDA. NIDA Notes. NIDA’s Latest Research Report Focuses on MDMA (Ecstasy) Abuse. 2005;19 http://www.drugabuse.gov/NIDA.notes/NNvol19N5/tearoff.html. [Google Scholar]
  • 2.NIDA. NIDA InfoFacts. Methamphetamine and MDMA (Esctasy) 2008 http://www.nida.nih.gov/DrugPages.
  • 3.NIDA. NIDA Community Drug Alert Bulletin–Methamphetamine. 2008 http://www.drugabuse.gov/MethAlert/MethAlert.
  • 4.Mitchell SJ, et al. Methamphetamine use and sexual activity among HIV-infected patients in care–San Francisco, 2004. AIDS Patient Care STDS. 2006;20:502–510. doi: 10.1089/apc.2006.20.502. [DOI] [PubMed] [Google Scholar]
  • 5.Kuczenski R, et al. Hippocampus norepinephrine, caudate dopamine and serotonin, and behavioral responses to the stereoisomers of amphetamine and methamphetamine. J Neurosci. 1995;15:1308–1317. doi: 10.1523/JNEUROSCI.15-02-01308.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Sulzer D, et al. Amphetamine redistributes dopamine from synaptic vesicles to the cytosol and promotes reverse transport. J Neurosci. 1995;15:4102–4108. doi: 10.1523/JNEUROSCI.15-05-04102.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Rothman RB, et al. Neurochemical neutralization of methamphetamine with high-affinity non-selective inhibitors of biogenic amine transporters: a pharmacological strategy for treating stimulant abuse. Synapse. 2000;35:222–227. doi: 10.1002/(SICI)1098-2396(20000301)35:3<222::AID-SYN7>3.0.CO;2-K. [DOI] [PubMed] [Google Scholar]
  • 8.Sulzer D, et al. Mechanisms of neurotransmitter release by amphetamines: a review. Prog Neurobiol. 2005;75:406–433. doi: 10.1016/j.pneurobio.2005.04.003. [DOI] [PubMed] [Google Scholar]
  • 9.Rothman RB, Baumann MH. Monoamine transporters and psychostimulant drugs. Eur J Pharmacol. 2003;479:23–40. doi: 10.1016/j.ejphar.2003.08.054. [DOI] [PubMed] [Google Scholar]
  • 10.Nash JF, Yamamoto BK. Methamphetamine neurotoxicity and striatal glutamate release: comparison to 3,4-methylenedioxymethamphetamine. Brain Res. 1992;581:237–243. doi: 10.1016/0006-8993(92)90713-j. [DOI] [PubMed] [Google Scholar]
  • 11.Rocher C, Gardier AM. Effects of repeated systemic administration of d-Fenfluramine on serotonin and glutamate release in rat ventral hippocampus: comparison with methamphetamine using in vivo microdialysis. Naunyn Schmiedebergs Arch Pharmacol. 2001;363:422–428. doi: 10.1007/s002100000381. [DOI] [PubMed] [Google Scholar]
  • 12.Cunningham MO, et al. Dual effects of gabapentin and pregabalin on glutamate release at rat entorhinal synapses in vitro. Eur J Neurosci. 2004;20:1566–1576. doi: 10.1111/j.1460-9568.2004.03625.x. [DOI] [PubMed] [Google Scholar]
  • 13.Seiden LS, Fischman MW, Schuster CR. Long-term methamphetamine induced changes in brain catecholamines in tolerant rhesus monkeys. Drug Alcohol Depend. 1976;1:215–219. doi: 10.1016/0376-8716(76)90030-2. [DOI] [PubMed] [Google Scholar]
  • 14.Hotchkiss AJ, Gibb JW. Long-term effects of multiple doses of methamphetamine on tryptophan hydroxylase and tyrosine hydroxylase activity in rat brain. J Pharmacol Exp Ther. 1980;214:257–262. [PubMed] [Google Scholar]
  • 15.Morgan ME, Gibb JW. Short-term and long-term effects of methamphetamine on biogenic amine metabolism in extra-striatal dopaminergic nuclei. Neuropharmacology. 1980;19:989–995. doi: 10.1016/0028-3908(80)90010-6. [DOI] [PubMed] [Google Scholar]
  • 16.Ricaurte GA, Schuster CR, Seiden LS. Long-term effects of repeated methylamphetamine administration on dopamine and serotonin neurons in the rat brain: a regional study. Brain Res. 1980;193:153–163. doi: 10.1016/0006-8993(80)90952-x. [DOI] [PubMed] [Google Scholar]
  • 17.Wagner GC, et al. Long-lasting depletions of striatal dopamine and loss of dopamine uptake sites following repeated administration of methamphetamine. Brain Res. 1980;181:151–160. doi: 10.1016/0006-8993(80)91265-2. [DOI] [PubMed] [Google Scholar]
  • 18.Ricaurte GA, et al. Dopamine nerve terminal degeneration produced by high doses of methylamphetamine in the rat brain. Brain Res. 1982;235:93–103. doi: 10.1016/0006-8993(82)90198-6. [DOI] [PubMed] [Google Scholar]
  • 19.Ricaurte GA, et al. (+/-)3,4-Methylenedioxymethamphetamine selectively damages central serotonergic neurons in nonhuman primates. JAMA. 1988;260:51–55. [PubMed] [Google Scholar]
  • 20.Seiden LS, Ricaurte GA. Neurotoxicity of methamphetamine and related drugs. In: Meltzer H, editor. Psychopharmacology: The Third Generation of Progress. Raven Press; New York, NY: 1987. pp. 359–366. [Google Scholar]
  • 21.Commins DL, et al. Biochemical and histological evidence that methylenedioxymethylamphetamine (MDMA) is toxic to neurons in the rat brain. J Pharmacol Exp Ther. 1987;241:338–345. [PubMed] [Google Scholar]
  • 22.Pu C, Broening HW, Vorhees CV. Effect of methamphetamine on glutamate-positive neurons in the adult and developing rat somatosensory cortex. Synapse. 1996;23:328–334. doi: 10.1002/(SICI)1098-2396(199608)23:4<328::AID-SYN11>3.0.CO;2-T. [DOI] [PubMed] [Google Scholar]
  • 23.O’Callaghan JP, Miller DB. Neurotoxicity profiles of substituted amphetamines in the C57BL/6J mouse. J Pharmacol Exp Ther. 1994;270:741–751. [PubMed] [Google Scholar]
  • 24.Johnson EA, O’Callaghan JP, Miller DB. Chronic treatment with supraphysiological levels of corticosterone enhances D-MDMA-induced dopaminergic neurotoxicity in the C57BL/6J female mouse. Brain Res. 2002;933:130–138. doi: 10.1016/s0006-8993(02)02310-7. [DOI] [PubMed] [Google Scholar]
  • 25.Preston KL, et al. Long-term effects of repeated methylamphetamine administration on monoamine neurons in the rhesus monkey brain. Brain Res. 1985;338:243–248. doi: 10.1016/0006-8993(85)90153-2. [DOI] [PubMed] [Google Scholar]
  • 26.Woolverton WL, et al. Long-term effects of chronic methamphetamine administration in rhesus monkeys. Brain Res. 1989;486:73–78. doi: 10.1016/0006-8993(89)91279-1. [DOI] [PubMed] [Google Scholar]
  • 27.Seiden LS, et al. Neurotoxicity in dopamine and 5-hydroxytryptamine terminal fields: a regional analysis in nigrostriatal and mesolimbic projections. Ann N Y Acad Sci. 1988;537:161–172. doi: 10.1111/j.1749-6632.1988.tb42104.x. [DOI] [PubMed] [Google Scholar]
  • 28.Ricaurte G, et al. Hallucinogenic amphetamine selectively destroys brain serotonin nerve terminals. Science. 1985;229:986–988. doi: 10.1126/science.4023719. [DOI] [PubMed] [Google Scholar]
  • 29.Stone DM, et al. The effects of 3,4-methylenedioxymethamphetamine (MDMA) and 3,4-methylenedioxyamphetamine (MDA) on monoaminergic systems in the rat brain. Eur J Pharmacol. 1986;128:41–48. doi: 10.1016/0014-2999(86)90555-8. [DOI] [PubMed] [Google Scholar]
  • 30.Battaglia G, et al. In vitro and in vivo irreversible blockade of cortical S2 serotonin receptors by N-ethoxycarbonyl-2-ethoxy-1,2-dihydroquinoline: a technique for investigating S2 serotonin receptor recovery. J Neurochem. 1986;46:589–593. doi: 10.1111/j.1471-4159.1986.tb13008.x. [DOI] [PubMed] [Google Scholar]
  • 31.Finnegan KT, et al. Orally administered MDMA causes a long-term depletion of serotonin in rat brain. Brain Res. 1988;447:141–144. doi: 10.1016/0006-8993(88)90974-2. [DOI] [PubMed] [Google Scholar]
  • 32.O’Hearn E, et al. Methylenedioxyamphetamine (MDA) and methylenedioxymethamphetamine (MDMA) cause selective ablation of serotonergic axon terminals in forebrain: immunocytochemical evidence for neurotoxicity. J Neurosci. 1988;8:2788–2803. doi: 10.1523/JNEUROSCI.08-08-02788.1988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Wilson JM, et al. Striatal dopamine nerve terminal markers in human, chronic methamphetamine users. Nat Med. 1996;2:699–703. doi: 10.1038/nm0696-699. [DOI] [PubMed] [Google Scholar]
  • 34.Volkow ND, et al. Low level of brain dopamine D2 receptors in methamphetamine abusers: association with metabolism in the orbitofrontal cortex. Am J Psychiatry. 2001;158:2015–2021. doi: 10.1176/appi.ajp.158.12.2015. [DOI] [PubMed] [Google Scholar]
  • 35.Volkow ND, et al. Loss of dopamine transporters in methamphetamine abusers recovers with protracted abstinence. J Neurosci. 2001;21:9414–9418. doi: 10.1523/JNEUROSCI.21-23-09414.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Volkow ND, et al. Association of dopamine transporter reduction with psychomotor impairment in methamphetamine abusers. Am J Psychiatry. 2001;158:377–382. doi: 10.1176/appi.ajp.158.3.377. [DOI] [PubMed] [Google Scholar]
  • 37.Boileau I, et al. Increased vesicular monoamine transporter binding during early abstinence in human methamphetamine users: is VMAT2 a stable dopamine neuron biomarker? J Neurosci. 2008;28:9850–9856. doi: 10.1523/JNEUROSCI.3008-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Sekine Y, et al. Brain serotonin transporter density and aggression in abstinent methamphetamine abusers. Arch Gen Psychiatry. 2006;63:90–100. doi: 10.1001/archpsyc.63.1.90. [DOI] [PubMed] [Google Scholar]
  • 39.Kish SJ, et al. Brain serotonin transporter in human methamphetamine users. Psychopharmacology (Berl) 2008;202:649–661. doi: 10.1007/s00213-008-1346-x. [DOI] [PubMed] [Google Scholar]
  • 40.McCann UD, et al. Quantitative PET studies of the serotonin transporter in MDMA users and controls using [11C]McN5652 and [11C]DASB. Neuropsychopharmacology. 2005;30:1741–1750. doi: 10.1038/sj.npp.1300736. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Colado MI, O’Shea E, Green AR. Acute and long-term effects of MDMA on cerebral dopamine biochemistry and function. Psychopharmacology (Berl) 2004;173:249–263. doi: 10.1007/s00213-004-1788-8. [DOI] [PubMed] [Google Scholar]
  • 42.Itzhak Y, Achat-Mendes C. Methamphetamine and MDMA (ecstasy) neurotoxicity: ‘of mice and men’. IUBMB Life. 2004;56:249–255. doi: 10.1080/15216540410001727699. [DOI] [PubMed] [Google Scholar]
  • 43.Yamamoto BK, Bankson MG. Amphetamine neurotoxicity: cause and consequence of oxidative stress. Crit Rev Neurobiol. 2005;17:87–117. doi: 10.1615/critrevneurobiol.v17.i2.30. [DOI] [PubMed] [Google Scholar]
  • 44.Quinton MS, Yamamoto BK. Causes and consequences of methamphetamine and MDMA toxicity. AAPS J. 2006;8:E337–E347. doi: 10.1007/BF02854904. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Cadet JL, et al. Neurotoxicity of substituted amphetamines: molecular and cellular mechanisms. Neurotox Res. 2007;11:183–202. doi: 10.1007/BF03033567. [DOI] [PubMed] [Google Scholar]
  • 46.Fleckenstein AE, et al. New insights into the mechanism of action of amphetamines. Annu Rev Pharmacol Toxicol. 2007;47:681–698. doi: 10.1146/annurev.pharmtox.47.120505.105140. [DOI] [PubMed] [Google Scholar]
  • 47.Gudelsky GA, Yamamoto BK. Actions of 3,4-methylenedioxymethamphetamine (MDMA) on cerebral dopaminergic, serotonergic and cholinergic neurons. Pharmacol Biochem Behav. 2008;90:198–207. doi: 10.1016/j.pbb.2007.10.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Yamamoto BK, Raudensky J. The role of oxidative stress, metabolic compromise, and inflammation in neuronal injury produced by amphetamine-related drugs of abuse. J Neuroimmune Pharmacol. 2008;3:203–217. doi: 10.1007/s11481-008-9121-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Mirecki A, et al. Brain antioxidant systems in human methamphetamine users. J Neurochem. 2004;89:1396–1408. doi: 10.1111/j.1471-4159.2004.02434.x. [DOI] [PubMed] [Google Scholar]
  • 50.Fitzmaurice PS, et al. Levels of 4-hydroxynonenal and malondialdehyde are increased in brain of human chronic users of methamphetamine. J Pharmacol Exp Ther. 2006;319:703–709. doi: 10.1124/jpet.106.109173. [DOI] [PubMed] [Google Scholar]
  • 51.Wrona MZ, et al. Potential new insights into the molecular mechanisms of methamphetamine-induced neurodegeneration. NIDA Res Monogr. 1997;173:146–174. [PubMed] [Google Scholar]
  • 52.Hanson GR, Rau KS, Fleckenstein AE. The methamphetamine experience: a NIDA partnership. Neuropharmacology. 2004;47(Suppl. 1):92–100. doi: 10.1016/j.neuropharm.2004.06.004. [DOI] [PubMed] [Google Scholar]
  • 53.Sprague JE, Everman SL, Nichols DE. An integrated hypothesis for the serotonergic axonal loss induced by 3,4-methylenedioxymethamphetamine. Neurotoxicology. 1998;19:427–441. [PubMed] [Google Scholar]
  • 54.Breier JM, Bankson MG, Yamamoto BK. L-tyrosine contributes to (+)-3,4-methylenedioxymethamphetamine-induced serotonin depletions. J Neurosci. 2006;26:290–299. doi: 10.1523/JNEUROSCI.3353-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Jiang XR, et al. Reactions of the putative neurotoxin tryptamine-4,5-dione with L-cysteine and other thiols. Chem Res Toxicol. 2004;17:357–369. doi: 10.1021/tx020084k. [DOI] [PubMed] [Google Scholar]
  • 56.Jones DC, et al. Serotonergic neurotoxic metabolites of ecstasy identified in rat brain. J Pharmacol Exp Ther. 2005;313:422–431. doi: 10.1124/jpet.104.077628. [DOI] [PubMed] [Google Scholar]
  • 57.Erives GV, Lau SS, Monks TJ. Accumulation of neurotoxic thioether metabolites of 3,4-(+/−)-methylenedioxymethamphetamine in rat brain. J Pharmacol Exp Ther. 2008;324:284–291. doi: 10.1124/jpet.107.128785. [DOI] [PubMed] [Google Scholar]
  • 58.Bruno V, Scapagnini U, Canonico PL. Excitatory amino acids and neurotoxicity. Funct Neurol. 1993;8:279–292. [PubMed] [Google Scholar]
  • 59.Nicholls DG. Mitochondrial dysfunction and glutamate excitotoxicity studied in primary neuronal cultures. Curr Mol Med. 2004;4:149–177. doi: 10.2174/1566524043479239. [DOI] [PubMed] [Google Scholar]
  • 60.Uehara T. Accumulation of misfolded protein through nitrosative stress linked to neurodegenerative disorders. Antioxid Redox Signal. 2007;9:597–601. doi: 10.1089/ars.2006.1517. [DOI] [PubMed] [Google Scholar]
  • 61.Stephans SE, Yamamoto BK. Methamphetamine-induced neurotoxicity: roles for glutamate and dopamine efflux. Synapse. 1994;17:203–209. doi: 10.1002/syn.890170310. [DOI] [PubMed] [Google Scholar]
  • 62.Mark KA, Soghomonian JJ, Yamamoto BK. High-dose methamphetamine acutely activates the striatonigral pathway to increase striatal glutamate and mediate long-term dopamine toxicity. J Neurosci. 2004;24:11449–11456. doi: 10.1523/JNEUROSCI.3597-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Battaglia G, et al. Selective blockade of mGlu5 metabotropic glutamate receptors is protective against methamphetamine neurotoxicity. J Neurosci. 2002;22:2135–2141. doi: 10.1523/JNEUROSCI.22-06-02135.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Itzhak Y, Ali SF. The neuronal nitric oxide synthase inhibitor, 7-nitroindazole, protects against methamphetamine-induced neurotoxicity in vivo. J Neurochem. 1996;67:1770–1773. doi: 10.1046/j.1471-4159.1996.67041770.x. [DOI] [PubMed] [Google Scholar]
  • 65.Anderson KL, Itzhak Y. Methamphetamine-induced selective dopaminergic neurotoxicity is accompanied by an increase in striatal nitrate in the mouse. Ann N Y Acad Sci. 2006;1074:225–233. doi: 10.1196/annals.1369.021. [DOI] [PubMed] [Google Scholar]
  • 66.Wang J, et al. Connection between the striatal neurokinin-1 receptor and nitric oxide formation during methamphetamine exposure. Ann N Y Acad Sci. 2008;1139:164–171. doi: 10.1196/annals.1432.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Straiko MM, et al. The effect of amphetamine analogs on cleaved microtubule-associated protein-tau formation in the rat brain. Neuroscience. 2007;144:223–231. doi: 10.1016/j.neuroscience.2006.08.073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Staszewski RD, Yamamoto BK. Methamphetamine-induced spectrin proteolysis in the rat striatum. J Neurochem. 2006;96:1267–1276. doi: 10.1111/j.1471-4159.2005.03618.x. [DOI] [PubMed] [Google Scholar]
  • 69.Warren MW, et al. Calpain- and caspase-mediated alphaII-spectrin and tau proteolysis in rat cerebrocortical neuronal cultures after ecstasy or methamphetamine exposure. Int J Neuropsychopharmacol. 2007;10:479–489. doi: 10.1017/S1461145706007061. [DOI] [PubMed] [Google Scholar]
  • 70.Klongpanichapak S, et al. Attenuation of cocaine and methamphetamine neurotoxicity by coenzyme Q10. Neurochem Res. 2006;31:303–311. doi: 10.1007/s11064-005-9025-3. [DOI] [PubMed] [Google Scholar]
  • 71.Brown JM, Quinton MS, Yamamoto BK. Methamphetamine-induced inhibition of mitochondrial complex II: roles of glutamate and peroxynitrite. J Neurochem. 2005;95:429–436. doi: 10.1111/j.1471-4159.2005.03379.x. [DOI] [PubMed] [Google Scholar]
  • 72.Burrows KB, Gudelsky G, Yamamoto BK. Rapid and transient inhibition of mitochondrial function following methamphetamine or 3,4-methylenedioxymethamphetamine administration. Eur J Pharmacol. 2000;398:11–18. doi: 10.1016/s0014-2999(00)00264-8. [DOI] [PubMed] [Google Scholar]
  • 73.Alves E, et al. Monoamine oxidase-B mediates ecstasy-induced neurotoxic effects to adolescent rat brain mitochondria. J Neurosci. 2007;27:10203–10210. doi: 10.1523/JNEUROSCI.2645-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Burrows KB, Nixdorf WL, Yamamoto BK. Central administration of methamphetamine synergizes with metabolic inhibition to deplete striatal monoamines. J Pharmacol Exp Ther. 2000;292:853–860. [PubMed] [Google Scholar]
  • 75.Darvesh AS, Yamamoto BK, Gudelsky GA. Evidence for the involvement of nitric oxide in 3,4-methylenedioxymethamphetamine-induced serotonin depletion in the rat brain. J Pharmacol Exp Ther. 2005;312:694–701. doi: 10.1124/jpet.104.074849. [DOI] [PubMed] [Google Scholar]
  • 76.Stephans SE, et al. Substrates of energy metabolism attenuate methamphetamine-induced neurotoxicity in striatum. J Neurochem. 1998;71:613–621. doi: 10.1046/j.1471-4159.1998.71020613.x. [DOI] [PubMed] [Google Scholar]
  • 77.Bowyer JF, et al. Further studies of the role of hyperthermia in methamphetamine neurotoxicity. J Pharmacol Exp Ther. 1994;268:1571–1580. [PubMed] [Google Scholar]
  • 78.Ali SF, et al. Low environmental temperatures or pharmacologic agents that produce hypothermia decrease methamphetamine neurotoxicity in mice. Brain Res. 1994;658:33–38. doi: 10.1016/s0006-8993(09)90007-5. [DOI] [PubMed] [Google Scholar]
  • 79.Broening HW, Bowyer JF, Slikker W., Jr Age-dependent sensitivity of rats to the long-term effects of the serotonergic neurotoxicant (+/-)-3,4-methylenedioxymethamphetamine (MDMA) correlates with the magnitude of the MDMA-induced thermal response. J Pharmacol Exp Ther. 1995;275:325–333. [PubMed] [Google Scholar]
  • 80.Bowyer JF. The role of hyperthermia in amphetamine’s interactions with NMDA receptors, nitric oxide, and age to produce neurotoxicity. Ann N Y Acad Sci. 1995;765:309–310. doi: 10.1111/j.1749-6632.1995.tb16594.x. [DOI] [PubMed] [Google Scholar]
  • 81.Sonsalla PK, Riordan DE, Heikkila RE. Competitive and noncompetitive antagonists at N-methyl-D-aspartate receptors protect against methamphetamine-induced dopaminergic damage in mice. J Pharmacol Exp Ther. 1991;256:506–512. [PubMed] [Google Scholar]
  • 82.Albers DS, Sonsalla PK. Methamphetamine-induced hyperthermia and dopaminergic neurotoxicity in mice: pharmacological profile of protective and non-protective agents. J Pharmacol Exp Ther. 1995;275:1104–1114. [PubMed] [Google Scholar]
  • 83.Ali SF, Newport GD, Slikker W., Jr Methamphetamine-induced dopaminergic toxicity in mice. Role of environmental temperature and pharmacological agents. Ann N Y Acad Sci. 1996;801:187–198. doi: 10.1111/j.1749-6632.1996.tb17441.x. [DOI] [PubMed] [Google Scholar]
  • 84.Fleckenstein AE, et al. Interaction between hyperthermia and oxygen radical formation in the 5-hydroxytryptaminergic response to a single methamphetamine administration. J Pharmacol Exp Ther. 1997;283:281–285. [PubMed] [Google Scholar]
  • 85.Commins DL, Seiden LS. alpha-Methyltyrosine blocks methylamphetamine-induced degeneration in the rat somatosensory cortex. Brain Res. 1986;365:15–20. doi: 10.1016/0006-8993(86)90717-1. [DOI] [PubMed] [Google Scholar]
  • 86.Ellison G, Switzer RC., 3rd Dissimilar patterns of degeneration in brain following four different addictive stimulants. Neuroreport. 1993;5:17–20. doi: 10.1097/00001756-199310000-00004. [DOI] [PubMed] [Google Scholar]
  • 87.Sonsalla PK, et al. Treatment of mice with methamphetamine produces cell loss in the substantia nigra. Brain Res. 1996;738:172–175. doi: 10.1016/0006-8993(96)00995-x. [DOI] [PubMed] [Google Scholar]
  • 88.Schmued LC, Bowyer JF. Methamphetamine exposure can produce neuronal degeneration in mouse hippocampal remnants. Brain Res. 1997;759:135–140. doi: 10.1016/s0006-8993(97)00173-x. [DOI] [PubMed] [Google Scholar]
  • 89.Eisch AJ, Marshall JF. Methamphetamine neurotoxicity: dissociation of striatal dopamine terminal damage from parietal cortical cell body injury. Synapse. 1998;30:433–445. doi: 10.1002/(SICI)1098-2396(199812)30:4<433::AID-SYN10>3.0.CO;2-O. [DOI] [PubMed] [Google Scholar]
  • 90.Eisch AJ, Schmued LC, Marshall JF. Characterizing cortical neuron injury with Fluoro-Jade labeling after a neurotoxic regimen of methamphetamine. Synapse. 1998;30:329–333. doi: 10.1002/(SICI)1098-2396(199811)30:3<329::AID-SYN10>3.0.CO;2-V. [DOI] [PubMed] [Google Scholar]
  • 91.Scallet AC, et al. Neuropathological evaluation by combined immunohistochemistry and degeneration-specific methods: application to methylenedioxymethamphetamine. Neurotoxicology. 1988;9:529–537. [PubMed] [Google Scholar]
  • 92.O’Dell SJ, Marshall JF. Repeated administration of methamphetamine damages cells in the somatosensory cortex: overlap with cytochrome oxidase-rich barrels. Synapse. 2000;37:32–37. doi: 10.1002/(SICI)1098-2396(200007)37:1<32::AID-SYN4>3.0.CO;2-X. [DOI] [PubMed] [Google Scholar]
  • 93.Deng X, et al. Methamphetamine causes widespread apoptosis in the mouse brain: evidence from using an improved TUNEL histochemical method. Brain Res Mol Brain Res. 2001;93:64–69. doi: 10.1016/s0169-328x(01)00184-x. [DOI] [PubMed] [Google Scholar]
  • 94.Yu J, et al. Histological evidence supporting a role for the striatal neurokinin-1 receptor in methamphetamine-induced neurotoxicity in the mouse brain. Brain Res. 2004;1007:124–131. doi: 10.1016/j.brainres.2004.01.077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Kadota T, Kadota K. Neurotoxic morphological changes induced in the medial prefrontal cortex of rats behaviorally sensitized to methamphetamine. Arch Histol Cytol. 2004;67:241–251. doi: 10.1679/aohc.67.241. [DOI] [PubMed] [Google Scholar]
  • 96.Kuczenski R, et al. Escalating dose-multiple binge methamphetamine exposure results in degeneration of the neocortex and limbic system in the rat. Exp Neurol. 2007;207:42–51. doi: 10.1016/j.expneurol.2007.05.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Deng X, et al. Methamphetamine administration causes death of dopaminergic neurons in the mouse olfactory bulb. Biol Psychiatry. 2007;61:1235–1243. doi: 10.1016/j.biopsych.2006.09.010. [DOI] [PubMed] [Google Scholar]
  • 98.Atianjoh FE, et al. Amphetamine causes dopamine depletion and cell death in the mouse olfactory bulb. Eur J Pharmacol. 2008;589:94–97. doi: 10.1016/j.ejphar.2008.05.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Prudencio C, et al. Structural and functional cellular alterations underlying the toxicity of methamphetamine in rat retina and prefrontal cortex. Ann N Y Acad Sci. 2002;965:522–528. doi: 10.1111/j.1749-6632.2002.tb04193.x. [DOI] [PubMed] [Google Scholar]
  • 100.Gerfen CR. The neostriatal mosaic: multiple levels of compartmental organization in the basal ganglia. Annu Rev Neurosci. 1992;15:285–320. doi: 10.1146/annurev.ne.15.030192.001441. [DOI] [PubMed] [Google Scholar]
  • 101.Thiriet N, et al. Neuropeptide Y protects against methamphetamine-induced neuronal apoptosis in the mouse striatum. J Neurosci. 2005;25:5273–5279. doi: 10.1523/JNEUROSCI.4893-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Zhu JP, Xu W, Angulo JA. Methamphetamine-induced cell death: selective vulnerability in neuronal subpopulations of the striatum in mice. Neuroscience. 2006;140:607–622. doi: 10.1016/j.neuroscience.2006.02.055. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Jayanthi S, et al. Calcineurin/NFAT-induced upregulation of the Fas ligand/Fas death pathway is involved in methamphetamine-induced neuronal apoptosis. Proc Natl Acad Sci USA. 2005;102:868–873. doi: 10.1073/pnas.0404990102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Gotz T, et al. Functional properties of AMPA and NMDA receptors expressed in identified types of basal ganglia neurons. J Neurosci. 1997;17:204–215. doi: 10.1523/JNEUROSCI.17-01-00204.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Kwok KH, et al. Cellular localization of GluR1, GluR2/3 and GluR4 glutamate receptor subunits in neurons of the rat neostriatum. Brain Res. 1997;778:43–55. doi: 10.1016/s0006-8993(97)00950-5. [DOI] [PubMed] [Google Scholar]
  • 106.Jayanthi S, et al. Methamphetamine induces neuronal apoptosis via cross-talks between endoplasmic reticulum and mitochondria-dependent death cascades. FASEB J. 2004;18:238–251. doi: 10.1096/fj.03-0295com. [DOI] [PubMed] [Google Scholar]
  • 107.Cregan SP, et al. Apoptosis-inducing factor is involved in the regulation of caspase-independent neuronal cell death. J Cell Biol. 2002;158:507–517. doi: 10.1083/jcb.200202130. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Deng X, et al. Methamphetamine induces apoptosis in an immortalized rat striatal cell line by activating the mitochondrial cell death pathway. Neuropharmacology. 2002;42:837–845. doi: 10.1016/s0028-3908(02)00034-5. [DOI] [PubMed] [Google Scholar]
  • 109.Yu SW, et al. Mediation of poly(ADP-ribose) polymerase-1-dependent cell death by apoptosis-inducing factor. Science. 2002;297:259–263. doi: 10.1126/science.1072221. [DOI] [PubMed] [Google Scholar]
  • 110.Zhang Y, Bhavnani BR. Glutamate-induced apoptosis in neuronal cells is mediated via caspase-dependent and independent mechanisms involving calpain and caspase-3 proteases as well as apoptosis inducing factor (AIF) and this process is inhibited by equine estrogens. BMC Neurosci. 2006;7:49. doi: 10.1186/1471-2202-7-49. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Xu W, Zhu JP, Angulo JA. Induction of striatal pre- and postsynaptic damage by methamphetamine requires the dopamine receptors. Synapse. 2005;58:110–121. doi: 10.1002/syn.20185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Cepeda C, Buchwald NA, Levine MS. Neuromodulatory actions of dopamine in the neostriatum are dependent upon the excitatory amino acid receptor subtypes activated. Proc Natl Acad Sci USA. 1993;90:9576–9580. doi: 10.1073/pnas.90.20.9576. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Lee FJ, et al. Dual regulation of NMDA receptor functions by direct protein-protein interactions with the dopamine D1 receptor. Cell. 2002;111:219–230. doi: 10.1016/s0092-8674(02)00962-5. [DOI] [PubMed] [Google Scholar]
  • 114.Pei L, et al. Regulation of dopamine D1 receptor function by physical interaction with the NMDA receptors. J Neurosci. 2004;24:1149–1158. doi: 10.1523/JNEUROSCI.3922-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Angulo JA, McEwen BS. Molecular aspects of neuropeptide regulation and function in the corpus striatum and nucleus accumbens. Brain Res Brain Res Rev. 1994;19:1–28. doi: 10.1016/0165-0173(94)90002-7. [DOI] [PubMed] [Google Scholar]
  • 116.Camp DM, Loeffler DA, LeWitt PA. L-DOPA does not enhance hydroxyl radical formation in the nigrostriatal dopamine system of rats with a unilateral 6-hydroxydopamine lesion. J Neurochem. 2000;74:1229–1240. doi: 10.1046/j.1471-4159.2000.741229.x. [DOI] [PubMed] [Google Scholar]
  • 117.Lazzeri G, et al. Mechanisms involved in the formation of dopamine-induced intracellular bodies within striatal neurons. J Neurochem. 2007;101:1414–1427. doi: 10.1111/j.1471-4159.2006.04429.x. [DOI] [PubMed] [Google Scholar]
  • 118.Kawaguchi Y, et al. Striatal interneurones: chemical, physiological and morphological characterization. Trends Neurosci. 1995;18:527–535. doi: 10.1016/0166-2236(95)98374-8. [DOI] [PubMed] [Google Scholar]
  • 119.Yu J, Cadet JL, Angulo JA. Neurokinin-1 (NK-1) receptor antagonists abrogate methamphetamine-induced striatal dopaminergic neurotoxicity in the murine brain. J Neurochem. 2002;83:613–622. doi: 10.1046/j.1471-4159.2002.01155.x. [DOI] [PubMed] [Google Scholar]
  • 120.Zhu JP, Xu W, Angulo JA. Distinct mechanisms mediating methamphetamine-induced neuronal apoptosis and dopamine terminal damage share the neuropeptide substance p in the striatum of mice. Ann N Y Acad Sci. 2006;1074:135–148. doi: 10.1196/annals.1369.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Jakab RL, Goldman-Rakic P. Presynaptic and postsynaptic subcellular localization of substance P receptor immunoreactivity in the neostriatum of the rat and rhesus monkey (Macaca mulatta) J Comp Neurol. 1996;369:125–136. doi: 10.1002/(SICI)1096-9861(19960520)369:1<125::AID-CNE9>3.0.CO;2-5. [DOI] [PubMed] [Google Scholar]
  • 122.Eisch AJ, O’Dell SJ, Marshall JF. Striatal and cortical NMDA receptors are altered by a neurotoxic regimen of methamphetamine. Synapse. 1996;22:217–225. doi: 10.1002/(SICI)1098-2396(199603)22:3<217::AID-SYN3>3.0.CO;2-F. [DOI] [PubMed] [Google Scholar]
  • 123.O’Dell SJ, Marshall JF. Effects of vibrissae removal on methamphetamine-induced damage to rat somatosensory cortical neurons. Synapse. 2002;43:122–130. doi: 10.1002/syn.10016. [DOI] [PubMed] [Google Scholar]
  • 124.Neumann H. Control of glial immune function by neurons. Glia. 2001;36:191–199. doi: 10.1002/glia.1108. [DOI] [PubMed] [Google Scholar]
  • 125.Thomas DM, Kuhn DM. MK-801 and dextromethorphan block microglial activation and protect against methamphetamine-induced neurotoxicity. Brain Res. 2005;1050:190–198. doi: 10.1016/j.brainres.2005.05.049. [DOI] [PubMed] [Google Scholar]
  • 126.de Bock F, Dornand J, Rondouin G. Release of TNF alpha in the rat hippocampus following epileptic seizures and excitotoxic neuronal damage. Neuroreport. 1996;7:1125–1129. doi: 10.1097/00001756-199604260-00004. [DOI] [PubMed] [Google Scholar]
  • 127.Vezzani A, et al. Interleukin-1beta immunoreactivity and microglia are enhanced in the rat hippocampus by focal kainate application: functional evidence for enhancement of electrographic seizures. J Neurosci. 1999;19:5054–5065. doi: 10.1523/JNEUROSCI.19-12-05054.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Marini H, et al. Modulation of IL-1 beta gene expression by lipid peroxidation inhibition after kainic acid-induced rat brain injury. Exp Neurol. 2004;188:178–186. doi: 10.1016/j.expneurol.2004.03.023. [DOI] [PubMed] [Google Scholar]
  • 129.Chaparro-Huerta V, et al. Proinflammatory cytokines and apoptosis following glutamate-induced excitotoxicity mediated by p38 MAPK in the hippocampus of neonatal rats. J Neuroimmunol. 2005;165:53–62. doi: 10.1016/j.jneuroim.2005.04.025. [DOI] [PubMed] [Google Scholar]
  • 130.Zou JY, Crews FT. TNF alpha potentiates glutamate neurotoxicity by inhibiting glutamate uptake in organotypic brain slice cultures: neuroprotection by NF kappa B inhibition. Brain Res. 2005;1034:11–24. doi: 10.1016/j.brainres.2004.11.014. [DOI] [PubMed] [Google Scholar]
  • 131.Casamenti F, et al. Interleukin-1beta activates forebrain glial cells and increases nitric oxide production and cortical glutamate and GABA release in vivo: implications for Alzheimer’s disease. Neuroscience. 1999;91:831–842. doi: 10.1016/s0306-4522(98)00680-0. [DOI] [PubMed] [Google Scholar]
  • 132.Guilarte TR, et al. Methamphetamine-induced deficits of brain monoaminergic neuronal markers: distal axotomy or neuronal plasticity. Neuroscience. 2003;122:499–513. doi: 10.1016/s0306-4522(03)00476-7. [DOI] [PubMed] [Google Scholar]
  • 133.Thomas DM, et al. Microglial activation is a pharmacologically specific marker for the neurotoxic amphetamines. Neurosci Lett. 2004;367:349–354. doi: 10.1016/j.neulet.2004.06.065. [DOI] [PubMed] [Google Scholar]
  • 134.Thomas DM, et al. Methamphetamine neurotoxicity in dopamine nerve endings of the striatum is associated with microglial activation. J Pharmacol Exp Ther. 2004;311:1–7. doi: 10.1124/jpet.104.070961. [DOI] [PubMed] [Google Scholar]
  • 135.Bowyer JF, et al. Neurotoxic-related changes in tyrosine hydroxylase, microglia, myelin, and the blood–brain barrier in the caudate-putamen from acute methamphetamine exposure. Synapse. 2008;62:193–204. doi: 10.1002/syn.20478. [DOI] [PubMed] [Google Scholar]
  • 136.Fantegrossi WE, et al. A comparison of the physiological, behavioral, neurochemical and microglial effects of methamphetamine and 3,4-methylenedioxymethamphetamine in the mouse. Neuroscience. 2008;151:533–543. doi: 10.1016/j.neuroscience.2007.11.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Pubill D, et al. Different glial response to methamphetamine- and methylenedioxymethamphetamine-induced neurotoxicity. Naunyn Schmiedebergs Arch Pharmacol. 2003;367:490–499. doi: 10.1007/s00210-003-0747-y. [DOI] [PubMed] [Google Scholar]
  • 138.LaVoie MJ, Card JP, Hastings TG. Microglial activation precedes dopamine terminal pathology in methamphetamine-induced neurotoxicity. Exp Neurol. 2004;187:47–57. doi: 10.1016/j.expneurol.2004.01.010. [DOI] [PubMed] [Google Scholar]
  • 139.Escubedo E, et al. Microgliosis and down-regulation of adenosine transporter induced by methamphetamine in rats. Brain Res. 1998;814:120–126. doi: 10.1016/s0006-8993(98)01065-8. [DOI] [PubMed] [Google Scholar]
  • 140.Thomas DM, Francescutti-Verbeem DM, Kuhn DM. The newly synthesized pool of dopamine determines the severity of methamphetamine-induced neurotoxicity. J Neurochem. 2008;105:605–616. doi: 10.1111/j.1471-4159.2007.05155.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Sekine Y, et al. Methamphetamine causes microglial activation in the brains of human abusers. J Neurosci. 2008;28:5756–5761. doi: 10.1523/JNEUROSCI.1179-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Coutinho A, et al. Chronic methamphetamine induces structural changes in frontal cortex neurons and upregulates type I interferons. J Neuroimmune Pharmacol. 2008;3:241–245. doi: 10.1007/s11481-008-9113-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Sriram K, Miller DB, O’Callaghan JP. Minocycline attenuates microglial activation but fails to mitigate striatal dopaminergic neurotoxicity: role of tumor necrosis factor-alpha. J Neurochem. 2006;96:706–718. doi: 10.1111/j.1471-4159.2005.03566.x. [DOI] [PubMed] [Google Scholar]
  • 144.Goncalves J, et al. Methamphetamine-induced early increase of IL-6 and TNF-alpha mRNA expression in the mouse brain. Ann N Y Acad Sci. 2008;1139:103–111. doi: 10.1196/annals.1432.043. [DOI] [PubMed] [Google Scholar]
  • 145.Facchinetti F, et al. Dopamine inhibits responses of astroglia-enriched cultures to lipopolysaccharide via a beta-adrenoreceptor-mediated mechanism. J Neuroimmunol. 2004;150:29–36. doi: 10.1016/j.jneuroim.2004.01.014. [DOI] [PubMed] [Google Scholar]
  • 146.Kuhn DM, Francescutti-Verbeem DM, Thomas DM. Dopamine quinones activate microglia and induce a neurotoxic gene expression profile: relationship to methamphetamine-induced nerve ending damage. Ann N Y Acad Sci. 2006;1074:31–41. doi: 10.1196/annals.1369.003. [DOI] [PubMed] [Google Scholar]
  • 147.LaVoie MJ, Hastings TG. Dopamine quinone formation and protein modification associated with the striatal neurotoxicity of methamphetamine: evidence against a role for extracellular dopamine. J Neurosci. 1999;19:1484–1491. doi: 10.1523/JNEUROSCI.19-04-01484.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Orio L, et al. 3,4-Methylenedioxymethamphetamine increases interleukin-1beta levels and activates microglia in rat brain: studies on the relationship with acute hyperthermia and 5-HT depletion. J Neurochem. 2004;89:1445–1453. doi: 10.1111/j.1471-4159.2004.02443.x. [DOI] [PubMed] [Google Scholar]
  • 149.O’Shea E, et al. 3,4-Methylenedioxymethamphetamine increases pro-interleukin-1beta production and caspase-1 protease activity in frontal cortex, but not in hypothalamus, of Dark Agouti rats: role of interleukin-1beta in neurotoxicity. Neuroscience. 2005;135:1095–1105. doi: 10.1016/j.neuroscience.2005.06.084. [DOI] [PubMed] [Google Scholar]
  • 150.Boyle NT, Connor TJ. MDMA (“Ecstasy”) suppresses the innate IFN-gamma response in vivo: a critical role for the anti-inflammatory cytokine IL-10. Eur J Pharmacol. 2007;572:228–238. doi: 10.1016/j.ejphar.2007.07.020. [DOI] [PubMed] [Google Scholar]
  • 151.Hozumi H, et al. Protective effects of interferon-gamma against methamphetamine-induced neurotoxicity. Toxicol Lett. 2008;177:123–129. doi: 10.1016/j.toxlet.2008.01.005. [DOI] [PubMed] [Google Scholar]
  • 152.Lin YC, et al. Attenuation of methamphetamine-induced nigrostriatal dopaminergic neurotoxicity in mice by lipopolysaccharide pretreatment. Chin J Physiol. 2007;50:51–56. [PubMed] [Google Scholar]
  • 153.Hu S, et al. Cytokine effects on glutamate uptake by human astrocytes. Neuroimmunomodulation. 2000;7:153–159. doi: 10.1159/000026433. [DOI] [PubMed] [Google Scholar]
  • 154.Shaked I, et al. Protective autoimmunity: interferon-gamma enables microglia to remove glutamate without evoking inflammatory mediators. J Neurochem. 2005;92:997–1009. doi: 10.1111/j.1471-4159.2004.02954.x. [DOI] [PubMed] [Google Scholar]
  • 155.Nakajima A, et al. Role of tumor necrosis factor-alpha in methamphetamine-induced drug dependence and neurotoxicity. J Neurosci. 2004;24:2212–2225. doi: 10.1523/JNEUROSCI.4847-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Parpura V, Haydon PG. Physiological astrocytic calcium levels stimulate glutamate release to modulate adjacent neurons. Proc Natl Acad Sci USA. 2000;97:8629–8634. doi: 10.1073/pnas.97.15.8629. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Miyatake M, et al. Glutamatergic neurotransmission and protein kinase C play a role in neuron-glia communication during the development of methamphetamine-induced psychological dependence. Eur J Neurosci. 2005;22:1476–1488. doi: 10.1111/j.1460-9568.2005.04325.x. [DOI] [PubMed] [Google Scholar]
  • 158.Vincent VA, Tilders FJ, Van Dam AM. Inhibition of endotoxin-induced nitric oxide synthase production in microglial cells by the presence of astroglial cells: a role for transforming growth factor beta. Glia. 1997;19:190–198. doi: 10.1002/(sici)1098-1136(199703)19:3<190::aid-glia2>3.0.co;2-3. [DOI] [PubMed] [Google Scholar]
  • 159.Aloisi F. Immune function of microglia. Glia. 2001;36:165–179. doi: 10.1002/glia.1106. [DOI] [PubMed] [Google Scholar]
  • 160.Oh JW, Schwiebert LM, Benveniste EN. Cytokine regulation of CC and CXC chemokine expression by human astrocytes. J Neurovirol. 1999;5:82–94. doi: 10.3109/13550289909029749. [DOI] [PubMed] [Google Scholar]
  • 161.Block ML, Hong JS. Chronic microglial activation and progressive dopaminergic neurotoxicity. Biochem Soc Trans. 2007;35:1127–1132. doi: 10.1042/BST0351127. [DOI] [PubMed] [Google Scholar]
  • 162.Kawasaki T, et al. Protective effect of the radical scavenger edaravone against methamphetamine-induced dopaminergic neurotoxicity in mouse striatum. Eur J Pharmacol. 2006;542:92–99. doi: 10.1016/j.ejphar.2006.05.012. [DOI] [PubMed] [Google Scholar]
  • 163.Mehlhase J, et al. Oxidation-induced ferritin turnover in microglial cells: role of proteasome. Free Radic Biol Med. 2005;38:276–285. doi: 10.1016/j.freeradbiomed.2004.10.025. [DOI] [PubMed] [Google Scholar]
  • 164.Yamamoto BK, Zhu W. The effects of methamphetamine on the production of free radicals and oxidative stress. J Pharmacol Exp Ther. 1998;287:107–114. [PubMed] [Google Scholar]
  • 165.Olzmann JA, Li L, Chin LS. Aggresome formation and neurodegenerative diseases: therapeutic implications. Curr Med Chem. 2008;15:47–60. doi: 10.2174/092986708783330692. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Golde S, et al. Decreased iNOS synthesis mediates dexamethasone-induced protection of neurons from inflammatory injury in vitro. Eur J Neurosci. 2003;18:2527–2537. doi: 10.1046/j.1460-9568.2003.02917.x. [DOI] [PubMed] [Google Scholar]
  • 167.Chung KK, et al. S-nitrosylation of parkin regulates ubiquitination and compromises parkin’s protective function. Science. 2004;304:1328–1331. doi: 10.1126/science.1093891. [DOI] [PubMed] [Google Scholar]
  • 168.Nakamura T, Lipton SA. Emerging roles of S-nitrosylation in protein misfolding and neurodegenerative diseases. Antioxid Redox Signal. 2008;10:87–101. doi: 10.1089/ars.2007.1858. [DOI] [PubMed] [Google Scholar]
  • 169.Osna NA, et al. Peroxynitrite alters the catalytic activity of rodent liver proteasome in vitro and in vivo. Hepatology. 2004;40:574–582. doi: 10.1002/hep.20352. [DOI] [PubMed] [Google Scholar]
  • 170.Yang H, et al. Downregulation of parkin damages antioxidant defenses and enhances proteasome inhibition-induced toxicity in PC12 cells. J Neuroimmune Pharmacol. 2007;2:276–283. doi: 10.1007/s11481-007-9082-2. [DOI] [PubMed] [Google Scholar]
  • 171.Papa L, Rockwell P. Persistent mitochondrial dysfunction and oxidative stress hinder neuronal cell recovery from reversible proteasome inhibition. Apoptosis. 2008;13:588–599. doi: 10.1007/s10495-008-0182-0. [DOI] [PubMed] [Google Scholar]
  • 172.Fornai F, et al. Striatal postsynaptic ultrastructural alterations following methylenedioxymethamphetamine administration. Ann N Y Acad Sci. 2002;965:381–398. doi: 10.1111/j.1749-6632.2002.tb04180.x. [DOI] [PubMed] [Google Scholar]
  • 173.Fornai F, et al. Amphetamines induce ubiquitin-positive inclusions within striatal cells. Neurol Sci. 2003;24:182–183. doi: 10.1007/s10072-003-0120-4. [DOI] [PubMed] [Google Scholar]
  • 174.Fornai F, et al. DNA damage and ubiquitinated neuronal inclusions in the substantia nigra and striatum of mice following MDMA (ecstasy) Psychopharmacology (Berl) 2004;173:353–363. doi: 10.1007/s00213-003-1708-3. [DOI] [PubMed] [Google Scholar]
  • 175.Fornai F, et al. Methamphetamine produces neuronal inclusions in the nigrostriatal system and in PC12 cells. J Neurochem. 2004;88:114–123. doi: 10.1046/j.1471-4159.2003.02137.x. [DOI] [PubMed] [Google Scholar]
  • 176.Fornai F, et al. Occurrence of neuronal inclusions combined with increased nigral expression of alpha-synuclein within dopaminergic neurons following treatment with amphetamine derivatives in mice. Brain Res Bull. 2005;65:405–413. doi: 10.1016/j.brainresbull.2005.02.022. [DOI] [PubMed] [Google Scholar]
  • 177.Quan L, et al. Ubiquitin-immunoreactive structures in the midbrain of methamphetamine abusers. Leg Med (Tokyo) 2005;7:144–150. doi: 10.1016/j.legalmed.2004.11.002. [DOI] [PubMed] [Google Scholar]
  • 178.Fornai F, et al. Fine structure and biochemical mechanisms underlying nigrostriatal inclusions and cell death after proteasome inhibition. J Neurosci. 2003;23:8955–8966. doi: 10.1523/JNEUROSCI.23-26-08955.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.McNaught KS, et al. Systemic exposure to proteasome inhibitors causes a progressive model of Parkinson’s disease. Ann Neurol. 2004;56:149–162. doi: 10.1002/ana.20186. [DOI] [PubMed] [Google Scholar]
  • 180.Gesi M, et al. Inclusion dynamics in PC12 is comparable between amphetamines and MPTP. Ann N Y Acad Sci. 2006;1074:315–319. doi: 10.1196/annals.1369.028. [DOI] [PubMed] [Google Scholar]
  • 181.Iacovelli L, et al. The neurotoxicity of amphetamines: bridging drugs of abuse and neurodegenerative disorders. Exp Neurol. 2006;201:24–31. doi: 10.1016/j.expneurol.2006.02.130. [DOI] [PubMed] [Google Scholar]
  • 182.De Blasi A, et al. Presence of beta-arrestin in cellular inclusions in metamphetamine-treated PC12 cells. Neurol Sci. 2003;24:164–165. doi: 10.1007/s10072-003-0111-5. [DOI] [PubMed] [Google Scholar]
  • 183.Iacovelli L, et al. Role of G protein-coupled receptor kinase 4 and beta-arrestin 1 in agonist-stimulated metabotropic glutamate receptor 1 internalization and activation of mitogen-activated protein kinases. J Biol Chem. 2003;278:12433–12442. doi: 10.1074/jbc.M203992200. [DOI] [PubMed] [Google Scholar]
  • 184.Macey TA, Gurevich VV, Neve KA. Preferential interaction between the dopamine D2 receptor and arrestin2 in neostriatal neurons. Mol Pharmacol. 2004;66:1635–1642. doi: 10.1124/mol.104.001495. [DOI] [PubMed] [Google Scholar]
  • 185.Fiorentini C, et al. Reciprocal regulation of dopamine D1 and D3 receptor function and trafficking by heterodimerization. Mol Pharmacol. 2008;74:59–69. doi: 10.1124/mol.107.043885. [DOI] [PubMed] [Google Scholar]
  • 186.Oluwatosin-Chigbu Y, et al. Parkin suppresses wild-type alpha-synuclein-induced toxicity in SHSY-5Y cells. Biochem Biophys Res Commun. 2003;309:679–684. doi: 10.1016/j.bbrc.2003.08.059. [DOI] [PubMed] [Google Scholar]
  • 187.Lo Bianco C, et al. alpha-Synucleinopathy and selective dopaminergic neuron loss in a rat lentiviral-based model of Parkinson’s disease. Proc Natl Acad Sci USA. 2002;99:10813–10818. doi: 10.1073/pnas.152339799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Conway KA, et al. Kinetic stabilization of the alpha-synuclein protofibril by a dopamine-alpha-synuclein adduct. Science. 2001;294:1346–1349. doi: 10.1126/science.1063522. [DOI] [PubMed] [Google Scholar]
  • 189.Li HT, et al. Inhibition of alpha-synuclein fibrillization by dopamine analogs via reaction with the amino groups of alpha-synuclein. Implication for dopaminergic neurodegeneration. Febs J. 2005;272:3661–3672. doi: 10.1111/j.1742-4658.2005.04792.x. [DOI] [PubMed] [Google Scholar]
  • 190.Rochet JC, et al. Interactions among alpha-synuclein, dopamine, and biomembranes: some clues for understanding neurodegeneration in Parkinson’s disease. J Mol Neurosci. 2004;23:23–34. doi: 10.1385/jmn:23:1-2:023. [DOI] [PubMed] [Google Scholar]
  • 191.Castino R, et al. Suppression of autophagy precipitates neuronal cell death following low doses of methamphetamine. J Neurochem. 2008;106:1426–1439. doi: 10.1111/j.1471-4159.2008.05488.x. [DOI] [PubMed] [Google Scholar]
  • 192.Connor B, Dragunow M. The role of neuronal growth factors in neurodegenerative disorders of the human brain. Brain Res Brain Res Rev. 1998;27:1–39. doi: 10.1016/s0165-0173(98)00004-6. [DOI] [PubMed] [Google Scholar]
  • 193.Reichardt LF. Neurotrophin-regulated signalling pathways. Philos Trans R Soc Lond B Biol Sci. 2006;361:1545–1564. doi: 10.1098/rstb.2006.1894. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Airaksinen MS, Saarma M. The GDNF family: signalling, biological functions and therapeutic value. Nat Rev Neurosci. 2002;3:383–394. doi: 10.1038/nrn812. [DOI] [PubMed] [Google Scholar]
  • 195.Cass WA. GDNF selectively protects dopamine neurons over serotonin neurons against the neurotoxic effects of methamphetamine. J Neurosci. 1996;16:8132–8139. doi: 10.1523/JNEUROSCI.16-24-08132.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Cass WA, Manning MW, Bailey SL. Restorative effects of GDNF on striatal dopamine release in rats treated with neurotoxic doses of methamphetamine. Ann N Y Acad Sci. 2000;914:127–136. doi: 10.1111/j.1749-6632.2000.tb05190.x. [DOI] [PubMed] [Google Scholar]
  • 197.Melega WP, et al. Long-term methamphetamine-induced decreases of [(11)C]WIN 35,428 binding in striatum are reduced by GDNF: PET studies in the vervet monkey. Synapse. 2000;35:243–249. doi: 10.1002/(SICI)1098-2396(20000315)35:4<243::AID-SYN1>3.0.CO;2-N. [DOI] [PubMed] [Google Scholar]
  • 198.Cass WA, et al. Protection by GDNF and other trophic factors against the dopamine-depleting effects of neurotoxic doses of methamphetamine. Ann N Y Acad Sci. 2006;1074:272–281. doi: 10.1196/annals.1369.024. [DOI] [PubMed] [Google Scholar]
  • 199.Chao CC, Lee EH. Neuroprotective mechanism of glial cell line-derived neurotrophic factor on dopamine neurons: role of antioxidation. Neuropharmacology. 1999;38:913–916. doi: 10.1016/s0028-3908(99)00030-1. [DOI] [PubMed] [Google Scholar]
  • 200.Sawada H, et al. Neuroprotective mechanism of glial cell line-derived neurotrophic factor in mesencephalic neurons. J Neurochem. 2000;74:1175–1184. doi: 10.1046/j.1471-4159.2000.741175.x. [DOI] [PubMed] [Google Scholar]
  • 201.Boger HA, et al. Long-term consequences of methamphetamine exposure in young adults are exacerbated in glial cell line-derived neurotrophic factor heterozygous mice. J Neurosci. 2007;27:8816–8825. doi: 10.1523/JNEUROSCI.1067-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Matsuzaki H, et al. Brain-derived neurotrophic factor rescues neuronal death induced by methamphetamine. Biol Psychiatry. 2004;55:52–60. doi: 10.1016/s0006-3223(03)00785-6. [DOI] [PubMed] [Google Scholar]
  • 203.Bowyer JF, Ali S. High doses of methamphetamine that cause disruption of the blood-brain barrier in limbic regions produce extensive neuronal degeneration in mouse hippocampus. Synapse. 2006;60:521–532. doi: 10.1002/syn.20324. [DOI] [PubMed] [Google Scholar]
  • 204.Kiyatkin EA, Brown PL, Sharma HS. Brain edema and breakdown of the blood-brain barrier during methamphetamine intoxication: critical role of brain hyperthermia. Eur J Neurosci. 2007;26:1242–1253. doi: 10.1111/j.1460-9568.2007.05741.x. [DOI] [PubMed] [Google Scholar]
  • 205.Sharma HS, Sjoquist PO, Ali SF. Drugs of abuse-induced hyperthermia, blood-brain barrier dysfunction and neurotoxicity: neuroprotective effects of a new antioxidant compound H-290/51. Curr Pharm Des. 2007;13:1903–1923. doi: 10.2174/138161207780858375. [DOI] [PubMed] [Google Scholar]
  • 206.Sharma HS, Kiyatkin EA. Rapid morphological brain abnormalities during acute methamphetamine intoxication in the rat: an experimental study using light and electron microscopy. J Chem Neuroanat. 2009;37:18–32. doi: 10.1016/j.jchemneu.2008.08.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Stanimirovic DB, et al. Free radical-induced endothelial membrane dysfunction at the site of blood-brain barrier: relationship between lipid peroxidation, Na,K-ATPase activity, and 51Cr release. Neurochem Res. 1995;20:1417–1427. doi: 10.1007/BF00970589. [DOI] [PubMed] [Google Scholar]
  • 208.Yang Y, et al. Matrix metalloproteinase-mediated disruption of tight junction proteins in cerebral vessels is reversed by synthetic matrix metalloproteinase inhibitor in focal ischemia in rat. J Cereb Blood Flow Metab. 2007;27:697–709. doi: 10.1038/sj.jcbfm.9600375. [DOI] [PubMed] [Google Scholar]
  • 209.Persidsky Y, et al. Blood-brain barrier: structural components and function under physiologic and pathologic conditions. J Neuroimmune Pharmacol. 2006;1:223–236. doi: 10.1007/s11481-006-9025-3. [DOI] [PubMed] [Google Scholar]
  • 210.Conant K, et al. Human immunodeficiency virus type 1 Tat and methamphetamine affect the release and activation of matrix-degrading proteinases. J Neurovirol. 2004;10:21–28. doi: 10.1080/13550280490261699. [DOI] [PubMed] [Google Scholar]
  • 211.Mahajan SD, et al. Methamphetamine alters blood brain barrier permeability via the modulation of tight junction expression: Implication for HIV-1 neuropathogenesis in the context of drug abuse. Brain Res. 2008;1203:133–148. doi: 10.1016/j.brainres.2008.01.093. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Fazo N, Raudensky BK, Yamamoto BK. Matrix metalloproteinase-9 in the hippocampus after chronic unpredictable stress and methamphetamine exposure. Society for Neuroscience Annual Meeting; Washington, DC. November 15–19.2008. [Google Scholar]
  • 213.Haorah J, et al. Oxidative stress activates protein tyrosine kinase and matrix metalloproteinases leading to blood-brain barrier dysfunction. J Neurochem. 2007;101:566–576. doi: 10.1111/j.1471-4159.2006.04393.x. [DOI] [PubMed] [Google Scholar]
  • 214.Harkness KA, et al. Dexamethasone regulation of matrix metalloproteinase expression in CNS vascular endothelium. Brain. 2000;123(Pt 4):698–709. doi: 10.1093/brain/123.4.698. [DOI] [PubMed] [Google Scholar]
  • 215.McColl BW, Rothwell NJ, Allan SM. Systemic inflammation alters the kinetics of cerebrovascular tight junction disruption after experimental stroke in mice. J Neurosci. 2008;28:9451–9462. doi: 10.1523/JNEUROSCI.2674-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Amantea D, et al. Early upregulation of matrix metalloproteinases following reperfusion triggers neuroinflammatory mediators in brain ischemia in rat. Int Rev Neurobiol. 2007;82:149–169. doi: 10.1016/S0074-7742(07)82008-3. [DOI] [PubMed] [Google Scholar]
  • 217.Kim YS, et al. Matrix metalloproteinase-3: a novel signaling proteinase from apoptotic neuronal cells that activates microglia. J Neurosci. 2005;25:3701–3711. doi: 10.1523/JNEUROSCI.4346-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Reyes MG, et al. Nigral degeneration in acquired immune deficiency syndrome (AIDS) Acta Neuropathol. 1991;82:39–44. doi: 10.1007/BF00310921. [DOI] [PubMed] [Google Scholar]
  • 219.Sardar AM, Czudek C, Reynolds GP. Dopamine deficits in the brain: the neurochemical basis of parkinsonian symptoms in AIDS. Neuroreport. 1996;7:910–912. doi: 10.1097/00001756-199603220-00015. [DOI] [PubMed] [Google Scholar]
  • 220.Ferris MJ, Mactutus CF, Booze RM. Neurotoxic profiles of HIV, psychostimulant drugs of abuse, and their concerted effect on the brain: current status of dopamine system vulnerability in NeuroAIDS. Neurosci Biobehav Rev. 2008;32:883–909. doi: 10.1016/j.neubiorev.2008.01.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 221.Hayman M, et al. Neurotoxicity of peptide analogues of the transactivating protein tat from Maedi-Visna virus and human immunodeficiency virus. Neuroscience. 1993;53:1–6. doi: 10.1016/0306-4522(93)90278-n. [DOI] [PubMed] [Google Scholar]
  • 222.Bansal AK, et al. Neurotoxicity of HIV-1 proteins gp120 and Tat in the rat striatum. Brain Res. 2000;879:42–49. doi: 10.1016/s0006-8993(00)02725-6. [DOI] [PubMed] [Google Scholar]
  • 223.Zauli G, et al. HIV-1 Tat-mediated inhibition of the tyrosine hydroxylase gene expression in dopaminergic neuronal cells. J Biol Chem. 2000;275:4159–4165. doi: 10.1074/jbc.275.6.4159. [DOI] [PubMed] [Google Scholar]
  • 224.Theodore S, Cass WA, Maragos WF. Methamphetamine and human immunodeficiency virus protein Tat synergize to destroy dopaminergic terminals in the rat striatum. Neuroscience. 2006;137:925–935. doi: 10.1016/j.neuroscience.2005.10.056. [DOI] [PubMed] [Google Scholar]
  • 225.Cadet JL, Krasnova IN. Interactions of HIV and methamphetamine: cellular and molecular mechanisms of toxicity potentiation. Neurotox Res. 2007;12:181–204. doi: 10.1007/BF03033915. [DOI] [PubMed] [Google Scholar]
  • 226.Haughey NJ, et al. HIV-1 Tat through phosphorylation of NMDA receptors potentiates glutamate excitotoxicity. J Neurochem. 2001;78:457–467. doi: 10.1046/j.1471-4159.2001.00396.x. [DOI] [PubMed] [Google Scholar]
  • 227.Silvers JM, et al. Neurotoxicity of HIV-1 Tat protein: involvement of D1 dopamine receptor. Neurotoxicology. 2007;28:1184–1190. doi: 10.1016/j.neuro.2007.07.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228.Aksenova MV, et al. HIV-1 Tat neurotoxicity in primary cultures of rat midbrain fetal neurons: changes in dopamine transporter binding and immunoreactivity. Neurosci Lett. 2006;395:235–239. doi: 10.1016/j.neulet.2005.10.095. [DOI] [PubMed] [Google Scholar]
  • 229.Chang L, et al. Additive effects of HIV and chronic methamphetamine use on brain metabolite abnormalities. Am J Psychiatry. 2005;162:361–369. doi: 10.1176/appi.ajp.162.2.361. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 230.Langford D, et al. Patterns of selective neuronal damage in methamphetamine-user AIDS patients. J Acquir Immune Defic Syndr. 2003;34:467–474. doi: 10.1097/00126334-200312150-00004. [DOI] [PubMed] [Google Scholar]
  • 231.Chana G, et al. Cognitive deficits and degeneration of interneurons in HIV+ methamphetamine users. Neurology. 2006;67:1486–1489. doi: 10.1212/01.wnl.0000240066.02404.e6. [DOI] [PubMed] [Google Scholar]
  • 232.Liang H, et al. Methamphetamine enhances HIV infection of macrophages. Am J Pathol. 2008;172:1617–1624. doi: 10.2353/ajpath.2008.070971. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 233.Turchan J, et al. Estrogen protects against the synergistic toxicity by HIV proteins, methamphetamine and cocaine. BMC Neurosci. 2001;2:3. doi: 10.1186/1471-2202-2-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 234.Langford D, et al. The role of mitochondrial alterations in the combined toxic effects of human immunodeficiency virus Tat protein and methamphetamine on calbindin positive-neurons. J Neurovirol. 2004;10:327–337. doi: 10.1080/13550280490520961. [DOI] [PubMed] [Google Scholar]
  • 235.Maragos WF, et al. Human immunodeficiency virus-1 Tat protein and methamphetamine interact synergistically to impair striatal dopaminergic function. J Neurochem. 2002;83:955–963. doi: 10.1046/j.1471-4159.2002.01212.x. [DOI] [PubMed] [Google Scholar]
  • 236.Flora G, et al. Methamphetamine potentiates HIV-1 Tat protein-mediated activation of redox-sensitive pathways in discrete regions of the brain. Exp Neurol. 2003;179:60–70. doi: 10.1006/exnr.2002.8048. [DOI] [PubMed] [Google Scholar]
  • 237.Theodore S, Cass WA, Maragos WF. Involvement of cytokines in human immunodeficiency virus-1 protein Tat and methamphetamine interactions in the striatum. Exp Neurol. 2006;199:490–498. doi: 10.1016/j.expneurol.2006.01.009. [DOI] [PubMed] [Google Scholar]
  • 238.Cass WA, et al. HIV-1 protein Tat potentiation of methamphetamine-induced decreases in evoked overflow of dopamine in the striatum of the rat. Brain Res. 2003;984:133–142. doi: 10.1016/s0006-8993(03)03122-6. [DOI] [PubMed] [Google Scholar]
  • 239.Conant K, et al. Cerebrospinal fluid levels of MMP-2, 7, and 9 are elevated in association with human immunodeficiency virus dementia. Ann Neurol. 1999;46:391–398. doi: 10.1002/1531-8249(199909)46:3<391::aid-ana15>3.0.co;2-0. [DOI] [PubMed] [Google Scholar]
  • 240.Liuzzi GM, et al. Increased activity of matrix metalloproteinases in the cerebrospinal fluid of patients with HIV-associated neurological diseases. J Neurovirol. 2000;6:156–163. doi: 10.3109/13550280009013159. [DOI] [PubMed] [Google Scholar]
  • 241.Carrasco GA, Van de Kar LD. Neuroendocrine pharmacology of stress. Eur J Pharmacol. 2003;463:235–272. doi: 10.1016/s0014-2999(03)01285-8. [DOI] [PubMed] [Google Scholar]
  • 242.Leonard BE. HPA and immune axes in stress: involvement of the serotonergic system. Neuroimmunomodulation. 2006;13:268–276. doi: 10.1159/000104854. [DOI] [PubMed] [Google Scholar]
  • 243.Morton AJ, Hickey MA, Dean LC. Methamphetamine toxicity in mice is potentiated by exposure to loud music. Neuroreport. 2001;12:3277–3281. doi: 10.1097/00001756-200110290-00026. [DOI] [PubMed] [Google Scholar]
  • 244.Conrad CD, Jackson JL, Wise LS. Chronic stress enhances ibotenic acid-induced damage selectively within the hippocampal CA3 region of male, but not female rats. Neuroscience. 2004;125:759–767. doi: 10.1016/j.neuroscience.2004.01.049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 245.Matuszewich L, Yamamoto BK. Chronic stress augments the long-term and acute effects of methamphetamine. Neuroscience. 2004;124:637–646. doi: 10.1016/j.neuroscience.2003.12.007. [DOI] [PubMed] [Google Scholar]
  • 246.Smith LK, et al. Stress accelerates neural degeneration and exaggerates motor symptoms in a rat model of Parkinson’s disease. Eur J Neurosci. 2008;27:2133–2146. doi: 10.1111/j.1460-9568.2008.06177.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 247.Tata DA, Yamamoto BK. Chronic stress enhances methamphetamine-induced extracellular glutamate and excitotoxicity in the rat striatum. Synapse. 2008;62:325–336. doi: 10.1002/syn.20497. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 248.Watanabe Y, Gould E, McEwen BS. Stress induces atrophy of apical dendrites of hippocampal CA3 pyramidal neurons. Brain Res. 1992;588:341–345. doi: 10.1016/0006-8993(92)91597-8. [DOI] [PubMed] [Google Scholar]
  • 249.Bremner JD, et al. MRI-based measurement of hippocampal volume in patients with combat-related posttraumatic stress disorder. Am J Psychiatry. 1995;152:973–981. doi: 10.1176/ajp.152.7.973. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 250.Kibel A, Drenjancevic-Peric I. Impact of glucocorticoids and chronic stress on progression of Parkinson’s disease. Med Hypotheses. 2008;71:952–956. doi: 10.1016/j.mehy.2008.06.036. [DOI] [PubMed] [Google Scholar]
  • 251.Quinton MS, Yamamoto BK. Neurotoxic effects of chronic restraint stress in the striatum of methamphetamine-exposed rats. Psychopharmacology (Berl) 2007;193:341–350. doi: 10.1007/s00213-007-0796-x. [DOI] [PubMed] [Google Scholar]
  • 252.Tata DA, Raudensky J, Yamamoto BK. Augmentation of methamphetamine-induced toxicity in the rat striatum by unpredictable stress: contribution of enhanced hyperthermia. Eur J Neurosci. 2007;26:739–748. doi: 10.1111/j.1460-9568.2007.05688.x. [DOI] [PubMed] [Google Scholar]
  • 253.Scaccianoce S, Lombardo K, Angelucci L. Nerve growth factor brain concentration and stress: changes depend on type of stressor and age. Int J Dev Neurosci. 2000;18:469–479. doi: 10.1016/s0736-5748(00)00014-9. [DOI] [PubMed] [Google Scholar]
  • 254.Hahn M, Lorez H, Fischer G. Effect of calcitriol in combination with corticosterone, interleukin-1beta, and transforming growth factor-beta1 on nerve growth factor secretion in an astroglial cell line. J Neurochem. 1997;69:102–109. doi: 10.1046/j.1471-4159.1997.69010102.x. [DOI] [PubMed] [Google Scholar]
  • 255.Nitta A, et al. Brain-derived neurotrophic factor prevents neuronal cell death induced by corticosterone. J Neurosci Res. 1999;57:227–235. doi: 10.1002/(SICI)1097-4547(19990715)57:2<227::AID-JNR8>3.0.CO;2-E. [DOI] [PubMed] [Google Scholar]
  • 256.Nitta A, Zheng WH, Quirion R. Insulin-like growth factor 1 prevents neuronal cell death induced by corticosterone through activation of the PI3k/Akt pathway. J Neurosci Res. 2004;76:98–103. doi: 10.1002/jnr.20057. [DOI] [PubMed] [Google Scholar]
  • 257.Schaaf MJ, Cidlowski JA. Molecular determinants of glucocorticoid receptor mobility in living cells: the importance of ligand affinity. Mol Cell Biol. 2003;23:1922–1934. doi: 10.1128/MCB.23.6.1922-1934.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 258.Conway-Campbell BL, et al. Nuclear targeting of the growth hormone receptor results in dysregulation of cell proliferation and tumorigenesis. Proc Natl Acad Sci USA. 2007;104:13331–13336. doi: 10.1073/pnas.0600181104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 259.Kumar M, et al. HIV-1 infection and its impact on the HPA axis, cytokines, and cognition. Stress. 2003;6:167–172. doi: 10.1080/10253890310001605376. [DOI] [PubMed] [Google Scholar]
  • 260.Johnson SD. Substance abuse, post-traumatic stress disorder and violence. Curr Opin Psychiatry. 2008;21:242–246. doi: 10.1097/YCO.0b013e3282fc9889. [DOI] [PubMed] [Google Scholar]

RESOURCES