Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Jul 1.
Published in final edited form as: Int Rev Neurobiol. 2009;88:101–119. doi: 10.1016/S0074-7742(09)88005-7

MOLECULAR BASES OF METHAMPHETAMINE-INDUCED NEURODEGENERATION

Jean Lud Cadet 1, Irina N Krasnova 1
PMCID: PMC8247532  NIHMSID: NIHMS1716399  PMID: 19897076

Abstract

Methamphetamine (METH) is a highly addictive psychostimulant drug, whose abuse has reached epidemic proportions worldwide. The addiction to METH is a major public concern because its chronic abuse is associated with serious health complications including deficits in attention, memory, and executive functions in humans. These neuropsychiatric complications might, in part, be related to drug-induced neurotoxic effects, which include damage to dopaminergic and serotonergic terminals, neuronal apoptosis, as well as activated astroglial and microglial cells in the brain. Thus, the purpose of the present paper is to review cellular and molecular mechanisms that might be responsible for METH neurotoxicity. These include oxidative stress, activation of transcription factors, DNA damage, excitotoxicity, blood–brain barrier breakdown, microglial activation, and various apoptotic pathways. Several approaches that allow protection against METH-induced neurotoxic effects are also discussed. Better understanding of the cellular and molecular mechanisms involved in METH toxicity should help to generate modern therapeutic approaches to prevent or attenuate the long-term consequences of psychostimulant use disorders in humans.

I. Epidemiology of Methamphetamine Abuse

Methamphetamine (METH) is an illegal psychostimulant that is abused by more than 25 million people in the world, which exceeds the amount of people who use heroin and cocaine (Rawson and Condon, 2007; Romanelli and Smith, 2006). METH is the most commonly synthesized illegal drug in the United States and has been cited by law enforcement officials as the leading crime problem in the country (Gettig et al., 2006). The inexpensive production of the drug, its low cost, and its long duration of action make it very desirable to abusers. Recently, METH abuse has reached epidemic proportions in the United States, with populations in the Western, Southern, and Midwest states being the most affected (Gettig et al., 2006; Rawson and Condon, 2007; Topolski, 2007). A 2003 survey found that 5.2% of the American adults have used METH at least once (Roehr, 2005). In addition, 5.3% and 6.2% of high school sophomores and seniors have tried METH (Gettig et al., 2006). METH-related emergency room admissions have also increased from 10 to 52 per 100,000 people between 1992 and 2002 (Roehr, 2005). Although METH abuse has been associated, traditionally, with blue-collar construction workers, truck drivers, and motorcycle gangs, the profile of the typical METH abusing individual has shifted due to the increased popularity among college students, women, and young professionals (Gettig et al., 2006; Romanelli and Smith, 2006). METH use has also highly increased in men who have sex with men resulting in a greater frequency of METH abuse in homosexual and bisexual men than in the general population (Shoptaw, 2006).

II. Clinical Toxicology of METH Abuse

METH abuse causes serious health complications in humans. METH users develop acute clinical symptoms that include agitation, anxiety, aggressive behaviors, paranoia, hypertension, hyperthermia, and psychosis (Albertson et al., 1999; Lynch and House, 1992; Murray, 1998). Ingestions of large doses of the drug can cause life-threatening hyperthermia above 41 °C, cardiac arrythmias, heart attacks, cerebrovascular hemorrhages, strokes, seizures, renal, and liver failure (Albertson et al., 1999; Perez et al., 1999).

Clinical studies have also shown the potential neuropathological consequences of METH abuse, which include deficits in attention, working memory, and executive functions in chronic psychostimulant abusers (Gonzalez et al., 2004; Rippeth et al., 2004; Salo et al., 2002; Sim et al., 2002; Simon et al., 2002, 2004; Woods et al., 2005). The drug can also cause neurodegenerative changes in the brains of human addicts. These include persistent loss of dopamine transporters (DAT) observed in the cortex and caudate-putamen (Sekine et al., 2003; Volkow et al., 2001), loss of serotonin transporters (5-HTT) in the cortex (Sekine et al., 2006), and decrease in the levels of dopamine (DA) and its metabolites in the caudate-putamen of METH abusers (Moszczynska et al., 2004; Wilson et al., 1996).

In agreement with clinical findings, a number of animal studies have shown that METH can cause long-term damage to presynaptic dopaminergic and serotonergic axons in rodents (Ricaurte et al., 1980; Wagner et al., 1980). More recently, it has also been shown that the drug can cause death of the neuronal bodies in the brain through apoptosis (Deng and Cadet, 1999, 2000; Eisch et al., 1998; O’Dell and Marshall, 2000; Zhu et al., 2006). In what follows, we discuss some of the mechanisms that might underlie these METH-induced neurodegenerative effects.

III. Role of Oxidative Stress in METH Toxicity

Toxic effects of METH are thought to depend on the similarity of its chemical structure to DA, which allows the drug to enter DA axons (Iversen, 2006), followed by DA release from synaptic vesicles into cytoplasm and by reverse transport into the synaptic cleft (Sulzer et al., 2005). METH neurotoxicity depends on the formation of DA quinones and superoxide radicals within nerve terminals (LaVoie and Hastings, 1999). DA metabolism by MAO can also increase hydrogen peroxide production, followed by its interactions with metal ions to form very toxic hydroxyl radicals (Cadet and Brannock, 1998). Evidence has accumulated to indicate that METH can also cause oxidative stress by switching the balance between ROS production and the capacity of antioxidant enzyme system to scavenge ROS (Gluck et al., 2001; Harold et al., 2000; Jayanthi et al., 1998). Excessive production of ROS that overwhelms this system can damage cellular components such as lipids, proteins, mitochondrial and nuclear DNA (Potashkin and Meredith, 2006).

A role for oxidative mechanisms in the drug toxicity is consistent with findings that pretreatment with N-acetyl-l-cysteine, ascorbic acid, or vitamin E allows protection against METH-induced depletion of monoaminergic axons (De Vito and Wagner, 1989; Fukami et al., 2004; Wagner et al., 1985). The role for superoxide radicals in the neurotoxic effects of METH on DA axons was tested by injecting METH to transgenic mice that overexpress the human CuZn superoxide dismutase (CuZnSOD) gene (Cadet et al., 1994; Hirata et al., 1996; Jayanthi et al., 1998). These mice have much higher CuZnSOD enzyme activity than control wild-type animals (Jayanthi et al., 1998, 1999) and were protected against METH toxicity. In addition, bromocriptine, which scavengers hydroxyl radicals, also attenuates METH-induced DA depletion in mice (Kondo et al., 1994). Together, these findings support the proposition that DA release caused by METH is accompanied by redox cycling of DA quinones and formation of superoxide radicals. The hypothesis that oxygen-based free radicals are involved in METH toxicity (Cadet and Brannock, 1998) is also supported by reports that the drug can reduce the levels of glutathione (Harold et al., 2000) and antioxidant enzymes (Jayanthi et al., 1998), increase lipid peroxidation (Gluck et al., 2001; Jayanthi et al., 1998), and cause the formation of protein carbonyls (Gluck et al., 2001).

Recent evidence also indicates that changes in nitric oxide (NO) metabolism can contribute to METH-induced oxidative stress and neurotoxicity (Itzhak and Ali, 2006). NO can react with superoxide radicals to form peroxynitrite, a strong oxidant and major neurotoxin (Pacher et al., 2007). Indeed, increased levels of a marker for peroxynitrite production 3-NT have been measured in vivo and in vitro in response to METH treatment (Imam et al., 2001a). Antioxidants selenium and melationin completely block the formation of 3-NT and striatal DA depletion (Imam et al., 2001a). In addition, free radical scavenger, edaravone, blocked METH-related increase in 3-NT immunoreactivity with subsequent attenuation of DA depletion and of reduction in TH immunoreactivity in the striatum (Kawasaki et al., 2006). Furthermore, the selective neuronal nitric oxide synthase (nNOS) inhibitor, 7-nitroindazole, protects against drug-induced formation of 3-NT, as well as DA and 5-HT depletion in the striatum (Ali and Itzhak, 1998; Di Monte et al., 1996; Itzhak and Ali, 1996). The participation of NO in METH toxicity is also proposed by findings that nNOS knockout mice were protected against the formation of 3-NT and DA terminal degeneration in the striatum (Imam et al., 2001b). Together, these data strongly support the hypothesis that NO and peroxynitrite, in particular, are involved in the mechanisms underlying METH-induced monoaminergic neurotoxicity (Imam et al., 2001b; Itzhak and Ali, 2006; Itzhak et al., 1998).

IV. Involvement of AP-1-Related Transcription Factors in METH-Induced Neurotoxicity

The accumulated data suggested that effects of METH might be mediated, in part, by activation of AP-1 transcription factors. These include upregulation of c-jun, c-fos, jun B, and jun D expression within 2 h after METH administration (Cadet et al., 2001). These changes, in turn, might be related to METH-induced generation of ROS. Specifically, hydroxyl and superoxide radicals can induce the expression of many genes via regulation of AP-1 transcription factors (Dalton et al., 1999). The role for c-fos in METH-induced neuropathological changes has been confirmed by using c-fos heterozygote mice that show increased degeneration of DA axons and increased cell death after psychostimulant treatment (Deng et al., 1999). These findings support a protective role for c-fos against METH damage. The factors that could be involved in this protection include cell adhesion receptors integrins because of decreased basal levels of integrin expression in c-fos heterozygote mice and the further reduction of these receptors in response to toxic doses of METH (Betts et al., 2002). This idea is also supported by the observations that integrins promote cell survival after injury and apoptotic insults via PI3K-Akt pathway which leads to phosphorylation of proapoptotic protein BAD, therefore, reducing its ability to block the antiapoptotic effects of Bcl-2 (Gilcrease, 2007). In contrast, inhibition of integrins increases apoptotic cell death (Gilcrease, 2007).

c-jun is another AP-1 transcription factor that might be involved in METH toxicity, because c-jun knockout mice show partial protection against damaging effects of the drug (Deng et al., 2002b). Moreover, because the c-jun knockout mice and their wild-type littermates show similar degree of dopaminergic toxicity after METH treatment, c-jun appears to only play role in the mediation of neuronal apoptosis in cells postsynaptic to DA axons.

V. Role of DNA Damage in METH-Induced Toxicity

As mentioned earlier, METH can cause neuronal apoptosis in several brain regions, including striatum, cortex, hippocampus, and olfactory bulb (Deng et al., 2001). Because apoptosis is associated with DNA damage, it was possible that treatment with METH might induce responses involved in the repair of the drug-related DNA damage. Data obtained using microarray analyses have shown that METH administration caused changes in the expression of a number of genes that participate in DNA repair, including APEX, PolB, and LIG1 (Cadet et al., 2002). These changes are probably related to METH-induced increase in the levels of free radicals because oxidative stress can cause single and double DNA strand breaks (Li and Trush, 1993). Thus, the upregulation of DNA repair genes following METH treatment suggests that these changes might be compensatory to counteract METH-related ROS-induced DNA damage. If the psychostimulant can cause similar DNA damages in humans, this might account for developmental deficits observed in children born of METH abusing mothers (Smith et al., 2006).

VI. METH Toxicity and Excitotoxicity

METH neurotoxicity might also occur via excitotoxic damage following glutamate release and activation of glutamate receptors. Glutamate toxicity is dependent, in part, on the production of NO (Chung et al., 2005). The hypothesis on the involvement of glutamate in METH toxicity is supported by findings that METH causes glutamate release in the brain (Abekawa et al., 1994; Baldwin et al., 1993; Mark et al., 2004; Marshall et al., 1993; Nash et al., 1988). In addition, some glutamate antagonists can attenuate METH-induced dopaminergic toxicity (Battaglia et al., 2002; Sonsalla et al., 1989). Glutamate-mediated NO formation might also be involved in METH toxicity because knockout mice deficient in either nNOS or iNOS (inducible nitric oxide synthase) are protected against psychostimulant-induced damage to monoaminergic axons (Itzhak et al., 1998). These data have provided strong support for the idea that the glutamate/NO pathway plays major role in METH neurotoxicity (Imam et al., 2001b; Itzhak and Ali, 2006; Itzhak et al., 1998). Finally, various nNOS inhibitors can also protect against depletion of monoaminergic axons caused by METH administration (Itzhak et al., 2000; Sanchez et al., 2003). In addition to their roles in the damage of monoaminergic axons, oxygen-based radicals and NO may be involved in METH-related neuronal death because CuZnSOD transgenic mice show partial protection against drug-induced apoptosis (Deng and Cadet, 2000).

VII. Role of Blood–Brain Barrier Dysfunction in METH Toxicity

Several recent papers have examined the effects of METH on the blood–brain barrier (BBB) and their potential relationships to METH toxicity (Bowyer and Ali, 2006; Bowyer et al., 2008; Kiyatkin et al., 2007; Sharma and Ali, 2006; Sharma et al., 2007). Using protein tracers and albumin immunohistochemistry, METH was shown to cause marked disruption of BBB at the levels of the cortex, hippocampus, thalamus, hypothalamus, cerebellum, and amygdala (Bowyer and Ali, 2006; Kiyatkin et al., 2007; Sharma et al., 2007). METH-induced BBB breakdown was evidenced by leakage of serum albumin into the brain tissue (Sharma et al., 2007). Doses of METH that cause BBB disturbances also induce neuronal damage, myelin degeneration, and reactive astrocytosis in the parietal and occipital cortices (Sharma et al., 2007). These doses also cause extensive degeneration of pyramidal cells and activation of microglia in amygdala and hippocampus of rats (Bowyer and Ali, 2006). These effects appear to depend on hyperthermia because the psychostimulant failed to induce BBB damage and neurodegeneration in the brains of animals that did not show increased temperature (Bowyer and Ali, 2006). Interestingly, mild BBB dysfunction found in the caudate-putamen after METH treatment was exacerbated by hyperthermia (Bowyer et al., 2008). It is interesting to point out that METH-induced leakage of serum albumin into brain tissue was attenuated by pretreatment with antioxidant, H-290/51, suggesting the involvement of free radicals in BBB damage (Sharma et al., 2007) and further supporting a role for oxidative stress in drug toxicity. Also of interest is the fact the antioxidant was able to attenuate METH-induced hyperthermia, neuronal damage, myelin degradation, and glial response (Sharma et al., 2007).

VIII. Involvement of Mitochondrial Death Pathway in METH-Induced Apoptosis

In addition to glutamate and NO, the Bcl-2 family of proteins may also be involved in the mechanisms underlying METH neurotoxicity (Cadet et al., 2001; Jayanthi et al., 2001; Stumm et al., 1999). In particular, METH induced increases in proapoptotic proteins, BAX and BID, and decreases in antiapoptotic proteins, Bcl-2 and Bcl-XL. The upregulation of proapoptotic proteins is consistent with findings that METH treatment caused release of mitochondrial proteins cytochrome c and apoptosis inducing factor (AIF) into the cytosol (Deng et al., 2002a; Jayanthi et al., 2004). AIF and Smac/DIABLO released from mitochondria have also been shown to participate in METH-induced apoptosis (Jayanthi et al., 2004). Their release is followed by activation of caspases 9 and 3, and the breakdown of several structural cellular proteins (Jayanthi et al., 2004). Thus, these findings implicate the mitochondrial pathway in METH-related neuronal death in the brain (Cadet et al., 2005, 2007). This idea is consistent with the findings that overexpression of Bcl-2 can protect against drug-induced apoptosis (Cadet et al., 1997).

IX. Involvement of the Endoplasmic Reticulum-Dependent Death Pathway in METH-Induced Apoptosis

Oxidative stress can trigger cellular damage by causing dysfunctions of cellular organelles such as the endoplasmic reticulum (ER) (Gorlach et al., 2006). In addition to regulating synthesis, folding, and transport of proteins, ER also constitutes the main intracellular store for Ca2+, whose excess can contribute to cell death (Gorlach et al., 2006). At physiological levels, Ca2+ released from the ER is taken up by mitochondria to enhance metabolite flow on the outer mitochondrial membrane and to increase ATP production (Kroemer et al., 2007). However, sustained release of Ca2+ from the ER stores may initiate calcium-dependent apoptosis via the permeabilization of the outer mitochondrial membrane (Kroemer et al., 2007). ER stress and dysregulation of calcium homeostasis appear to participate in METH-induced cell death because the drug can induce activation of calpain (Jayanthi et al., 2004), a calcium-responsive cytosolic protease involved in ER-dependent apoptosis (Nakagawa and Yuan, 2000). METH has been shown to increase calpain-mediated protolysis of cytoskeletal protein spectrin and microtubule protein tau in the rat cortex, hippocampus (Warren et al., 2005), and striatum (Staszewski and Yamamoto, 2006). In contrast, the calpain inhibitors can attenuate psychostimulant-induced spectrin and tau proteolysis (Warren et al., 2007) as well as neuronal death (Samantaray et al., 2006), strongly implicating ER stress and calpain activation in the mechanisms of METH neuronal degeneration. A role for the ER in METH toxicity is further supported by the findings that apoptotic doses of the drug can increase the expression of proteins such as caspase-12, GRP78/BiP, and CHOP/GADD153 (Jayanthi et al., 2004) that participate in ER-induced apoptosis (Marciniak and Ron, 2006). This METH-related ER stress might be secondary to oxidative stress (Cadet and Brannock, 1998; Cadet et al., 1994; Jayanthi et al., 1998) and to increases in BAX/Bcl-2 ratios induced by this illicit drug (Jayanthi et al., 2001).

X. Microglial Reactions and METH Toxicity

Microglia are the resident immune cells within CNS that function to protect the brain against injury and chemical damage (Raivich, 2005). In the healthy mature brain, microglia typically exist in a resting state characterized by ramified morphology, and monitor the neuronal environment (Block et al., 2007; Raivich, 2005). However, in response to brain injury or damage, microglia are readily activated, undergoing a dramatic transformation from a resting ramified state into an amoeboid morphology. Although microglial activation is necessary for host defense and neuron survival, the overactivation of microglia results in deleterious and neurotoxic consequences. Specifically, microglia contributes to the progress of many neurodegenerative diseases, including Parkinson’s (Kim and Joh, 2006), Alzheimer’s (Xiang et al., 2006), and Huntington (Sapp et al., 2001) diseases, AIDS-related neuropathy (Gonzalez-Scarano and Martin-Garcia, 2005) as well as in the neurotoxic effects of MPTP (Gao et al., 2002) and kainic acid (Chen et al., 2005). Once activated, microglia become big, migrate to the site of the injury, and cause phagocytosis of dying and dead cells. In addition, microglia secrete a variety of cytokines, reactive oxygen and nitrogen species, and prostaglandins that are known to cause neuronal damage (Block et al., 2007; Perry et al., 2007).

Recently, emerging data have implicated microglial activation as an early event in the neurotoxic cascade that is initiated by METH treatment. Specifically, METH causes strong microglial response in the areas of the brain that show DA axonal degeneration (Thomas et al., 2004b). In contrast, attenuation of METH neurotoxicity by MK-801 and dextromethorphan inhibits microglial activation (Thomas and Kuhn, 2005). Moreover, anti-inflammatory drug ketoprofen allows some protection against METH toxicity and also reduces psychostimulant-induced microgliosis (Asanuma et al., 2003). However, attenuation of microglial activation itself is insufficient to protect against METH neurotoxicity (Sriram et al., 2006b). This microglial activation precedes METH-induced DA axonal degeneration in the striatum, suggesting that microglia might contribute to drug toxicity (LaVoie et al., 2004).

While neurotoxic amphetamines METH, MDMA, amphetamine and p-chloroamphetamine cause microglial activation (Thomas et al., 2004a) in addition to reactive astrocytosis (Deng and Cadet, 1999; Krasnova et al., 2005; O’Callaghan et al., 1995; Xu et al., 2005), nonneurotoxic drugs such as fenfluramine and DOI fail to activate microglia (Thomas et al., 2004a). These data establish a link between the neurotoxic amphetamines and microglial activation, suggesting that microglia might be a selective marker for neuronal axonal damage (Thomas et al., 2004a) in agreement with the ability for the same drugs to cause reactive astrocytosis, an established hallmark of neurotoxicity (O’Callaghan and Sriram, 2005).

Microglial cells might potentiate METH-related damage by releasing toxic substances such as superoxide radicals and NO which have already been implicated in drug neurotoxicity (see discussion above). In addition, METH causes increase in the levels of TNF-α and IL-β (Flora et al., 2002; Sriram et al., 2006a), proinflammatory cytokines that can also contribute to toxicity of the drug. Consistent with these findings, METH neurotoxicity and an increase in a marker for microglial activation PK11195 binding were attenuated in IL-6 null mice (Ladenheim et al., 2000). Together, these observations show that inflammatory reactions in microglia might participate in the molecular pathway underlying METH toxicity in the brain.

A schematic diagram showing molecular mechanisms that lead to METH-induced neuronal degeneration is presented in Fig. 1.

FIG. 1.

FIG. 1.

Mechanisms implicated in METH-induced neurotoxicity. The figure summarizes the pathways that are reviewed in this chapter.

XI. Neuroprotective Mechanisms and METH Toxicity

Several attempts have been made to identify ways to protect the brain against METH toxicity. DA uptake inhibitors and DA receptor antagonists have been shown to provide protection against METH-induced degeneration of striatal DA terminals (Angulo et al., 2004; Jayanthi et al., 2005; Marek et al., 1990; O’Dell et al., 1993; Schmidt and Gibb, 1985; Sonsalla et al., 1986). The DA D1 antagonist, SCH23390, also protects against METH-induced cell death in the striatum (Jayanthi et al., 2005). These neuroprotective effects may depend on changes in DA release because DA receptor antagonists, SCH23390 and eticlopride, were reported to partially block METH-related increases in DA release (O’Dell et al., 1993). DA D1-depending mechanisms are also mediated via Fas/FasL-related events since prior treatment with SCH23390 decreased translocation of NFATc3 and NFATc4 from the cytoplasm to the nucleus and reduced increases in calcineurin expression caused by METH administration (Jayanthi et al., 2005). Pretreatment with SCH23390 also resulted in significant inhibition of the METH-induced increases in the expression of FasL and caspase-3 in rat striatal cells (Jayanthi et al., 2005). Because the dose of SCH23390 used in that study completely blocks METH-induced decreases in DA levels while providing only partial protection against death of striatal neurons (Jayanthi et al., 2005), the possibility exists that METH-induced cell death might involved additional mechanisms independent of stimulation of DA D1 receptors.

In addition to DAT inhibitors and DA receptor antagonists, some trophic factors also provide protection against the toxic effects of METH in vitro (Chou et al., 2008; Mamounas et al., 1995; Matsuzaki et al., 2004; Zhou et al., 2004) and in vivo (Cass, 1996; Chou et al., 2008; Melega et al., 2000). Specifically, brain-derived neurotrophic factor (BDNF) was shown to prevent METH-induced caspase-3 activation and death in primary cultures of cortical neurons (Matsuzaki et al., 2004). The protective effect was blocked by phosphatydilinositol-3-kinase inhibitors and by the expression of kinase-deficient Akt, thus implicating the PI3K/Akt pathway in the effects of BDNF (Matsuzaki et al., 2004). Glial cell line-derived neurotrophic factor (GDNF) administration can also prevent METH-mediated reductions in DA levels in the rat striatum (Cass, 1996; Cass et al., 2006) and hasten the recovery of striatal DA functions in METH-treated rats (Cass et al., 2000). GDNF pretreatment could also partially prevent the loss of DAT binding caused by METH administration in the striatum of vervet monkeys (Melega et al., 2000). Bone morphogenetivc protein 7 (BMP7) can also attenuate METH-induced decrease in TH immunoreactivity and cell death in primary DA neurons and in the mouse striatum (Chou et al., 2008) while nerve growth factor was reported to protect R2 cells against drug-related DNA fragmentation and apoptosis (Zhou et al., 2004).

Several studies have demonstrated that METH neurotoxicity can be mediated, in part, via activation of neuroinflammatory responses (Kuhn et al., 2006; Ladenheim et al., 2000; Thomas et al., 2004b). In line with these findings, some cytokines have been shown to induce protection against METH toxicity. For example, METH-induced damage to DA axons was prevented by pretreatment with interferon γ, possibly, by suppression of its neuroinflammatory effects (Hozumi et al., 2008). In addition, interferon γ might cause neuroprotective effects by increasing GDNF production in astrocytes (Appel et al., 1997). Another cytokine, TNF-α, which can enhance autoimmunity and inflammation (Sriram and O’Callaghan, 2007), is protective against METH-induced toxicity (Nakajima et al., 2004). The neuroprotective effects of TNF-α may be mediated by inhibition of METH-related increases in extracellular DA levels in the striatum and by potentiating DA uptake into synaptosomes and synaptic vesicles (Nakajima et al., 2004). The neuroprotection against METH toxicity caused by TNF-α may also be mediated via its effects on temperature regulation, because TNF-α pretreatment prior to drug administration also induced hypothermia, whereas METH-induced hyperthermia was exacerbated in TNF-α knockout mice (Nakajima et al., 2004). In addition, TNF-α has been shown to cause upregulation of MnSOD expression through activation of NFκB in hippocampal neurons (Mattson et al., 1997) because MnSOD transgenic mice are protected against METH toxicity (Maragos et al., 2000).

Estrogen can also protect against METH-induced damage to nigrostriatal DA system (Dluzen and McDermott, 2006). Estrogen treatment caused neuroprotective effects against METH neurotoxicity in female, but not in male mice (D’Astous et al., 2004, 2005; Dluzen et al., 2002; Liu and Dluzen, 2006). Pretreatment with tamoxifen, an estrogen receptor ligand, attenuates METH-induced striatal DA depletion and the decrease in DAT binding via mechanisms that were independent of thermoregulation (Bourque et al., 2007; Dluzen et al., 2001). In addition to its effects on estrogen receptors, tamoxifen can also cause anti-inflammatory responses in microglial cells (Suuronen et al., 2005) and protect glial cells against glutamate toxicity (Shy et al., 2000). Because glutamate release and microgliosis contribute to METH toxicity (see above), the protective effects of tamoxifen probably occur via suppression of both glutamate- and microglia-dependent toxic mechanisms.

Acknowledgments

This work is supported by the Intramural Research Program of the National Institute on Drug Abuse, NIH, DHHS.

References

  1. Abekawa T, Ohmori T, and Koyama T (1994). Effects of repeated administration of a high dose of methamphetamine on dopamine and glutamate release in rat striatum and nucleus accumbens. Brain Res 643, 276–281. [DOI] [PubMed] [Google Scholar]
  2. Albertson TE, Derlet RW, and Van Hoozen BE (1999). Methamphetamine and the expanding complications of amphetamines. West. J. Med 170, 214–219. [PMC free article] [PubMed] [Google Scholar]
  3. Ali SF, and Itzhak Y (1998). Effects of 7-nitroindazole, an NOS inhibitor on methamphetamine-induced dopaminergic and serotonergic neurotoxicity in mice. Ann. NY Acad. Sci 844, 122–130. [PubMed] [Google Scholar]
  4. Angulo JA, Angulo N, and Yu J (2004). Antagonists of the neurokinin-1 or dopamine D1 receptors confer protection from methamphetamine on dopamine terminals of the mouse striatum. Ann. NY Acad. Sci 1025, 171–180. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Appel E, Kolman O, Kazimirsky G, Blumberg PM, and Brodie C (1997). Regulation of GDNF expression in cultured astrocytes by inflammatory stimuli. Neuroreport 8, 3309–3312. [DOI] [PubMed] [Google Scholar]
  6. Asanuma M, Tsuji T, Miyazaki I, Miyoshi K, and Ogawa N (2003). Methamphetamine-induced neurotoxicity in mouse brain is attenuated by ketoprofen, a non-steroidal anti-inflammatory drug. Neurosci. Lett 352, 13–16. [DOI] [PubMed] [Google Scholar]
  7. Baldwin HA, Colado MI, Murray TK, De Souza RJ, and Green AR (1993). Striatal dopamine release in vivo following neurotoxic doses of methamphetamine and effect of the neuroprotective drugs, chlormethiazole and dizocilpine. Br. J. Pharmacol 108, 590–596. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Battaglia G, Fornai F, Busceti CL, Aloisi G, Cerrito F, De Blasi A, Melchiorri D, and Nicoletti F (2002). Selective blockade of mGlu5 metabotropic glutamate receptors is protective against methamphetamine neurotoxicity. J. Neurosci 22, 2135–2141. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Betts ES, Krasnova IN, McCoy MT, Ladenheim B, and Cadet JL (2002). Analysis of methamphetamine-induced changes in the expression of integrin family members in the cortex of wild-type and c-fos knockout mice. Neurotox. Res 4, 617–623. [DOI] [PubMed] [Google Scholar]
  10. Block ML, Zecca L, and Hong JS (2007). Microglia-mediated neurotoxicity: Uncovering the molecular mechanisms. Nat. Rev. Neurosci 8, 57–69. [DOI] [PubMed] [Google Scholar]
  11. Bourque M, Liu B, Dluzen DE, and Di Paolo T (2007). Tamoxifen protects male mice nigrostriatal dopamine against methamphetamine-induced toxicity. Biochem. Pharmacol 74, 1413–1423. [DOI] [PubMed] [Google Scholar]
  12. Bowyer JF, and Ali S (2006). High doses of methamphetamine that cause disruption of the blood-brain barrier in limbic regions produce extensive neuronal degeneration in mouse hippocampus. Synapse 60, 521–532. [DOI] [PubMed] [Google Scholar]
  13. Bowyer JF, Robinson B, Ali S, and Schmued LC (2008). Neurotoxic-related changes in tyrosine hydroxylase, microglia, myelin, and the blood-brain barrier in the caudate-putamen from acute methamphetamine exposure. Synapse 62, 193–204. [DOI] [PubMed] [Google Scholar]
  14. Cadet JL, and Brannock C (1998). Free radicals and the pathobiology of brain dopamine systems. Neurochem. Int 32, 117–131. [DOI] [PubMed] [Google Scholar]
  15. Cadet JL, Ali S, and Epstein C (1994). Involvement of oxygen-based radicals in methamphetamine-induced neurotoxicity: Evidence from the use of CuZnSOD transgenic mice. Ann. NY Acad. Sci 738, 388–391. [DOI] [PubMed] [Google Scholar]
  16. Cadet JL, Ordonez SV, and Ordonez JV (1997). Methamphetamine induces apoptosis in immortalized neural cells: Protection by the proto-oncogene, bcl-2. Synapse 25, 176–184. [DOI] [PubMed] [Google Scholar]
  17. Cadet JL, Jayanthi S, McCoy MT, Vawter M, and Ladenheim B (2001). Temporal profiling of methamphetamine-induced changes in gene expression in the mouse brain: Evidence from cDNA array. Synapse 41, 40–48. [DOI] [PubMed] [Google Scholar]
  18. Cadet JL, McCoy MT, and Ladenheim B (2002). Distinct gene expression signatures in the striata of wild-type and heterozygous c-fos knockout mice following methamphetamine administration: Evidence from cDNA array analyses. Synapse 44, 211–226. [DOI] [PubMed] [Google Scholar]
  19. Cadet JL, Jayanthi S, and Deng X (2005). Methamphetamine-induced neuronal apoptosis involves the activation of multiple death pathways. Rev. Neurotox. Res 8, 199–206. [DOI] [PubMed] [Google Scholar]
  20. Cadet JL, Krasnova IN, Jayanthi S, and Lyles J (2007). Neurotoxicity of substituted amphetamines: Molecular and cellular mechanisms. Neurotox. Res 11, 183–202. [DOI] [PubMed] [Google Scholar]
  21. Cass WA (1996). GDNF selectively protects dopamine neurons over serotonin neurons against the neurotoxic effects of methamphetamine. J. Neurosci 16, 8132–8139. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Cass WA, Manning MW, and Bailey SL (2000). Restorative effects of GDNF on striatal dopamine release in rats treated with neurotoxic doses of methamphetamine. Ann. NY Acad. Sci 914, 127–136. [DOI] [PubMed] [Google Scholar]
  23. Cass WA, Peters LE, Harned ME, and Seroogy KB (2006). Protection by GDNF and other trophic factors against the dopamine-depleting effects of neurotoxic doses of methamphetamine. Ann. NY Acad. Sci 1074, 272–281. [DOI] [PubMed] [Google Scholar]
  24. Chen Z, Duan RS, Quezada HC, Mix E, Nennesmo I, Adem A, Winblad B, and Zhu J (2005). Increased microglial activation and astrogliosis after intranasal administration of kainic acid in C57BL/6 mice. J. Neurobiol 62, 207–218. [DOI] [PubMed] [Google Scholar]
  25. Chou J, Luo Y, Kuo CC, Powers K, Shen H, Harvey BK, Hoffer BJ, and Wang Y (2008). Bone morphogenetic protein-7 reduces toxicity induced by high doses of methamphetamine in rodents. Neuroscience 151, 92–103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Chung KK, Dawson TM, and Dawson VL (2005). Nitric oxide, S-nitrosylation and neurodegeneration. Cell. Mol. Biol. (Noisy-le-grand) 51, 247–254. [PubMed] [Google Scholar]
  27. D’Astous M, Gajjar TM, Dluzen DE, and Di Paolo T (2004). Dopamine transporter as a marker of neuroprotection in methamphetamine-lesioned mice treated acutely with estradiol. Neuroendocrinology 79, 296–304. [DOI] [PubMed] [Google Scholar]
  28. D’Astous M, Mickley KR, Dluzen DE, and Di Paolo T (2005). Differential protective properties of estradiol and tamoxifen against methamphetamine-induced nigrostriatal dopaminergic toxicity in mice. Neuroendocrinology 82, 111–120. [DOI] [PubMed] [Google Scholar]
  29. Dalton TP, Shertzer HG, and Puga A (1999). Regulation of gene expression by reactive oxygen. Annu. Rev. Pharmacol. Toxicol 39, 67–101. [DOI] [PubMed] [Google Scholar]
  30. Deng X, and Cadet JL (1999). Methamphetamine administration causes overexpression of nNOS in the mouse striatum. Brain Res 851, 254–257. [DOI] [PubMed] [Google Scholar]
  31. Deng X, and Cadet JL (2000). Methamphetamine-induced apoptosis is attenuated in the striata of copper-zinc superoxide dismutase transgenic mice. Brain Res. Mol. Brain Res 83, 121–124. [DOI] [PubMed] [Google Scholar]
  32. Deng X, Ladenheim B, Tsao LI, and Cadet JL (1999). Null mutation of c-fos causes exacerbation of methamphetamine-induced neurotoxicity. J. Neurosci 19, 10107–10115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Deng X, Wang Y, Chou J, and Cadet JL (2001). Methamphetamine causes widespread apoptosis in the mouse brain: Evidence from using an improved TUNEL histochemical method. Brain Res. Mol. Brain Res 93, 64–69. [DOI] [PubMed] [Google Scholar]
  34. Deng X, Cai NS, McCoy MT, Chen W, Trush MA, and Cadet JL (2002a). Methamphetamine induces apoptosis in an immortalized rat striatal cell line by activating the mitochondrial cell death pathway. Neuropharmacology 42, 837–845. [DOI] [PubMed] [Google Scholar]
  35. Deng X, Jayanthi S, Ladenheim B, Krasnova IN, and Cadet JL (2002b). Mice with partial deficiency of c-Jun show attenuation of methamphetamine-induced neuronal apoptosis. Mol. Pharmacol 62, 993–1000. [DOI] [PubMed] [Google Scholar]
  36. De Vito MJ, and Wagner GC (1989). Methamphetamine-induced neuronal damage: A possible role for free radicals. Neuropharmacology 28, 1145–1150. [DOI] [PubMed] [Google Scholar]
  37. Di Monte DA, Royland JE, Jakowec MW, and Langston JW (1996). Role of nitric oxide in methamphetamine neurotoxicity: Protection by 7-nitroindazole, an inhibitor of neuronal nitric oxide synthase. J. Neurochem 67, 2443–2450. [DOI] [PubMed] [Google Scholar]
  38. Dluzen DE, and McDermott JL (2006). Estrogen, testosterone, and methamphetamine toxicity. Ann. NY Acad. Sci 1074, 282–294. [DOI] [PubMed] [Google Scholar]
  39. Dluzen DE, McDermott JL, and Anderson LI (2001). Tamoxifen diminishes methamphetamine-induced striatal dopamine depletion in intact female and male mice. J. Neuroendocrinol 13, 618–624. [DOI] [PubMed] [Google Scholar]
  40. Dluzen DE, Anderson LI, and Pilati CF (2002). Methamphetamine-gonadal steroid hormonal interactions: Effects upon acute toxicity and striatal dopamine concentrations. Neurotoxicol. Teratol 24, 267–273. [DOI] [PubMed] [Google Scholar]
  41. Eisch AJ, Schmued LC, and Marshall JF (1998). Characterizing cortical neuron injury with Fluoro-Jade labeling after a neurotoxic regimen of methamphetamine. Synapse 30, 329–333. [DOI] [PubMed] [Google Scholar]
  42. Flora G, Lee YW, Nath A, Maragos W, Hennig B, and Toborek M (2002). Methamphetamine-induced TNF-alpha gene expression and activation of AP-1 in discrete regions of mouse brain: Potential role of reactive oxygen intermediates and lipid peroxidation. Neuromolecular Med 2, 71–85. [DOI] [PubMed] [Google Scholar]
  43. Fukami G, Hashimoto K, Koike K, Okamura N, Shimizu E, and Iyo M (2004). Effect of antioxidant N-acetyl-L-cysteine on behavioral changes and neurotoxicity in rats after administration of methamphetamine. Brain Res 1016, 90–95. [DOI] [PubMed] [Google Scholar]
  44. Gao HM, Jiang J, Wilson B, Zhang W, Hong JS, and Liu B (2002). Microglial activation-mediated delayed and progressive degeneration of rat nigral dopaminergic neurons: Relevance to Parkinson’s disease. J. Neurochem 81, 1285–1297. [DOI] [PubMed] [Google Scholar]
  45. Gettig JP, Grady SE, and Nowosadzka I (2006). Methamphetamine: Putting the brakes on speed. J. Sch. Nurs 22, 66–73. [PubMed] [Google Scholar]
  46. Gilcrease MZ (2007). Integrin signaling in epithelial cells. Cancer Lett 247, 1–25. [DOI] [PubMed] [Google Scholar]
  47. Gluck MR, Moy LY, Jayatilleke E, Hogan KA, Manzino L, and Sonsalla PK (2001). Parallel increases in lipid and protein oxidative markers in several mouse brain regions after methamphetamine treatment. J. Neurochem 79, 152–160. [DOI] [PubMed] [Google Scholar]
  48. Gonzalez R, Rippeth JD, Carey CL, Heaton RK, Moore DJ, Schweinsburg BC, Cherner M, and Grant I (2004). Neurocognitive performance of methamphetamine users discordant for history of marijuana exposure. Drug Alcohol Depend 76, 181–190. [DOI] [PubMed] [Google Scholar]
  49. Gonzalez-Scarano F, and Martin-Garcia J (2005). The neuropathogenesis of AIDS. Nat. Rev. Immunol 5, 69–81. [DOI] [PubMed] [Google Scholar]
  50. Gorlach A, Klappa P, and Kietzmann T (2006). The endoplasmic reticulum: Folding, calcium homeostasis, signaling, and redox control. Antioxid. Redox Signal 8, 1391–1418. [DOI] [PubMed] [Google Scholar]
  51. Harold C, Wallace T, Friedman R, Gudelsky G, and Yamamoto B (2000). Methamphetamine selectively alters brain glutathione. Eur. J. Pharmacol 400, 99–102. [DOI] [PubMed] [Google Scholar]
  52. Hirata H, Ladenheim B, Carlson E, Epstein C, and Cadet JL (1996). Autoradiographic evidence for methamphetamine-induced striatal dopaminergic loss in mouse brain: Attenuation in CuZn-superoxide dismutase transgenic mice. Brain Res 714, 95–103. [DOI] [PubMed] [Google Scholar]
  53. Hozumi H, Asanuma M, Miyazaki I, Fukuoka S, Kikkawa Y, Kimoto N, Kitamura Y, Sendo T, Kita T, and Gomita Y (2008). Protective effects of interferon-gamma against methamphetamine-induced neurotoxicity. Toxicol. Lett 177, 123–129. [DOI] [PubMed] [Google Scholar]
  54. Imam SZ, el-Yazal J, Newport GD, Itzhak Y, Cadet JL, Slikker W Jr., and Ali SF (2001a). Methamphetamine-induced dopaminergic neurotoxicity: Role of peroxynitrite and neuroprotective role of antioxidants and peroxynitrite decomposition catalysts. Ann. NY Acad. Sci 939, 366–380. [DOI] [PubMed] [Google Scholar]
  55. Imam SZ, Newport GD, Itzhak Y, Cadet JL, Islam F, Slikker W Jr., and Ali SF (2001b). Peroxynitrite plays a role in methamphetamine-induced dopaminergic neurotoxicity: Evidence from mice lacking neuronal nitric oxide synthase gene or overexpressing copper-zinc superoxide dismutase. J. Neurochem 76, 745–749. [DOI] [PubMed] [Google Scholar]
  56. Itzhak Y, and Ali SF (1996). The neuronal nitric oxide synthase inhibitor, 7-nitroindazole, protects against methamphetamine-induced neurotoxicity in vivo. J. Neurochem 67, 1770–1773. [DOI] [PubMed] [Google Scholar]
  57. Itzhak Y, and Ali SF (2006). Role of nitrergic system in behavioral and neurotoxic effects of amphetamine analogs. Pharmacol. Ther 109, 246–262. [DOI] [PubMed] [Google Scholar]
  58. Itzhak Y, Gandia C, Huang PL, and Ali SF (1998). Resistance of neuronal nitric oxide synthasedeficient mice to methamphetamine-induced dopaminergic neurotoxicity. J. Pharmacol. Exp. Ther 284, 1040–1047. [PubMed] [Google Scholar]
  59. Itzhak Y, Martin JL, and Ail SF (2000). nNOS inhibitors attenuate methamphetamine-induced dopaminergic neurotoxicity but not hyperthermia in mice. Neuroreport 11, 2943–2946. [DOI] [PubMed] [Google Scholar]
  60. Iversen L (2006). Neurotransmitter transporters and their impact on the development of psychopharmacology. Br. J. Pharmacol 147(Suppl. 1), S82–S88. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Jayanthi S, Ladenheim B, and Cadet JL (1998). Methamphetamine-induced changes in antioxidant enzymes and lipid peroxidation in copper/zinc-superoxide dismutase transgenic mice. Ann. NY Acad. Sci 844, 92–102. [PubMed] [Google Scholar]
  62. Jayanthi S, Ladenheim B, Andrews AM, and Cadet JL (1999). Overexpression of human copper/zinc superoxide dismutase in transgenic mice attenuates oxidative stress caused by methylenedioxymethamphetamine (Ecstasy). Neuroscience 91, 1379–1387. [DOI] [PubMed] [Google Scholar]
  63. Jayanthi S, Deng X, Bordelon M, McCoy MT, and Cadet JL (2001). Methamphetamine causes differential regulation of pro-death and anti-death Bcl-2 genes in the mouse neocortex. FASEB J 15, 1745–1752. [DOI] [PubMed] [Google Scholar]
  64. Jayanthi S, Deng X, Noailles PA, Ladenheim B, and Cadet JL (2004). Methamphetamine induces neuronal apoptosis via cross-talks between endoplasmic reticulum and mitochondria-dependent death cascades. FASEB J 18, 238–251. [DOI] [PubMed] [Google Scholar]
  65. Jayanthi S, Deng X, Ladenheim B, McCoy MT, Cluster A, Cai NS, and Cadet JL (2005). Calcineurin/NFAT-induced up-regulation of the Fas ligand/Fas death pathway is involved in methamphetamine-induced neuronal apoptosis. Proc. Natl. Acad. Sci. USA 102, 868–873. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Kawasaki T, Ishihara K, Ago Y, Nakamura S, Itoh S, Baba A, and Matsuda T (2006). Protective effect of the radical scavenger edaravone against methamphetamine-induced dopaminergic neurotoxicity in mouse striatum. Eur. J. Pharmacol 542, 92–99. [DOI] [PubMed] [Google Scholar]
  67. Kim YS, and Joh TH (2006). Microglia, major player in the brain inflammation: Their roles in the pathogenesis of Parkinson’s disease. Exp. Mol. Med 38, 333–347. [DOI] [PubMed] [Google Scholar]
  68. Kiyatkin EA, Brown PL, and Sharma HS (2007). Brain edema and breakdown of the blood-brain barrier during methamphetamine intoxication: Critical role of brain hyperthermia. Eur. J. Neurosci 26, 1242–1253. [DOI] [PubMed] [Google Scholar]
  69. Kondo T, Ito T, and Sugita Y (1994). Bromocriptine scavenges methamphetamine-induced hydroxyl radicals and attenuates dopamine depletion in mouse striatum. Ann. NY Acad. Sci 738, 222–229. [DOI] [PubMed] [Google Scholar]
  70. Krasnova IN, Ladenheim B, and Cadet JL (2005). Amphetamine induces apoptosis of medium spiny striatal projection neurons via the mitochondria-dependent pathway. FASEB J 19, 851–853. [DOI] [PubMed] [Google Scholar]
  71. Kroemer G, Galluzzi L, and Brenner C (2007). Mitochondrial membrane permeabilization in cell death. Physiol. Rev 87, 99–163. [DOI] [PubMed] [Google Scholar]
  72. Kuhn DM, Francescutti-Verbeem DM, and Thomas DM (2006). Dopamine quinones activate microglia and induce a neurotoxic gene expression profile: Relationship to methamphetamine-induced nerve ending damage. Ann. NY Acad. Sci 1074, 31–41. [DOI] [PubMed] [Google Scholar]
  73. Ladenheim B, Krasnova IN, Deng X, Oyler JM, Polettini A, Moran TH, Huestis MA, and Cadet JL (2000). Methamphetamine-induced neurotoxicity is attenuated in transgenic mice with a null mutation for interleukin-6. Mol. Pharmacol 58, 1247–1256. [DOI] [PubMed] [Google Scholar]
  74. LaVoie MJ, and Hastings TG (1999). Dopamine quinone formation and protein modification associated with the striatal neurotoxicity of methamphetamine: Evidence against a role for extracellular dopamine. J. Neurosci 19, 1484–1491. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. LaVoie MJ, Card JP, and Hastings TG (2004). Microglial activation precedes dopamine terminal pathology in methamphetamine-induced neurotoxicity. Exp. Neurol 187, 47–57. [DOI] [PubMed] [Google Scholar]
  76. Liu B, and Dluzen DE (2006). Effect of estrogen upon methamphetamine-induced neurotoxicity within the impaired nigrostriatal dopaminergic system. Synapse 60, 354–361. [DOI] [PubMed] [Google Scholar]
  77. Li Y, and Trush MA (1993). DNA damage resulting from the oxidation of hydroquinone by copper: Role for a Cu(II)/Cu(I) redox cycle and reactive oxygen generation. Carcinogenesis 14, 1303–1311. [DOI] [PubMed] [Google Scholar]
  78. Lynch J, and House MA (1992). Cardiovascular effects of methamphetamine. J. Cardiovasc. Nurs 6, 12–18. [PubMed] [Google Scholar]
  79. Mamounas LA, Blue ME, Siuciak JA, and Altar CA (1995). Brain-derived neurotrophic factor promotes the survival and sprouting of serotonergic axons in rat brain. J. Neurosci 15, 7929–7939. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Maragos WF, Jakel R, Chesnut D, Pocernich CB, Butterfield DA, St Clair D, and Cass WA (2000). Methamphetamine toxicity is attenuated in mice that overexpress human manganese superoxide dismutase. Brain Res 878, 218–222. [DOI] [PubMed] [Google Scholar]
  81. Marciniak SJ, and Ron D (2006). Endoplasmic reticulum stress signaling in disease. Physiol. Rev 86, 1133–1149. [DOI] [PubMed] [Google Scholar]
  82. Marek GJ, Vosmer G, and Seiden LS (1990). Dopamine uptake inhibitors block long-term neurotoxic effects of methamphetamine upon dopaminergic neurons. Brain Res 513, 274–279. [DOI] [PubMed] [Google Scholar]
  83. Mark KA, Soghomonian JJ, and Yamamoto BK (2004). High-dose methamphetamine acutely activates the striatonigral pathway to increase striatal glutamate and mediate long-term dopamine toxicity. J. Neurosci 24, 11449–11456. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Marshall JF, O’Dell SJ, and Weihmuller FB (1993). Dopamine-glutamate interactions in methamphetamine-induced neurotoxicity. J. Neural. Transm. Gen. Sect 91, 241–254. [DOI] [PubMed] [Google Scholar]
  85. Matsuzaki H, Namikawa K, Kiyama H, Mori N, and Sato K (2004). Brain-derived neurotrophic factor rescues neuronal death induced by methamphetamine. Biol. Psychiatry 55, 52–60. [DOI] [PubMed] [Google Scholar]
  86. Mattson MP, Goodman Y, Luo H, Fu W, and Furukawa K (1997). Activation of NF-kappaB protects hippocampal neurons against oxidative stress-induced apoptosis: Evidence for induction of manganese superoxide dismutase and suppression of peroxynitrite production and protein tyrosine nitration. J. Neurosci. Res 49, 681–697. [DOI] [PubMed] [Google Scholar]
  87. Melega WP, Lacan G, Desalles AA, and Phelps ME (2000). 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 35, 243–249. [DOI] [PubMed] [Google Scholar]
  88. Moszczynska A, Fitzmaurice P, Ang L, Kalasinsky KS, Schmunk GA, Peretti FJ, Aiken SS, Wickham DJ, and Kish SJ (2004). Why is parkinsonism not a feature of human methamphetamine users? Brain 127, 363–370. [DOI] [PubMed] [Google Scholar]
  89. Murray JB (1998). Psychophysiological aspects of amphetamine-methamphetamine abuse. J. Psychol 132, 227–237. [DOI] [PubMed] [Google Scholar]
  90. Nakagawa T, and Yuan J (2000). Cross-talk between two cysteine protease families. Activation of caspase-12 by calpain in apoptosis. J. Cell. Biol 150, 887–894. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Nakajima A, Yamada K, Nagai T, Uchiyama T, Miyamoto Y, Mamiya T, He J, Nitta A, Mizuno M, Tran MH, Seto A, Yoshimura M, et al. (2004). Role of tumor necrosis factor-alpha in methamphetamine-induced drug dependence and neurotoxicity. J. Neurosci 24, 2212–2225. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Nash JF Jr., Meltzer HY, and Gudelsky GA (1988). Elevation of serum prolactin and corticosterone concentrations in the rat after the administration of 3,4-methylenedioxymethamphetamine. J. Pharmacol. Exp. Ther 245, 873–879. [PubMed] [Google Scholar]
  93. O’Callaghan JP, and Sriram K (2005). Glial fibrillary acidic protein and related glial proteins as biomarkers of neurotoxicity. Expert Opin. Drug Saf 4, 433–442. [DOI] [PubMed] [Google Scholar]
  94. O’Callaghan JP, Jensen KF, and Miller DB (1995). Quantitative aspects of drug and toxicant-induced astrogliosis. Neurochem. Int 26, 115–124. [DOI] [PubMed] [Google Scholar]
  95. O’Dell SJ, and Marshall JF (2000). Repeated administration of methamphetamine damages cells in the somatosensory cortex: Overlap with cytochrome oxidase-rich barrels. Synapse 37, 32–37. [DOI] [PubMed] [Google Scholar]
  96. O’Dell SJ, Weihmuller FB, and Marshall JF (1993). Methamphetamine-induced dopamine overflow and injury to striatal dopamine terminals: Attenuation by dopamine D1 or D2 antagonists. J. Neurochem 60, 1792–1799. [DOI] [PubMed] [Google Scholar]
  97. Pacher P, Beckman JS, and Liaudet L (2007). Nitric oxide and peroxynitrite in health and disease. Physiol. Rev 87, 315–424. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Perez JA Jr., Arsura EL, and Strategos S (1999). Methamphetamine-related stroke: Four cases. J. Emerg. Med 17, 469–471. [DOI] [PubMed] [Google Scholar]
  99. Perry VH, Cunningham C, and Holmes C (2007). Systemic infections and inflammation affect chronic neurodegeneration. Nat. Rev. Immunol 7, 161–167. [DOI] [PubMed] [Google Scholar]
  100. Potashkin JA, and Meredith GE (2006). The role of oxidative stress in the dysregulation of gene expression and protein metabolism in neurodegenerative disease. Antioxid. Redox Signal 8, 144–151. [DOI] [PubMed] [Google Scholar]
  101. Raivich G (2005). Like cops on the beat: The active role of resting microglia. Trends Neurosci 28, 571–573. [DOI] [PubMed] [Google Scholar]
  102. Rawson RA, and Condon TP (2007). Why do we need an Addiction supplement focused on methamphetamine? Addiction 102(Suppl. 1), 1–4. [DOI] [PubMed] [Google Scholar]
  103. Ricaurte GA, Schuster CR, and Seiden LS (1980). Long-term effects of repeated methylamphetamine administration on dopamine and serotonin neurons in the rat brain: A regional study. Brain Res 193, 153–163. [DOI] [PubMed] [Google Scholar]
  104. Rippeth JD, Heaton RK, Carey CL, Marcotte TD, Moore DJ, Gonzalez R, Wolfson T, and Grant I (2004). Methamphetamine dependence increases risk of neuropsychological impairment in HIV infected persons. J. Int. Neuropsychol. Soc 10, 1–14. [DOI] [PubMed] [Google Scholar]
  105. Roehr B (2005). Half a million Americans use methamphetamine every week. BMJ 331, 476. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Romanelli F, and Smith KM (2006). Clinical effects and management of methamphetamine abuse. Pharmacotherapy 26, 1148–1156. [DOI] [PubMed] [Google Scholar]
  107. Salo R, Nordahl TE, Possin K, Leamon M, Gibson DR, Galloway GP, Flynn NM, Henik A, Pfefferbaum A, and Sullivan EV (2002). Preliminary evidence of reduced cognitive inhibition in methamphetamine-dependent individuals. Psychiatry Res 111, 65–74. [DOI] [PubMed] [Google Scholar]
  108. Samantaray S, Ray SK, Ali SF, and Banik NL (2006). Calpain activation in apoptosis of motoneurons in cell culture models of experimental parkinsonism. Ann. NY Acad. Sci 1074, 349–356. [DOI] [PubMed] [Google Scholar]
  109. Sanchez V, Zeini M, Camarero J, O’Shea E, Bosca L, Green AR, and Colado MI (2003). The nNOS inhibitor, AR-R17477AR, prevents the loss of NF68 immunoreactivity induced by methamphetamine in the mouse striatum. J. Neurochem 85, 515–524. [DOI] [PubMed] [Google Scholar]
  110. Sapp E, Kegel KB, Aronin N, Hashikawa T, Uchiyama Y, Tohyama K, Bhide PG, Vonsattel JP, and DiFiglia M (2001). Early and progressive accumulation of reactive microglia in the Huntington disease brain. J. Neuropathol. Exp. Neurol 60, 161–172. [DOI] [PubMed] [Google Scholar]
  111. Schmidt CJ, and Gibb JW (1985). Role of the dopamine uptake carrier in the neurochemical response to methamphetamine: Effects of amfonelic acid. Eur. J. Pharmacol 109, 73–80. [DOI] [PubMed] [Google Scholar]
  112. Sekine Y, Minabe Y, Ouchi Y, Takei N, Iyo M, Nakamura K, Suzuki K, Tsukada H, Okada H, Yoshikawa E, Futatsubashi M, and Mori N (2003). Association of dopamine transporter loss in the orbitofrontal and dorsolateral prefrontal cortices with methamphetamine-related psychiatric symptoms. Am. J. Psychiatry 160, 1699–1701. [DOI] [PubMed] [Google Scholar]
  113. Sekine Y, Ouchi Y, Takei N, Yoshikawa E, Nakamura K, Futatsubashi M, Okada H, Minabe Y, Suzuki K, Iwata Y, Tsuchiya KJ, Tsukada H, et al. (2006). Brain serotonin transporter density and aggression in abstinent methamphetamine abusers. Arch. Gen. Psychiatry 63, 90–100. [DOI] [PubMed] [Google Scholar]
  114. Sharma HS, and Ali SF (2006). Alterations in blood-brain barrier function by morphine and methamphetamine. Ann. NY Acad. Sci 1074, 198–224. [DOI] [PubMed] [Google Scholar]
  115. Sharma HS, Sjoquist PO, and Ali SF (2007). Drugs of abuse-induced hyperthermia, blood-brain barrier dysfunction and neurotoxicity: Neuroprotective effects of a new antioxidant compound H-290/51. Curr. Pharm. Des 13, 1903–1923. [DOI] [PubMed] [Google Scholar]
  116. Shoptaw S (2006). Methamphetamine use in urban gay and bisexual populations. Top. HIV Med 14, 84–87. [PubMed] [Google Scholar]
  117. Shy H, Malaiyandi L, and Timiras PS (2000). Protective action of 17beta-estradiol and tamoxifen on glutamate toxicity in glial cells. Int. J. Dev. Neurosci 18, 289–297. [DOI] [PubMed] [Google Scholar]
  118. Simon SL, Domier CP, Sim T, Richardson K, Rawson RA, and Ling W (2002). Cognitive performance of current methamphetamine and cocaine abusers. J. Addict. Dis 21, 61–74. [DOI] [PubMed] [Google Scholar]
  119. Simon SL, Dacey J, Glynn S, Rawson R, and Ling W (2004). The effect of relapse on cognition in abstinent methamphetamine abusers. J. Subst. Abuse. Treat 27, 59–66. [DOI] [PubMed] [Google Scholar]
  120. Sim T, Simon SL, Domier CP, Richardson K, Rawson RA, and Ling W (2002). Cognitive deficits among methamphetamine users with attention deficit hyperactivity disorder symptomatology. J. Addict. Dis 21, 75–89. [DOI] [PubMed] [Google Scholar]
  121. Smith LM, LaGasse LL, Derauf C, Grant P, Shah R, Arria A, Huestis M, Haning W, Strauss A, Grotta SD, Liu J, Lester BM, et al. (2006). The infant development, environment, and lifestyle study: Effects of prenatal methamphetamine exposure, polydrug exposure, and poverty on intrauterine growth. Pediatrics 118, 1149–1156. [DOI] [PubMed] [Google Scholar]
  122. Sonsalla PK, Gibb JW, and Hanson GR (1986). Roles of D1 and D2 dopamine receptor subtypes in mediating the methamphetamine-induced changes in monoamine systems. J. Pharmacol. Exp. Ther 238, 932–937. [PubMed] [Google Scholar]
  123. Sonsalla PK, Nicklas WJ, and Heikkila RE (1989). Role for excitatory amino acids in methamphetamine-induced nigrostriatal dopaminergic toxicity. Science 243, 398–400. [DOI] [PubMed] [Google Scholar]
  124. Sriram K, and O’Callaghan JP (2007). Divergent roles for tumor necrosis factor-alpha in the brain. J. Neuroimmune Pharmacol 2, 140–153. [DOI] [PubMed] [Google Scholar]
  125. Sriram K, Matheson JM, Benkovic SA, Miller DB, Luster MI, and O’Callaghan JP (2006a). Deficiency of TNF receptors suppresses microglial activation and alters the susceptibility of brain regions to MPTP-induced neurotoxicity: Role of TNF-alpha. FASEB J 20, 670–682. [DOI] [PubMed] [Google Scholar]
  126. Sriram K, Miller DB, and O’Callaghan JP (2006b). Minocycline attenuates microglial activation but fails to mitigate striatal dopaminergic neurotoxicity: Role of tumor necrosis factor-alpha. J. Neurochem 96, 706–718. [DOI] [PubMed] [Google Scholar]
  127. Staszewski RD, and Yamamoto BK (2006). Methamphetamine-induced spectrin proteolysis in the rat striatum. J. Neurochem 96, 1267–1276. [DOI] [PubMed] [Google Scholar]
  128. Stumm G, Schlegel J, Schafer T, Wurz C, Mennel HD, Krieg JC, and Vedder H (1999). Amphetamines induce apoptosis and regulation of bcl-x splice variants in neocortical neurons. FASEB J 13, 1065–1072. [DOI] [PubMed] [Google Scholar]
  129. Sulzer D, Sonders MS, Poulsen NW, and Galli A (2005). Mechanisms of neurotransmitter release by amphetamines: A review. Prog. Neurobiol 75, 406–433. [DOI] [PubMed] [Google Scholar]
  130. Suuronen T, Nuutinen T, Huuskonen J, Ojala J, Thornell A, and Salminen A (2005). Anti-inflammatory effect of selective estrogen receptor modulators (SERMs) in microglial cells. Inflamm. Res 54, 194–203. [DOI] [PubMed] [Google Scholar]
  131. Thomas DM, and Kuhn DM (2005). Attenuated microglial activation mediates tolerance to the neurotoxic effects of methamphetamine. J. Neurochem 92, 790–797. [DOI] [PubMed] [Google Scholar]
  132. Thomas DM, Dowgiert J, Geddes TJ, Francescutti-Verbeem D, Liu X, and Kuhn DM (2004a). Microglial activation is a pharmacologically specific marker for the neurotoxic amphetamines. Neurosci. Lett 367, 349–354. [DOI] [PubMed] [Google Scholar]
  133. Thomas DM, Walker PD, Benjamins JA, Geddes TJ, and Kuhn DM (2004b). Methamphetamine neurotoxicity in dopamine nerve endings of the striatum is associated with microglial activation. J. Pharmacol. Exp. Ther 311, 1–7. [DOI] [PubMed] [Google Scholar]
  134. Topolski JM (2007). Epidemiology of methamphetamine abuse in Missouri. Mol. Med 104, 82–88. [PubMed] [Google Scholar]
  135. Volkow ND, Chang L, Wang GJ, Fowler JS, Leonido-Yee M, Franceschi D, Sedler MJ, Gatley SJ, Hitzemann R, Ding YS, Logan J, Wong C, et al. (2001). Association of dopamine transporter reduction with psychomotor impairment in methamphetamine abusers. Am. J. Psychiatry 158, 377–382. [DOI] [PubMed] [Google Scholar]
  136. Wagner GC, Ricaurte GA, Seiden LS, Schuster CR, Miller RJ, and Westley J (1980). Long-lasting depletions of striatal dopamine and loss of dopamine uptake sites following repeated administration of methamphetamine. Brain Res 181, 151–160. [DOI] [PubMed] [Google Scholar]
  137. Wagner GC, Carelli RM, and Jarvis MF (1985). Pretreatment with ascorbic acid attenuates the neurotoxic effects of methamphetamine in rats. Res. Commun. Chem. Pathol. Pharmacol 47, 221–228. [PubMed] [Google Scholar]
  138. Warren MW, Kobeissy FH, Liu MC, Hayes RL, Gold MS, and Wang KK (2005). Concurrent calpain and caspase-3 mediated proteolysis of alpha II-spectrin and tau in rat brain after methamphetamine exposure: A similar profile to traumatic brain injury. Life Sci 78, 301–309. [DOI] [PubMed] [Google Scholar]
  139. Warren MW, Zheng W, Kobeissy FH, Cheng Liu M, Hayes RL, Gold MS, Larner SF, and Wang KK (2007). Calpain- and caspase-mediated alphaII-spectrin and tau proteolysis in rat cerebrocortical neuronal cultures after ecstasy or methamphetamine exposure. Int. J. Neuropsychopharmacol 10, 479–489. [DOI] [PubMed] [Google Scholar]
  140. Wilson JM, Kalasinsky KS, Levey AI, Bergeron C, Reiber G, Anthony RM, Schmunk GA, Shannak K, Haycock JW, and Kish SJ (1996). Striatal dopamine nerve terminal markers in human, chronic methamphetamine users. Nat. Med 2, 699–703. [DOI] [PubMed] [Google Scholar]
  141. Woods SP, Rippeth JD, Conover E, Gongvatana A, Gonzalez R, Carey CL, Cherner M, Heaton RK, and Grant I (2005). Deficient strategic control of verbal encoding and retrieval in individuals with methamphetamine dependence. Neuropsychology 19, 35–43. [DOI] [PubMed] [Google Scholar]
  142. Xiang Z, Haroutunian V, Ho L, Purohit D, and Pasinetti GM (2006). Microglia activation in the brain as inflammatory biomarker of Alzheimer’s disease neuropathology and clinical dementia. Dis. Markers 22, 95–102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. Xu W, Zhu JP, and Angulo JA (2005). Induction of striatal pre- and postsynaptic damage by methamphetamine requires the dopamine receptors. Synapse 58, 110–121. [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Zhou JL, Liang JH, Zheng JW, and Li CL (2004). Nerve growth factor protects R2 cells against neurotoxicity induced by methamphetamine. Toxicol. Lett 150, 221–227. [DOI] [PubMed] [Google Scholar]
  145. Zhu JP, Xu W, and Angulo JA (2006). Methamphetamine-induced cell death: Selective vulnerability in neuronal subpopulations of the striatum in mice. Neuroscience 140, 607–622. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES