The Wayback Machine - https://web.archive.org/web/20210308145834/https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4418472/

Warning: The NCBI web site requires JavaScript to function. more...

Logo of nihpaAbout Author manuscriptsSubmit a manuscriptHHS Public Access; Author Manuscript; Accepted for publication in peer reviewed journal;
Int Rev Neurobiol. Author manuscript; available in PMC 2015 May 4.
Published in final edited form as:
PMCID: PMC4418472
NIHMSID: NIHMS684623
PMID: 25175865

Neuroimmune Basis of Methamphetamine Toxicity

Jennifer M. Loftis*‡,1 and Aaron Janowsky*§

Abstract

Although it is not known which antigen-specific immune responses (or if antigen-specific immune responses) are relevant or required for methamphetamine's neurotoxic effects, it is apparent that methamphetamine exposure is associated with significant effects on adaptive and innate immunity. Alterations in lymphocyte activity and number, changes in cytokine signaling, impairments in phagocytic functions, and glial activation and gliosis have all been reported. These drug-induced changes in immune response, particularly within the CNS, are now thought to play a critical role in the addiction process for methamphetamine dependence as well as for other substance use disorders. In Section 2, methamphetamine's effects on glial cell (e.g., microglia and astrocytes) activity and inflammatory signaling cascades are summarized, including how alterations in immune cell function can induce the neurotoxic and addictive effects of methamphetamine. Section 2 also describes neurotransmitter involvement in the modulation of methamphetamine's inflammatory effects. Section 3 discusses the very recent use of pharmacological and genetic animal models which have helped elucidate the behavioral effects of methamphetamine's neurotoxic effects and the role of the immune system. Section 4 is focused on the effects of methamphetamine on blood–brain barrier integrity and associated immune consequences. Clinical considerations such as the combined effects of methamphetamine and HIV and/or HCV on brain structure and function are included in Section 4. Finally, in Section 5, immune-based treatment strategies are reviewed, with a focus on vaccine development, neuroimmune therapies, and other anti-inflammatory approaches.

1. INTRODUCTION

The toxic effects of methamphetamine have been recognized for decades. Only recently, however, the role of the immune system in methamphetamine’s neurotoxic effects has been examined in detail. A number of molecular and cellular mechanisms are triggered following exposure of cells or animals to methamphetamine, and the cascade of events from exposure to neurotoxicity involves cellular components from receptors to immune system activation and inflammation, to energy metabolism. The term “neurotoxicity” can be ambiguous due to the array of methods and perspectives that are used to address methamphetamine’s effects. Here, the term is used to describe a condition that follows exposure to methamphetamine, which initiates a cascade of events resulting in altered behavior or cellular function in the absence of drug. This distinguishes neurotoxicity from intoxication and effects that occur simultaneously with drug exposure. This chapter reviews some of the more recent findings related to the mechanisms involved in the neuroimmune basis for methamphetamine toxicity.

2. PART 1

2.1. Microglia and astrocytes

Methamphetamine affects glial cell (e.g., microglia and astrocytes) activity, and glial cell activity canmodulate the neurotoxic and addictive effects of methamphetamine. Microglia are the major antigen-presenting cells in brain and when activated, they secrete an array of signaling molecules (e.g., proinflammatory cytokines and chemokines) that can cause neuronal damage. Astrocytes are the most numerous and diverse glial cells in the CNS, with a variety of functions including, but not limited to, maintenance of brain homeostasis, storage and distribution of energy substrates, synaptogenesis, and brain defense. Like microglia, astrocytes can also secret a number of signaling molecules that play a putative role in methamphetamine-induced neurotoxicity, such as pro- and anti-inflammatory cytokines including interleukins (ILs), interferons (IFNs), and tumor necrosis factors (TNFs), as well as chemokines (e.g., macrophage inflammatory proteins and monocyte chemoattractant proteins) (Whitney, Eidem, Peng, Huang, & Zheng, 2009). Activation of microglia and astrocytes are normal compensatory reactions to brain injury, but excess neuroinflammation can lead to further brain damage. Indeed, repeated or neurotoxic (i.e., high dose) methamphetamine exposure induces alterations in glial cell functions that contribute to a complex cascade of events, leading to neuroinflammation, neuronal damage, and behavioral impairments (Fig. 7.1). Table 7.1 provides a review of studies that describe acute and chronic effects of methamphetamine exposure on peripheral and central immune function.

An external file that holds a picture, illustration, etc.
Object name is nihms684623f1.jpg

Neuroimmune mechanisms of methamphetamine-induced CNS toxicity. The simplified schematic of brain and periphery illustrates the blood–brain barrier (BBB), immune and neural cells, and pro- and anti-inflammatory factors before (panel on left) and after (panel on right) methamphetamine. Following exposure to methamphetamine disruption of the BBB is observed (reviewed in Kousik, Napier, & Carvey, 2012) as well as activation of brain microglia and astrocytes, infiltration of macrophages/monocytes, alterations in the expression and balance of proinflammatory factors (e.g., IL-1β), myelin damage, and other indicators of neurodegeneration. Although not depicted in this figure, methamphetamine-induced neurotoxicity is accompanied not only by changes in immunity but also by alterations in neurotransmitter expression and function (such as excitotoxicity from glutamatergic overactivation) and changes in cellular bioenergetics (including oxidative and nitrosative stress). Collectively, these pathological processes contribute to neurotoxicity (e.g., increased BBB permeability, inflammation, neuronal degeneration, cell death) and neuropsychiatric impairments (e.g., cognitive deficits, mood disorders)—processes that can perpetuate addictive behaviors and hinder recovery efforts.

Table 7.1

Effects of methamphetamine exposure on immunologic outcomes in human and animal models Immune effects measured

ResultsReferences
Splenocyte immune factor production, following in vivo methamphetamine exposure and in vitro antigen stimulationDecreased ConA and LPS-induced production of IL-2 and IFN-γ; normal production of IL-4 and IL-6; increased production of TNF-αYu et al. (2002)
Peritoneal macrophage function, following in vivo methamphetamine exposure and in vitro antigen exposureReduced macrophage counts and reduced phagocytosis. Following exposure to LPS, reduced cytotoxic activity against melanoma cells; reduced antiviral activity against poly I:C; reduced NO2 production; reduced TNF-α, IL-1, and IL-6 production; reduced expression of CD14 receptorsIn, Son, Rhee, and Pyo (2004)
Lymphocyte proliferation and NK cell activity, following in vivo methamphetamine exposure and in vitro antigen exposureIn monkeys, following ConA and PHA exposure, increased lymphocyte proliferation and NK cell activitySaito et al. (2006)
Dendritic cell immune factor production, following in vitro methamphetamine exposure and in vitro stimulationWithin a large gene microarray, altered IL-4/GMCSF induced production of functional classes of genes involved in chemokine and cytokine regulation, signal transduction mechanisms, apoptosis, and cell cycle regulationMahajan et al. (2006)
Splenocyte T-cell proliferation and macrophage and dendritic cell function, following in vitro methamphetamine exposure and in vitro antigen exposureFollowing OVA exposure, reduced spleen T-cell proliferation; following exposure to Candida albicans and Cryptococcus neoformans, reduced macrophage phagocytosis, and increased fungal replication; within macrophages, collapsed pH gradients and increased accumulation of autophagosomesTalloczy et al. (2008)
Splenocyte antibody and immune factor production, following in vivo methamphetamine exposure and in vivo antigen sensitizationFollowing sensitzation to OVA, reduced splenocyte production of OVA-specific IgM, IgG1, and IgG2a; reduced IL-4 and IFN-γ production upon ex vivo restimulation with OVAWey, Wu, Chang, and Jan (2008)
Dendritic cell immune factor production, following in vitro methamphetamine exposure and in vitro stimulationWithin a large protein array, altered IL-4/GMCSF induced production of a number of functional classes of proteins that modulate apoptosis, protein folding, protein kinase activity, metabolism, and intracellular signal transductionReynolds et al. (2009)
Splenocyte T-cell proliferation, and peritoneal macrophage function, following in vivo methamphetamine exposure and in vitro antigen exposure/Disease progression, lung immune factor expression, and antibody production, following in vivo methamphetamine exposure and in vivo antigen exposureFollowing exposure to histoplasmosis, reduced splenocyte T-cell proliferation and reduced macrophage function (e.g., reduced phagocytosis), resulting in increased fungal load and infection/Following exposure to histoplasmosis, increased disease progression (e.g., to death); increased lung expression of TNF-α, IFN-γ, IL-4, IL-10, and TGF-β; increased IgG2b levelsMartinez, Mihu, Gacser, Santambrogio, and Nosanchuk (2009)
Neuroinflammatory responses to a subsequent peripheral immune stimulus in mice administered a neurotoxic methamphetamine treatment regimenMethamphetamine exacerbated the LPS-induced increase in central cytokine mRNA. Methamphetamine alone increased microglial Iba1 expression (a marker for microglial activation) and expression was further increased when mice were exposed to both methamphetamine and LPSBuchanan, Sparkman, and Johnson (2010a, 2010b)
Peripheral blood T-cell proliferation, following in vitro methamphetamine exposure and in vitro T-cell stimulationReduced anti-CD3/CD28 induced T-cell proliferation and IL-2 productionPotula et al., (2010)
Peripheral and central immune factor levels, following in vivo methamphetamine exposure in humans and mice during different periods of withdrawal/remiss ionA number of significant methamphetamine-induced changes in cytokines, chemokines, and adhesion factors were observed. Of particular interest were monocyte chemoattractant protein 1 (MCP-1; a.k.a., CCL2) and intercellular adhesion molecule (ICAM-1; a.k.a. CD54), which were similarly increased in the plasma of methamphetamine exposed mice as well as humans. In human participants, methamphetamine-induced changes in the cytokine and chemokine milieu were accompanied by increased cognitive impairmentsLoftis, Choi, Hoffman, and Huckans (2011)
Phenotypic changes in leukocyte subsets (e.g., NK cells, T cells, macrophages) from spleen and lymph node of mice treated with an escalating methamphetamine dose scheduleMethamphetamine exposure was associated with an overall decrease in both proportion and number of splenic NK cells, dendritic cells, and a subset of monocytes/macro phages [i.e., monocytes expressing high levels of granulocyte antigen 1 (Gr-1high)]Harms, Morsey, Boyer, Fox, and Sarvetnick (2012)
Astrocyte immune factor production, following in vitro methamphetamine exposureIn human fetal astrocytes, methamphetamine increased IL-6 and IL-8 production (blocked by a metabotropic glutamate receptor-5 inhibitor)Shah, Silverstein, Singh, and Kumar (2012)

CCR5, C-C chemokine receptor type 5; ConA, concanavalin A; CXCR4, C-X-C chemokine receptor 4; DA, dopamine; GMCSF, granulocyte–macrophage colonly stimulating factorl; HCV, hepatitis C virus; HIV, human immunodeficiency virus; ICAM, intracellular adhesion molecule; IFN, interferon; Ig, immunoglobulin; IL, interleukin; LPS, lipopolysaccharide; MAPK, mitogen-activated protein kinase; MCP, monocyte chemotactic protein; MIP, macrophage inflammatory protein; mRNA, messenger ribonucleic acid; NK, natural killer; OVA, ovalbumin; PHA, phytochemagglutinin-P; STAT1, signal transducer and activator of transcription1; TGF, transforming growth factor; TNF, tumor necrosis factor.

2.1.1 Microglia

The role of microglia in mediating methamphetamine’s effects has been appreciated since the late 1990s when it was reported that following methamphetamine exposure microglia participate more in necrosis–phagocytosis of neuronal debris than in restoration of neuronal processes (Escubedo et al., 1998). Subsequently, glial cell activation associated with methamphetamine exposure has been well documented with significant contributions made from a number of laboratories (e.g., Asanuma, Miyazaki, Higashi, Tsuji, & Ogawa, 2004; Fantegrossi et al., 2008; Guilarte, Nihei, McGlothan, & Howard, 2003; Hebert & O’Callaghan, 2000; Thomas, Francescutti-Verbeem, Liu, & Kuhn, 2004; Thomas, Walker, Benjamins, Geddes, & Kuhn, 2004). Microglial activation appears to be an early step in the process of methamphetamine-induced neurotoxicity, especially following exposure to high doses of methamphetamine (Thomas, Francescutti-Verbeem, & Kuhn, 2008b; Thomas, Francescutti-Verbeem, et al., 2004; Thomas, Walker, et al., 2004). Although, in some brain regions (e.g., caudate–putamen) microglial activation may not occur until 1 day or more after exposure to high-dose methamphetamine (Bowyer, Robinson, Ali, & Schmued, 2008). Neurotoxic regimens of methamphetamine exposure result in reactive microglia and increased inflammatory gene expression, particularly in striatum—but evidence of methamphetamine-induced microgliosis has also been observed in other brain regions [e.g., cerebellum and hippocampus (Escubedo et al., 1998); somatosensory and piriform cortices and periaqueductal gray (LaVoie, Card, & Hastings, 2004)].

Methamphetamine activates microglia in a dose-dependent manner and via a time course that is generally concurrent with damage to the dopaminergic system (Thomas, Walker, et al., 2004). As a possible target for methamphetamine’s effects on microglia, the microglial-specific fractalkine receptor (CX3CR1) [a mediator of 1-methyl-4-phenyl-1,2,3,4-tetrahydropyridine (MPTP)-induced neurodegeneration of dopamine neurons (Cardona et al., 2006)] was evaluated for a role in methamphetamine-induced microglial activation and neurotoxicity. Using mice in which the CX3CR1 gene was deleted (CX3CR1 knock-out mice), Thomas, Francescutti-Verbeem, and Kuhn (2008a) determined that CX3CR1 signaling does not modulate methamphetamine-induced neurotoxicity or microglial activation. Specifically, methamphetamine exposure had similar effects in both the CX3CR1 knock-out mice and in the wild-type control mice (e.g., microglial activation, increases in body temperature, and reductions in dopamine) (Thomas et al., 2008a).

Once activated, microglia contribute to and potentially perpetuate methamphetamine-induced neuroinflammation and neurodegeneration through inflammatory processes, including the production of proinflammatory cytokines (e.g., TNF-α, IL-1β, and IL-6), or through oxidative mechanisms (Clark, Wiley, & Bradberry, 2013; Yamamoto & Raudensky, 2008) (Fig. 7.1). For example, the excess dopamine resulting from methamphetamine exposure produces dopamine quinones (DAQs) which can activate microglia. Kuhn, Francescutti-Verbeem, and Thomas (2006) demonstrated that DAQs cause time-dependent activation of cultured microglial cells. Importantly, microarray analysis of the effects of DAQs on microglial gene expression indicated that many of the genes differentially regulated by DAQs were those associated with inflammation and neurotoxicity, including cytokines, chemokines, and prostaglandins. Thus, following methamphetamine exposure, the generation of DAQs may induce early activation of microglial cells and increased expression of inflammatory signaling cascades. Of note, one study reported a global pattern of microglial activation and microgliosis in individuals with a history of methamphetamine addiction, which appeared to persist for at least 2 years into abstinence (Sekine et al., 2008).

2.1.2 Astrocytes

For astrocytes, methamphetamine’s effects are mediated, in part, by changes in: (1) transcription factor pathways, (2) astrocytic cytokine receptors, (3) excitatory amino acid transporters (EAATs), and (4) glucose uptake mechanisms (Abdul Muneer, Alikunju, Szlachetka, & Haorah, 2011). Methamphetamine can activate astrocytes and induce astrogliosis (e.g., in striatum) via activation of the Janus kinase 2 (JAK2)/signal transducer and activator of transcription 3 (STAT3) signaling cascade (Hebert & O’Callaghan, 2000; Robson et al., 2014)—a pathway that is similarly thought to contribute to astrogliosis following exposure to other neurotoxic substances (MPTP) (e.g., Sriram, Benkovic, Hebert, Miller, & O’Callaghan, 2004) and one that may promote the persistence of reactive gliosis following toxicant exposure (Hebert & O’Callaghan, 2000). For example, Friend and Keefe (2013) reported that astrocytes (but not microglia) remain reactive for at least 30 days following methamphetamine exposure. Consistent with a role for inflammatory signaling in maintaining methamphetamine’s activation of astrocytes, mice treated with a neurotoxic regimen of methamphetamine (i.e., four doses of 5 mg/kg administered every 2 h) show significant increases in oncostatin M receptor expression (an astrocytic receptor, Tamura, Morikawa, & Senba, 2003, for the proinflammatory IL-6-type cytokine oncostatin M) (Robson et al., 2014). Methamphetamine also alters EAATs on astrocytes [particularly type 2, EAAT-2 (reviewed in Cisneros & Ghorpade, 2012)] putatively via the production of reactive oxygen species (ROS) (Vaarmann, Gandhi, & Abramov, 2010). These changes in transporter expression could compromise the ability of EAATs to remove extracellular glutamate from the synapse (Cisneros & Ghorpade, 2012; Lau & Tymianski, 2010), thus contributing to methamphetamine-induced excitotoxicity and neurotoxic immune activation [a.k.a. inflammotoxicity (Wilhelm, Hashimoto, Roberts, & Loftis, 2014)].

2.2. Inflammatory signaling

A growing literature demonstrates that methamphetamine exposure and activation of microglia and astrocytes (as well as other cell types) alters peripheral and central immune functions (In, Son, Rhee, & Pyo, 2005; In et al., 2004; Liang et al., 2008; Martinez et al., 2009; Ye et al., 2008) and that immune factors such as cytokines (e.g., IL-1β, IL-6, and TNF-α), chemokines [e.g., monocyte chemotactic protein-1 (MCP-1)], and adhesion molecules play a critical role in the development and persistence of methamphetamine-induced neuronal injury and neuropsychiatric impairments (Clark et al., 2013; Loftis et al., 2011; Yamamoto, Moszczynska, & Gudelsky, 2010; Yamamoto & Raudensky, 2008). In an astrocytic cell line, methamphetamine exposure for 3 days increases IL-6 and IL-8 RNA levels by 4.6 ± 0.2 fold and 3.5 ± 0.2 fold, respectively (Shah et al., 2012). Consistent with these findings, mice treated with methamphetamine show brain region-specific increases in the expression of proinflammatory cytokines (e.g., IL-1β) up to 3 weeks after methamphetamine exposure (Loftis et al., 2011). In a rigorous study designed to evaluate whether or not deletion of the TNF-α gene in mice affects addictive-like behavior, including methamphetamine self-administration, motivation to self-administer methamphetamine, and cue-induced reinstatement of methamphetamine-seeking behavior, Yan, Nitta, Koseki, Yamada, and Nabeshima (2012) measured methamphetamine self-administration and reinstatement of drug-seeking behavior in TNF-α knock-out and wild-type mice. The authors observed an upward shift of dose responses to methamphetamine self-administration under a fixed ratio schedule of reinforcement in the TNF-α knockout as compared with wild-type mice, indicating that mice lacking TNF-α administered more methamphetamine than controls. Similarly, TNF-α knock-out mice also had a higher breaking point under a progressive ratio schedule of reinforcement, suggesting that TNF-α can also influence motivation to self-administer drug. Taken together, these findings demonstrate that TNF-α signaling affects methamphetamine self-administration and motivation to obtain the drug. However, TNF-α did not appear to contribute to methamphetamine-associated cue-induced relapsing behavior (Yan et al., 2012).

Interestingly, IFN-γ (another proinflammatory cytokine) injected systemically prior to repeated methamphetamine exposure protects against methamphetamine-induced neurotoxicity (as measured by a reduction in striatal dopamine transporters) and hyperthermia, putatively through intracerebralmolecular pathways (Hozumi et al., 2008). Thus, of the altered immune factors, some cytokines appear to facilitate or promote methamphetamine-induced toxicities but others may slow or prevent the development of adverse drug effects. Taken together, these observations describe a critical role for CNS immune signaling in methamphetamine neurotoxicity and dependence. More research is needed to elucidate the specific signaling pathways [e.g., STAT3, nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) (Wires et al., 2012)] and cell types (e.g., microglia, astrocytes, and vascular endothelial cells) that induce and perpetuate these neuroinflammatory processes—particularly because these immune effects can impact addictive and related behavioral responses.

2.3. Molecular/cellular: receptors and transporters

2.3.1 Neurotransmitter modulation of methamphetamine's inflammatory effects

Methamphetamine alters the disposition of a number of neurotransmitters and high doses cause reductions in a number of cell types, especially in the striatum. Loss of dopaminergic, GABAergic, and cholinergic cells has been described (Zhu, Xu, & Angulo, 2006). However, glial cells are also involved in neurotoxic reactions to methamphetamine, and receptors for neurotransmitters have been found on various cell types that are involved in immune response. Methamphetamine-induced alterations in neurotransmitter availability can affect receptor-mediated second messenger systems and downstream events inside the cell. Recent evidence for possible roles of methamphetamine in altering neurotransmitter-linked changes in intermediates of the immune response is described here.

As opposed to its indirect effects on neurotransmitter receptors via its actions as a serotonin, norepinephrine, dopamine, and vesicular monoamine (2) transporter substrate (Eshleman, Calligaro, & Eldefrawi, 1993; Sulzer et al., 1995; Wilhelm, Johnson, Lysko, Eshleman, & Janowsky, 2004), methamphetamine and its analogs, including methylenedioxymethamphetamine (MDMA), are direct agonists at trace amine associated receptor 1 (TAAR1). Endogenous agonists for the TAAR1 include tyramine and octopamine, as well as other neurotransmitter precursors and metabolites (Borowsky et al., 2001; Bunzow et al., 2001). TAAR1 mRNA and mRNA for other TAARs have been identified on mouse B and NK cells, and in stimulated human peripheral blood lymphocytes (Nelson, Tolbert, Singh, & Bost, 2007). In addition, TAAR1 protein has been characterized in rhesus monkey peripheral blood mononuclear cells (Panas et al., 2012). This latter report also characterized the effects of methamphetamine on phosphorylation of protein kinase A (PKA) and protein kinase C (PKC) activity in peripheral blood mononuclear cells as well as in rhesus monkey immortalized B lymphocytes. Generally, methamphetamine increased expression of the phosphorylated enzymes in PHA-treated cells. TAAR1-mediated stimulation of PKA and PKC phosphorylation appears to be necessary but not sufficient, i.e., PHA is also required for phosphorylation. In fact, it appears that PHA is required to increase the expression of TAAR1, which precedes methamphetamine stimulation of TAAR1 and subsequent phosphorylation of PKA and PKC. The authors also reported methamphetamine-induced increases in nuclear factor of activated T cells, part of the calcium response pathway, and an increase in cyclic AMP-response element binding protein (CREB) pathway in rhesus monkey TAAR1-transfected cells. Thus, methamphetamine stimulation of TAAR1 in lymphocytes appears to play an important role in lymphocyte function preceding immune response. Babusyte, Kotthoff, Fiedler, and Krautwurst (2013) subsequently demonstrated that activation of T-cell and B-cell TAAR1 (and 2) by endogenous ligands such as phenethylamine results in altered cytokine expression. Although the effects of methamphetamine were not examined, similar results could be expected. Reports also indicate that malignant B cells, including a Burkitt’s lymphoma cell line are sensitive to TAAR1 agonists, which reduce their viability (Wasik, Millan, Scanlan, & Barnes, 2012). Clearly, TAAR1 receptors on lymphocytes are involved in cell function and cellular response. However, it is not clear that TAAR1-mediated cellular mechanisms in neurons and lymphocytes are shared.

Glutamate receptors have also been implicated in the behavioral effects of methamphetamine (Achat-Mendes, Platt, & Spealman, 2012; Crawford, Roberts, & Beveridge, 2013; Herrold, Voigt, & Napier, 2013), and there are some data to indicate the role of specific glutamate receptors in neurotoxicity. Metabotropic glutamate receptor 5 (mGluR5) appears to play a role in the methamphetamine-induced change in immune response signaling. Shah et al. (2012) reported that exposure of the SVGA astrocyte cell line to methamphetamine caused a dose-dependent increase in mRNA and protein for the proinflammatory cytokines IL-6 and IL-8, as well as translocation of the nuclear subunit of NFκB, and the effects are most likely due to activation of Akt/PI3K pathways. Importantly, the effects of methamphetamine on IL-6 and IL-8 expression were reduced in the presence of 2-methyl-6-(phenylethynyl)pyridine (MPEP), an mGluR5 antagonist. The authors proposed a model consistent with direct stimulation of mGluR5 by methamphetamine. Additional studies suggest that mGluR5 regulates methamphetamine-induced dopamine release and the neurotoxic effects of methamphetamine on dopaminergic neurons, as MPEP also blocks the reduction in dopamine that is caused by high-dose methamphetamine, as well as methamphetamine-induced dopamine release (Gołembiowska, Konieczny, Wolfarth, & Ossowska, 2003).

The involvement of ionotropic N-methyl-d-aspartate (NMDA) and alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptor subtypes of glutamate receptors have also been implicated in methamphetamine-induced neurotoxicity by experimental results indicating that the NMDA receptor antagonist (dl)-amino-5-phosphonovaleric acid (AP5) or the AMPA receptor antagonist dinitroquinoxaline-2,3-dione (DNQX) prevent methamphetamine-induced c-fos expression and dopamine depletion, possibly as a result of reducing corticostriatal input (Gross, Duncker, & Marshall, 2011). Calcium release has been implicated in the mechanism of action of glutamate and methamphetamine neurotoxicity: NMDA + methamphetamine-induced neurotoxicity could be reduced by exposure of hippocampal slices in culture to the endoplasmic reticulum-bound sigma-1 receptor antagonist BD1047 which is coupled to calcium-mediated signal cascades (Smith, Butler, & Prendergast, 2010). Thus, evidence suggests both direct and indirect roles for glutamate receptors in methamphetamine-induced neurotoxic events.

A number of reports implicate the α7 subtype of nicotinic cholinergic receptor in the neurotoxic effects of methamphetamine. In the rat, 6-hydroxydopamine model of Parkinson’s disease, the α7 receptor agonist, 3-[(2,4-dimethoxy)benzylidene]-anabaseine dihydrochloride, reduces 6-hydroxydopamine-induced cell loss and the elevation of markers of neurotoxicity (Suzuki et al., 2013). The possibility that α7 receptors play a similar role in methamphetamine-induced neurotoxicity is supported by evidence that the α7 antagonist methyllycaconitine (MLA) prevents methamphetamine-induced neurotoxic events including decreases in cellular dopamine and in dopamine transporter expression, and methamphetamine-induced glutamate release (Northrop, Smith, Yamamoto, & Eyerman, 2011). The effect may be mediated by ROS, since MLA blocks methamphetamine-induced ROS production (Escubedo, Camarasa, Chipana, García-Ratés, & Pubill, 2009; Escubedo, Chipana, Pérez-Sánchez, Camarasa, & Pubill, 2005; Pubill et al., 2005). Therefore, the data support the hypothesis that methamphetamine indirectly causes release of the glutamate that is involved in methamphetamine-induced neurotoxicity. Although methamphetamine affects a number of receptors, including specific nicotinic cholinergic receptor subtypes, in its toxic effects, the role of muscarinic cholinergic receptors in methamphetamine’s cellular effects has not been described in detail.

Methamphetamine is a substrate for the dopamine transporter and causes neurotransmitter release (Eshleman et al., 2013), which may trigger drug-induced changes in behavior, but may not exert its neurotoxic effects via direct interaction with dopamine receptors. However, chronic changes in neurotransmitter interactions with receptors that accompany drug abuse can alter cellular function, and lymphocytes express a number of receptors that might be affected by methamphetamine (see also Table 7.1). In Section 3.1, effects of methamphetamine on signaling systems in brain are discussed.

3. PART 2

3.1. Animal models

Behavioral effects of methamphetamine’s neurotoxic effects and the role of the immune system have been characterized using pharmacological and genetic models. Many of the genetic models involve knock-out animals. For instance, dopamine D1 knock-out mice have less hyperthermia, gliosis, and inducible nitric oxide synthase in the striatum and less dopaminergic cell loss in the substantia nigra in response methamphetamine, as compared to wild-type control mice (Ares-Santos et al., 2012), and dopamine D2 receptor knock-out mice have reduced hyperthermia, loss of striatal dopaminergic markers, as well as a reduction in other markers of neurotoxicity (Granado, Ares-Santos, et al., 2011). Nrf2 is a neuroprotective gene that attenuates the effects of oxidative stress and is involved in innate immune response in the CNS. Nrf2 knock-out mice have an exaggerated response to methamphetamine, including increased glial activation, and expression of TNF-α and IL-15 (Granado, Lastres-Becker, et al., 2011). Prodynorphin-deficient mice were less sensitive to the dopamine depleting effects of methamphetamine as compared to wild-type mice. The mechanism for this effect may be via anti-inflammatory effects of dynorphin, which is found in very high concentrations in the nigrostriatal pathway (Wang et al., 2012). Likewise, protein kinase Cδ knock-out animals have reduced lipid peroxidation, protein oxidation, and behavioral deficits in response to methamphetamine. These studies were corroborated using pharmacological interventions, which also indicate increased tyrosine hydroxylase phosphorylation in response to methamphetamine (Shin et al., 2011, 2012).

Using pharmacological manipulations to reduce serotonin, Thomas, Angoa Pérez, Francescutti-Verbeem, Shah, and Kuhn (2010) demonstrated that serotonin may not play an essential role in methamphetamine neurotoxicity (reductions in dopaminergic cell markers such as dopamine, tyrosine hydroxylase, and the dopamine transporter). Importantly, this study also demonstrated that hyperthermia is not a prerequisite for toxicity, since depletion of serotonin is associated with increased hyperthermia. In experiments connecting the pharmacological effects of methamphetamine to its immunological effects, Buchanan et al. (2010b) demonstrated that a high-dose methamphetamine regimen increased the CNS cytokine response to subsequent i.p. administration of LPS, and also activated microglia.

4. PART 3

4.1. Blood–brain barrier and clinical considerations

4.1.1 Blood–brain barrier

The blood–brain barrier (BBB) is comprised of endothelial cells that line brain capillaries and contribute to supplying the brain with nutrients, regulating the transport of essential molecules (e.g., glucose and amino acids) to the brain, and protecting neural tissue from toxic substances—functions that collectively serve to maintain a stable environment. The BBB endothelial cells are surrounded by astrocytic perivascular endfeet which release specific factors, including transforming growth factor-β (TGF-β) and glial-derived neurotrophic factor (GDNF), important for the development and function of the BBB. Continuous tight junctions between the BBB endothelial cells as well as metabolic barriers (e.g., enzymes, transport systems) regulate transport across the BBB (e.g., limit the diffusion of molecules across the BBB and control the entry of neurotransmitters into the brain) (Abbott, Rönnbäck, & Hansson, 2006). Disturbances of brain homeostasis as a result of BBB damage are evident in a number of pathological conditions such as stroke, multiple sclerosis, Alzheimer’s disease, HIV, and substance use disorders (including methamphetamine dependence). Studies show that although the initial etiologies for BBB disruption are different among these pathologies, characteristic alterations and common intracellular pathways are involved. NFκB activation and proinflammatory signaling [e.g., via IL-6 and matrix metalloproteinases] often mediate dysfunction of the BBB’s structural and functional integrity (Abbott et al., 2006; Persidsky, Ramirez, Haorah, & Kanmogne, 2006). For example, proinflammatory cytokine-mediated release of matrix metalloproteinases (e.g., MMP-9) can increase BBB permeability by activating NF-κB, regulating microfilament reorganization and reducing the expression of tight junction proteins (or redistributing tight junction proteins from the membrane to the cytosol) (Aghajanian, Wittchen, Allingham, Garrett, & Burridge, 2008; Bogatcheva & Verin, 2008; Brown et al., 2003; Lu et al., 2009).

4.1.2 Methamphetamine exposure and BBB integrity

Psychostimulant exposure, and in particular methamphetamine abuse, is associated with disruption of the BBB (reviewed in Kousik et al., 2012) as well as with increased expression of proinflammatory factors. Pre-clinical in vivo studies and cultured human brain microvascular endothelial cell in vitro experiments show that methamphetamine contributes to inflammation-induced BBB dysfunction in a dose- and time-dependent manner by activating NF-κB, upregulating inflammatory gene expression (TNF-α), increasing the expression and activity of MMP-9, reducing the expression of tight junction proteins [e.g., zonula occludens (ZO)-1, claudin-5, and occludin], increasing the production of ROS (Lee, Hennig, Yao, & Toborek, 2001; Martins et al., 2011; Ramirez et al., 2009). Collectively, these adverse effects on BBB integrity contribute to neurodegeneration and may perpetuate toxic neural-immune interactions. For example, following exposure to high-dose methamphetamine, disruption of the BBB [as measured using immunoglobulin G (IgG) immunoreactivity to identify extravasated IgG in brain] is associated with neurodegeneration and activation of brain microglia and/or infiltration of macrophages. Methamphetamine-induced neurodegeneration is evident in brain regions with notable IgG immunoreactivity (i.e., hippocampus and amygdala) and during hyperthermic conditions (mouse body temperatures >40.5 °C). It was additionally reported that the hippocampal damage induced by methamphetamine and BBB disruption is likely sufficient to compromise cognitive function (Bowyer & Ali, 2006).

In a follow-up study, Bowyer et al. (2008) investigated BBB integrity to determine whether barrier disruption also plays a role in methamphetamine-induced neurotoxicity in the caudate–putamen—as it did in other brain regions (i.e., amygdala and hippocampus). Mice were administered high-dose methamphetamine and evaluated at 90 min, 4 h, 12 h, 1 and 3 days following drug exposure. The results showed that the BBB was only modestly altered (as indicated by IgG immunoreactivity) in the caudate–putamen at time points and subsequent to observation of neurodegeneration, and similar to previous work, these changes were dependent on hyperthermia (Bowyer et al., 2008; but see discussion of hyperthermia and serotonin involvement; Thomas et al., 2010).

Methamphetamine exposure, in combination with stress, can synergistically exacerbate BBB damage (i.e., reductions in occludin and claudin-5 immunoreactivity) via inflammatory mechanisms that persist for at least 7 days following methamphetamine (Northrop & Yamamoto, 2012). Collectively, these methamphetamine-induced effects on BBB contribute to immune dysfunction, such as increased leukocyte/monocyte transmigration across the endothelium and into the CNS (Buch et al., 2012; Ramirez et al., 2009, Fig. 7.1) as well as increased invasion of peripheral bacteria and viruses into the brain.

4.1.3 Clinical picture: immune signaling in addictive and related behaviors

In a classic positron emission tomography imaging study, Volkow et al. (2001) demonstrate that protracted abstinence can reverse some methamphetamine-induced brain insults (e.g., reductions in brain dopamine terminals) but suggest that other neurotoxic processes may persist and contribute to neuropsychiatric impairments during and following drug exposure. Research is beginning to link immune factor signaling with neural and behavioral aspects of addiction, such as impaired cognitive function (Loftis et al., 2011), drug-seeking behaviors, and resilience to relapse (Blednov, Mayfield, Belknap, & Harris, 2012; Blednov, Ponomarev, et al., 2012; Schwarz, Hutchinson, & Bilbo, 2011; Zhang et al., 2012). For example, in a pre-clinical study propentofylline (a phosphodiesterase inhibitor and glial cell modulator) significantly reduced the activation of astrocytes induced by methamphetamine, and treatment with propentofylline suppressed the rewarding effects of methamphetamine, as measured using conditioned place preference testing (Narita et al., 2006). These findings suggest that methamphetamine-induced astrocyte activation and the subsequent release of proinflammatory factors can play a role in the development of the rewarding effects of methamphetamine.

In addition to contributing to the addictive properties of methamphetamine, cytokines and other inflammatory factors can also be potent modulators of mood and cognitive function. This is important because neuropsychiatric impairments can persist following abstinence and are associated with poorer treatment outcomes, such as increased relapse rates, lower treatment retention rates, and reduced daily functioning (e.g., Zorick et al., 2010). Repeated methamphetamine exposure induces alterations in peripheral and central immune factor expression, and in clinical samples of methamphetamine-dependent humans, these peripheral alterations in immune factor expression are associated with cognitive impairments and mood disturbances (Letendre et al., 2005; Loftis et al., 2011). The neuroinflammatory effects of methamphetamine appear to be brain region specific and can lead to differential effects on cognitive function (Chang, Cloak, et al., 2005; Jernigan et al., 2005). Thus, interventions that target immune mechanisms for repairing persistent methamphetamine-induced CNS injury and neuropsychiatric impairments, both vital for successful recovery from methamphetamine dependence, may help reduce relapse rates and improve treatment outcomes in adults recovering from methamphetamine dependence and other substance use disorders (see Section 5.2).

4.1.4 Chronic infection with human immunodeficiency virus and/or hepatitis C virus

Exposure to methamphetamine and other substances of abuse can increase the invasion of peripheral bacteria and viruses, such as human immunodeficiency virus (HIV) and hepatitis C virus (HCV), into the brain (e.g., Gavrilin, Mathes, & Podell, 2002). Further, individuals who are infected with one virus have an increased risk of coinfection with the other. Given the high rates of comorbid viral infection with substance use disorders (Fuller, Loftis, Rodriguez, McQuesten, & Hauser, 2009; Loftis, Matthews, & Hauser, 2006), particularly psychostimulant dependence (e.g., Buchacz et al., 2005), the combined effects of substance use disorders and HIV and/or HCV on neural-immune interactions are of scientific and clinical interest (Cadet & Krasnova, 2007). In addition, the effects of chronic viral infection on brain are likely exacerbated by the adverse effects of comorbid methamphetamine dependence on BBB integrity and function (Liang et al., 2008) (Fig. 7.1).

HIV infection of glial cells results in their activation and release of cytokines, ROS, HIV-related proteins, and other neurotoxic mediators (Nath & Geiger, 1998). Reports indicate that HIV-induced neurotoxicity may preferentially affect brain regions rich with dopaminergic transporters (e.g., basal ganglia)—brain areas that also show damage following methamphetamine abuse (Nath, Maragos, Avison, Schmitt, & Berger, 2001). In fact, methamphetamine use has been linked to worse neuropsychiatric functioning in acute and early HIV infection, a critical period of immune activation (Weber et al., 2013).

One prominent neurotoxin associated with HIV is Tat (released from HIV-infected glia). Tat has been suggested to be pathogenically relevant in HIV-1-induced neuronal injury via diverse mechanisms. For example, coexposure to Tat and methamphetamine causes a synergistic loss of striatal dopamine and binding to the dopamine transporter, suggesting a loss of dopamine terminals (Theodore, Stolberg, Cass, & Maragos, 2006). In addition, methamphetamine and Tat can activate microglia and induce cytokine production. Interestingly, the interaction of Tat and methamphetamine is prevented in MCP-1-deficient mice and this also attenuates methamphetamine + Tat neurotoxicity (Theodore, Cass, & Maragos, 2006). Follow-up studies demonstrate that Tat contributes to increased TNF-α levels that exacerbates dopamine transporter damage from methamphetamine (Theodore, Cass, Nath, et al., 2006). Similarly, methamphetamine use in the presence of HIV produces additive interneuron loss (Chana et al., 2006) and abnormal brain metabolism in frontostriatal pathways (Chang, Ernst, Speck, & Grob, 2005).

The combination of HIV and methamphetamine is also associated with higher rates of neuropsychiatric impairment than either of the independent conditions (Rippeth et al., 2004). However, not all reports have been consistent regarding additive or synergistic effects of HIV and methamphetamine on brain structure and function (e.g., Archibald et al., 2012). The mechanisms of such injury are currently under investigation and inflammation, oxidative damage, and glutamate imbalance appear to contribute to neurotoxicity associated with methamphetamine dependence and comorbid HIV.

A few research groups are investigating CNS mechanisms associated with the neurotoxic effects of HCV and methamphetamine dependence. For example, using neuroimaging techniques, Taylor et al. (2004) found that N-acetylaspartate (NAA), a marker of neuronal integrity, was lower in the white matter of patients with HCV and a history of methamphetamine abuse, as compared to participants with HCV and no history of methamphetamine abuse. Importantly, this reduction in NAA was correlated with worse global neuropsychological deficit scores in the individuals with a history of methamphetamine addiction. However, at least one study of methamphetamine users (without HCV or HIV) found that impairments in cognitive function were unrelated to prior drug use histories (including self-reported age at first use, total years of use, route of consumption, or length of abstinence) (Cherner et al., 2010). Thus, more research is needed, as the additive or synergic effects of methamphetamine and viral infection on specific neurotoxicities and neuropsychiatric impairments are yet to be defined.

5. PART 4

5.1. Vaccine development

Anti-addiction or anti-drug vaccines are currently among the most developed immunotherapeutic approaches for substance use disorders [see National Institute on Drug Abuse (NIDA)-sponsored anti-drug vaccine animation at http://www.drugabuse.gov/news-events/nida-notes/antidrug-vaccine-animation]. These vaccines are designed to generate and attract antibodies to a substance so that it is too large to pass through the BBB (Cerny & Cerny, 2009; Gentry, Rüedi-Bettschen, & Owens, 2009; Kinsey, Jackson, & Orson, 2009; Li et al., 2011; Orson, Kinsey, Singh, Wu, & Kosten, 2009), thereby blocking its CNS actions and rewarding effects. To date, vaccines have been created against nicotine, morphine/heroin, cocaine, and methamphetamine, and a number of compounds are undergoing clinical trials or are in preclinical development (Goniewicz & Delijewski, 2013; Kosten, Domingo, Orson, & Kinsey, 2014; Shen et al., 2013). Researchers recently completed the first clinical study to establish the safety of a human-mouse chimeric monoclonal antibody (mAb) (i.e., ch-mAb7F9) designed to bind methamphetamine with high affinity and specificity in humans (Stevens et al., 2014). Results from the Phase 1a, double-blind, randomized, placebo-controlled clinical trial are not yet published (https://ClinicalTrials.gov Identifier: NCT01603147).

5.2. Neuroimmune-based and other anti-inflammatory treatment strategies

Ibudilast (AV-411 or MN-166) is a nonspecific phosphodiesterase inhibitor which suppresses glial cell activation and causes other anti-inflammatory effects. Prior to its development as a potential substance abuse treatment (e.g., for methamphetamine, alcohol, and opioid dependence), ibudilast had been used clinically for asthma, pulmonary, and cardiovascular diseases. Preclinically, administration of ibudilast following exposure to methamphetamine reduces: (1) the acute, chronic, and sensitization effects of the drug’s locomotor activity, (2) stress-induced methamphetamine relapse, and (3) methamphetamine self-administration—suggesting that glial cell activity can modulate methamphetamine’s behavioral effects (Beardsley, Shelton, Hendrick, & Johnson, 2010; Snider, Hendrick, & Beardsley, 2013; Snider et al., 2012). To date, a Phase I clinical trial of ibudilast with methamphetamine dependence has been completed. This initial human study was done in order to test the safety of the drug taken in combination with methamphetamine. Results from the Phase I safety interaction trial of ibudilast with methamphetamine are not yet publicly available (https://ClinicalTrials.gov Identifier: NCT01217970). Currently, the research is into its second phase, and results from the 12-week Phase II clinical trial are expected to be released in early 2015 (https://ClinicalTrials.gov Identifier: NCT01860807).

In addition to the development of therapeutic strategies for preventing and reducing methamphetamine use, interventions that target immune mechanisms for repairing persistent methamphetamine-induced CNS injury and neuropsychiatric impairments may also help reduce relapse rates and improve treatment outcomes in adults recovering from methamphetamine dependence and other substance use disorders. Given that a partial major histocompatibility complex (MHC)/neuroantigen peptide construct [pI-Ab/mMOG-35–55; a.k.a. recombinant T-cell receptor ligand (RTL)] effectively reduces the inflammatory and behavioral effects of experimental models of multiple sclerosis and stroke (Sinha et al., 2007; Subramanian et al., 2009; Vandenbark et al., 2003; Wang et al., 2006), it is possible that partial MHC/neuroantigen constructs could also effectively address the neuropsychiatric effects of chronic methamphetamine addiction. In a preclinical study, RTL-containing MHC coupled to myelin peptide [mouse myelin oligodendrocyte glycoprotein (mMOG)] improves the learning and memory impairments and CNS inflammation induced by repeated methamphetamine exposure in mouse models of chronic methamphetamine addiction (Loftis, Wilhelm, Vandenbark, & Huckans, 2013).

In addition to vaccine and neuroimmune-based strategies (e.g., ibudilast and partial MHC/neuroantigen peptide constructs), pharmacotherapies that regulate and reduce inflammation and oxidative stress are also under investigation. For example, minocycline (a broad-spectrum tetracycline antibiotic) has anti-inflammatory and antioxidant properties and has demonstrated some efficacy in the treatment of psychiatric disorders and neuropsychiatric symptoms. Preclinically, minocycline improves deficits in novel object recognition (a measure of cognitive function) induced by phencyclidine and methamphetamine treatment in mice (Fujita et al., 2008; Mizoguchi et al., 2008) and reduces the behavioral sensitization induced by methamphetamine and cocaine (Chen, Uz, & Manev, 2009; Zhang et al., 2006). In clinical studies, the results vary, with findings supporting the use of minocycline in schizophrenia, but showing less benefit for nicotine dependence (reviewed in Dean, Data-Franco, Giorlando, & Berk, 2012). Nonsteroidal anti-inflammatory drugs [(e.g., ketoprofen), traditional and selective inhibitors of cyclooxygenase (COX)-2] have also been evaluated for the treatment of psychiatric illness (Berk et al., 2013). For example, in one preclinical study, ketoprofen pretreatment (but not aspirin) dose-dependently attenuated methamphetamine-induced neuroxicity, as measured by reduction of dopamine transporters and accumulation of microglial cells in the striatum (Asanuma, Tsuji, Miyazaki, Miyoshi, & Ogawa, 2003).

Taken together, these preclinical and initial clinical findings indicate that immunotherapeutic strategies which target specific inflammatory pathways, reduce neurotoxicity, promote neuronal repair, and improve neuropsychiatric function may have potential as treatments for methamphetamine dependence and other substances of abuse.

5.3. Future directions

Genetic variation commonly influences psychiatric responses to drugs and medications, and this variation has implications for substance abuse treatment and the use of personalized therapeutic strategies. Thus, future research on methamphetamine dependence could investigate how drug use patterns as well as different treatment approaches may be influenced by genetic polymorphisms, as in alcohol dependence for example [naltrexone treatment seems more effective in carriers of a specific variant of the μ-opioid receptor gene (i.e., G polymorphism of SNP Rs1799971 in the gene OPRM1)] (Anton et al., 2008). Immunologic genes may be good candidates for affecting vulnerability to substance abuse treatment outcomes (Table 7.2). To date, the single nucleotide polymorphisms (SNPs) shown in Table 7.2 have not been directly studied in the context of methamphetamine addiction, but there is preliminary evidence implicating their possible involvement in neuropsychiatric impairments, particularly depression. A polymorphism near the IL28B gene encoding IFN lambda 3 is associated with depressive symptoms, including subjective appetite, energy, and sleep complaints (Lotrich et al., 2010). In addition to IL28B, the risk for developing adverse neuropsychiatric symptoms during or following methamphetamine abuse may be influenced by genetic polymorphisms in other cytokine genes, including IL-6 (IL6) (Bull et al., 2009), tumor necrosis factor-alpha (TNF) (Dogra, Chakravarti, Kar, & Chawla, 2011; Thio et al., 2004), and interferon-γ (IFNG) (Oxenkrug et al., 2011)—genetic variations which may have implications for the management and treatment of methamphetamine dependence and other substance use disorders (with or without comorbid viral infection).

Table 7.2

Candidate polymorphisms with preliminary evidence for possible involvement in adverse neuropsychiatric symptoms relevant to addiction

Gene
symbol
Gene nameID numberRole in neuropsychiatric symptomsSelect references
IDO1Indoleamine 2,3-dioxygenase 1rs9657182Homozygosity for the risk allele (CC) is associated with moderate–severe interferon (IFN)-α-induced depressive symptomsaSmith et al. (2012)
IL28BInterleukin 28B (interferon, lambda 3)rs8099917, rs1297860In adults with HCV, the allele (C) is associated with both better viral clearance and more subjective appetite, energy, and sleep complaintsGe et al. (2009); Lotrich, Loftis, Ferrell, Rabinovitz, and Hauser (2010)
IL6Interleukin 6rs1800795Afunctional G > CSNP in the promoter region is associated with differential IL-6 expression and IL-6 plasma concentrations. The C allele is associated with lower plasma IL-6 during immune activation than the G allele. The “low IL-6 synthesizing” genotype (CC) is associated with significantly fewer symptoms of IFN-α therapy induced depressionBull et al. (2009); Fishman et al. (1998)
IFNGInterferon-gammars2430561IFNG (+874)T/A polymorphism. The presence of T alleles represents a genetic risk factor for the development of IFN-α therapy induced depressionOxenkrug et al. (2011)
TNFTumor necrosis factor-alphars1800630, rs1800629A decreased frequency of the −863CC TNF-α promoter genotype (involved in high production of this proinflammatory cytokine) is present in patients recovered from HCV infection; the G-308A polymorphism of the TNF gene may be independently associated with hypertension, leptin levels, and hypercholesterol emia, leading to metabolic syndromeGupta et al. (2012); Lio et al. (2003)
aIFN-γ and IFN-α transcriptionally induce IDO, the rate-limiting enzyme of the kynurenine pathway of tryptophan metabolism.

It is hypothesized that IFN-induced upregulation of IDO contributes to depression, in part, by shifting tryptophan metabolism from the formation of serotonin to the production of neuroactive/toxic kynurenines (Loftis & Turner, 2010).

ACKNOWLEDGMENTS

This work was in part supported by National Institutes of Health Grant DA018165 to the Methamphetamine Abuse Research Center (MARC) in Portland, Oregon. This material is the result of work supported with resources and the use of facilities at the Portland Veterans Affairs Medical Center and Oregon Health & Science University. Jennifer M. Loftis, Ph.D., and Aaron Janowsky, Ph.D., are Research Scientists at the Portland Veterans Affairs Medical Center. The authors would like to thank Jason A. Laramie, Certified Medical Illustrator, for help in preparing Fig. 7.1. The authors acknowledge Elsevier for the use of Table 7.1, a version of which was previously published in Pharmacology & Therapeutics (i.e., Loftis and Huckans, 2013).

The Department of Veterans Affairs and Oregon Health & Science University own a technology referenced in this review (a partial MHC/neuroantigen peptide construct). The Department of Veterans Affairs, OHSU, and Dr. Loftis have rights to the royalties from the licensing agreement with Artielle (the company that has licensed the technology).

Footnotes

Conflict of interest statement: These potential conflicts of interest have been reviewed and managed by the Conflict of Interest Committees at the Portland Veterans Affairs Medical Center and Oregon Health & Science University.

REFERENCES

  • Abbott NJ, Rönnbäck L, Hansson E. Astrocyte-endothelial interactions at the blood-brain barrier. Nature Reviews Neuroscience. 2006;7:41–53. [PubMed] [Google Scholar]
  • Abdul Muneer PM, Alikunju S, Szlachetka AM, Haorah J. Methamphetamine inhibits the glucose uptake by human neurons and astrocytes: Stabilization by acetyl-L-carnitine. PLoS One. 2011;6:e19258. [PMC free article] [PubMed] [Google Scholar]
  • Achat-Mendes C, Platt DM, Spealman RD. Antagonism of metabotropic glutamate 1 receptors attenuates behavioral effects of cocaine and methamphetamine in squirrel monkeys. The Journal of Pharmacology and Experimental Therapeutics. 2012;343:214–224. [PMC free article] [PubMed] [Google Scholar]
  • Aghajanian A, Wittchen ES, Allingham MJ, Garrett TA, Burridge K. Endothelial cell junctions and the regulation of vascular permeability and leukocyte transmigration. Journal of Thrombosis and Haemostasis. 2008;6:1453–1460. [PMC free article] [PubMed] [Google Scholar]
  • Anton R, Oroszi G, O’Malley S, Couper D, Swift R, Pettinati H, et al. An evaluation of opioid receptor (OPRM1) as a predictor of naltrexone response in the treatment of alcohol dependence. Archives of General Psychiatry. 2008;65:135–144. [PMC free article] [PubMed] [Google Scholar]
  • Archibald SL, Jacobson MW, Fennema-Notestine C, Ogasawara M, Woods SP, Letendre S, et al. Functional interactions of HIV-infection and methamphetamine dependence during motor programming. Psychiatry Research. 2012;202:46–52. [PMC free article] [PubMed] [Google Scholar]
  • Ares-Santos S, Granado N, Oliva I, O’Shea E, Martin ED, Colado MI, et al. Dopamine D(1) receptor deletion strongly reduces neurotoxic effects of methamphetamine. Neurobiology of Disease. 2012;45:810–820. [PubMed] [Google Scholar]
  • Asanuma M, Miyazaki I, Higashi Y, Tsuji T, Ogawa N. Specific gene expression and possible involvement of inflammation in methamphetamine-induced neurotoxicity. Annals of the New York Academy of Sciences. 2004;1025:69–75. [PubMed] [Google Scholar]
  • Asanuma M, Tsuji T, Miyazaki I, Miyoshi K, Ogawa N. Methamphetamine-induced neurotoxicity in mouse brain is attenuated by ketoprofen, a non-steroidal anti-inflammatory drug. Neuroscience Letters. 2003;352:13–16. [PubMed] [Google Scholar]
  • Babusyte A, Kotthoff M, Fiedler J, Krautwurst D. Biogenic amines activate blood leukocytes via trace amine-associated receptors TAAR1 and TAAR2. Journal of Leukocyte Biology. 2013;93:387–394. [PubMed] [Google Scholar]
  • Beardsley PM, Shelton KL, Hendrick E, Johnson KW. The glial cell modulator and phosphodiesterase inhibitor, AV411 (ibudilast), attenuates prime- and stress-induced methamphetamine relapse. European Journal of Pharmacology. 2010;637:102–108. [PMC free article] [PubMed] [Google Scholar]
  • Berk M, Dean O, Drexhage H, McNeil JJ, Moylan S, O’Neil A, et al. Aspirin: A review of its neurobiological properties and therapeutic potential for mental illness. BMC Medicine. 2013;11:74. [PMC free article] [PubMed] [Google Scholar]
  • Blednov YA, Mayfield RD, Belknap J, Harris RA. Behavioral actions of alcohol: Phenotypic relations from multivariate analysis of mutant mouse data. Genes, Brain, and Behavior. 2012;11:424–435. [PMC free article] [PubMed] [Google Scholar]
  • Blednov YA, Ponomarev I, Geil C, Bergeson S, Koob GF, Harris RA. Neuroimmune regulation of alcohol consumption: Behavioral validation of genes obtained from genomic studies. Addiction Biology. 2012;17:108–120. [PMC free article] [PubMed] [Google Scholar]
  • Bogatcheva NV, Verin AD. The role of cytoskeleton in the regulation of vascular endothelial barrier function. Microvascular Research. 2008;76:202–207. [PMC free article] [PubMed] [Google Scholar]
  • Borowsky B, Adham N, Jones KA, Raddatz R, Artymyshyn R, Ogozalek KL, et al. Trace amines: Identification of a family of mammalian G protein-coupled receptors. Proceedings of the National Academy of Sciences of the United States of America. 2001;98:8966–8971. [PMC free article] [PubMed] [Google Scholar]
  • 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. [PubMed] [Google Scholar]
  • Bowyer JF, Robinson B, Ali S, Schmued LC. 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. [PubMed] [Google Scholar]
  • Brown RC, Mark KS, Egleton RD, Huber JD, Burroughs AR, Davis TP. Protection against hypoxia-induced increase in blood-brain barrier permeability: Role of tight junction proteins and NFkappaB. Journal of Cell Science. 2003;116:693–700. [PubMed] [Google Scholar]
  • Buch S, Yao H, Guo M, Mori T, Mathias-Costa B, Singh V, et al. Cocaine and HIV-1 interplay in CNS: Cellular and molecular mechanisms. Current HIV Research. 2012;10:425–428. [PMC free article] [PubMed] [Google Scholar]
  • Buchacz K, McFarland W, Kellogg TA, Loeb L, Holmberg SD, Dilley J, et al. Amphetamine use is associated with increased HIV incidence among men who have sex with men in San Francisco. AIDS. 2005;19:1423–1424. [PubMed] [Google Scholar]
  • Buchanan JB, Sparkman NL, Johnson RW. A neurotoxic regimen of methamphetamine exacerbates the febrile and neuroinflammatory response to a subsequent peripheral immune stimulus. Journal of Neuroinflammation. 2010a;7:82. [PMC free article] [PubMed] [Google Scholar]
  • Buchanan JB, Sparkman NL, Johnson RW. Methamphetamine sensitization attenuates the febrile and neuroinflammatory response to a subsequent peripheral immune stimulus. Brain, Behavior, and Immunity. 2010b;24:502–511. [PMC free article] [PubMed] [Google Scholar]
  • Bull SJ, Huezo-Diaz P, Binder EB, Cubells JF, Ranjith G, Maddock C, et al. Functional polymorphisms in the interleukin-6 and serotonin transporter genes, and depression and fatigue induced by interferon-alpha and ribavirin treatment. Molecular Psychiatry. 2009;14:1095–1104. [PMC free article] [PubMed] [Google Scholar]
  • Bunzow JR, Sonders MS, Arttamangkul S, Harrison LM, Zhang G, Quigley DI, et al. Amphetamine, 3,4-methylenedioxymethamphetamine, lysergic acid diethylamide, and metabolites of the catecholamine neurotransmitters are agonists of a rat trace amine receptor. Molecular Pharmacology. 2001;60:1181–1188. [PubMed] [Google Scholar]
  • Cadet JL, Krasnova IN. Interactions of methamphetamine and HIV: Cellular and molecular mechanisms of toxicity potentiation. Neurotoxicity Research. 2007;112:181–204. [PubMed] [Google Scholar]
  • Cardona AE, Pioro EP, Sasse ME, Kostenko V, Cardona SM, Dijkstra IM, et al. Control of microglial neurotoxicity by the fractalkine receptor. Nature Neuroscience. 2006;9:917–924. [PubMed] [Google Scholar]
  • Cerny EH, Cerny T. Vaccines against nicotine. Human Vaccines. 2009;5:200–205. [PubMed] [Google Scholar]
  • Chana G, Everall IP, Crews L, Langford D, Adame A, Grant I, et al. Cognitive deficits and degeneration of interneurons in HIV + methamphetamine users. Neurology. 2006;67:1486–1489. [PubMed] [Google Scholar]
  • Chang L, Cloak C, Patterson K, Grob C, Miller EN, Ernst T. Enlarged striatum in abstinent methamphetamine abusers: A possible compensatory response. Biological Psychiatry. 2005;57(9):967–974. [PMC free article] [PubMed] [Google Scholar]
  • Chang L, Ernst T, Speck O, Grob CS. Additive effects of HIV and chronic methamphetamine use on brain metabolite abnormalities. The American Journal of Psychiatry. 2005;162:361–369. [PMC free article] [PubMed] [Google Scholar]
  • Chen H, Uz T, Manev H. Minocycline affects cocaine sensitization in mice. Neuroscience Letters. 2009;452:258–261. [PMC free article] [PubMed] [Google Scholar]
  • Cherner M, Suarez P, Casey C, Deiss R, Letendre S, Marcotte T, et al. Methamphetamine use parameters do not predict neuropsychological impairment in currently abstinent dependent adults. Drug and Alcohol Dependence. 2010;106:154–163. [PMC free article] [PubMed] [Google Scholar]
  • Cisneros IE, Ghorpade A. HIV-1, methamphetamine and astrocyte glutamate regulation: Combined excitotoxic implications for neuro-AIDS. Current HIV Research. 2012;10:392–406. [PMC free article] [PubMed] [Google Scholar]
  • Clark KH, Wiley CA, Bradberry CW. Psychostimulant abuse and neuroinflammation: Emerging evidence of their interconnection. Neurotoxicity Research. 2013;23:174–188. [PubMed] [Google Scholar]
  • Crawford JT, Roberts DC, Beveridge TJ. The group II metabotropic glutamate receptor agonist, LY379268, decreases methamphetamine self-administration in rats. Drug and Alcohol Dependence. 2013;132:414–419. [PMC free article] [PubMed] [Google Scholar]
  • Dean OM, Data-Franco J, Giorlando F, Berk M. Minocycline: Therapeutic potential in psychiatry. CNS Drugs. 2012;26:391–401. [PubMed] [Google Scholar]
  • Dogra G, Chakravarti A, Kar P, Chawla YK. Polymorphism of tumor necrosis factor-α and interleukin-10 gene promoter region in chronic hepatitis C virus patients and their effect on pegylated interferon-α therapy response. Human Immunology. 2011;72:935–939. [PubMed] [Google Scholar]
  • Escubedo E, Camarasa J, Chipana C, García-Ratés S, Pubill D. Involvement of nicotinic receptors in methamphetamine- and MDMA-induced neurotoxicity: Pharmacological implications. International Review of Neurobiology. 2009;88:121–166. [PubMed] [Google Scholar]
  • Escubedo E, Chipana C, Pérez-Sánchez M, Camarasa J, Pubill D. Methyllycaconitine prevents methamphetamine-induced effects in mouse striatum: Involvement of alpha7 nicotinic receptors. The Journal of Pharmacology and Experimental Therapeutics. 2005;315:658–667. [PubMed] [Google Scholar]
  • Escubedo E, Guitart L, Sureda FX, Jiménez A, Pubill D, Pallàs M, et al. Microgliosis and down-regulation of adenosine transporter induced by methamphetamine in rats. Brain Research. 1998;814:120–126. [PubMed] [Google Scholar]
  • Eshleman AJ, Calligaro DO, Eldefrawi ME. Allosteric regulation by sodium of the binding of [3H]cocaine and [3H]GBR 12935 to rat and bovine striata. Membrane Biochemistry. 1993;10:129–144. [PubMed] [Google Scholar]
  • Eshleman AJ, Wolfrum KM, Hatfield MG, Johnson RA, Murphy KV, Janowsky A. Substituted methcathinones differ in transporter and receptor interactions. Biochemical Pharmacology. 2013;85:1803–1815. [PMC free article] [PubMed] [Google Scholar]
  • Fantegrossi WE, Ciullo JR, Wakabayashi KT, De La Garza R, 2nd, Traynor JR, Woods JH. A comparison of the physiological, behavioral, neurochemical and microglial effects of methamphetamine and 3,4-methylenedioxymethamphetamine in the mouse. Neuroscience. 2008;151:533–543. [PMC free article] [PubMed] [Google Scholar]
  • Fishman D, Faulds G, Jeffery R, Mohamed-Ali V, Yudkin JS, Humphries S, et al. The effect of novel polymorphisms in the interleukin-6 (IL-6) gene on IL-6 transcription and plasma IL-6 levels, and an association with systemic-onset juvenile chronic arthritis. The Journal of Clinical Investigation. 1998;102:1369–1376. [PMC free article] [PubMed] [Google Scholar]
  • Friend DM, Keefe KA. Glial reactivity in resistance to methamphetamine-induced neurotoxicity. Journal of Neurochemistry. 2013;125:566–574. [PMC free article] [PubMed] [Google Scholar]
  • Fujita Y, Ishima T, Kunitachi S, Hagiwara H, Zhang L, Iyo M, et al. Phencyclidine-induced cognitive deficits in mice are improved by subsequent subchronic administration of the antibiotic drug minocycline. Progress in Neuro-Psychopharmacology & Biological Psychiatry. 2008;32:336–339. [PubMed] [Google Scholar]
  • Fuller BE, Loftis JM, Rodriguez VL, McQuesten MJ, Hauser P. Psychiatric and substance use disorders comorbidities in veterans with hepatitis C virus and HIV coinfection. Current Opinion in Psychiatry. 2009;22:401–408. [PubMed] [Google Scholar]
  • Gavrilin MA, Mathes LE, Podell M. Methamphetamine enhances cell-associated feline immunodeficiency virus replication in astrocytes. Journal of Neurovirology. 2002;8:240–249. [PubMed] [Google Scholar]
  • Ge D, Fellay J, Thompson AJ, Simon JS, Shianna KV, Urban TJ, et al. Genetic variation in IL28B predicts hepatitis C treatment-induced viral clearance. Nature. 2009;461:399–401. [PubMed] [Google Scholar]
  • Gentry WB, Rüedi-Bettschen D, Owens SM. Development of active and passive human vaccines to treat methamphetamine addiction. Human Vaccines. 2009;5:206–213. [PMC free article] [PubMed] [Google Scholar]
  • Gołembiowska K, Konieczny J, Wolfarth S, Ossowska K. Neuroprotective action of MPEP, a selective mGluR5 antagonist, in methamphetamine-induced dopaminergic neurotoxicity is associated with a decrease in dopamine outflow and inhibition of hyperthermia in rats. Neuropharmacology. 2003;45:484–492. [PubMed] [Google Scholar]
  • Goniewicz ML, Delijewski M. Nicotine vaccines to treat tobacco dependence. Human Vaccines & Immunotherapeutics. 2013;9:13–25. [PMC free article] [PubMed] [Google Scholar]
  • Granado N, Ares-Santos S, Oliva I, O’Shea E, Martin ED, Colado MI, et al. Dopamine D2-receptor knockout mice are protected against dopaminergic neurotoxicity induced by methamphetamine or MDMA. Neurobiology of Disease. 2011;42:391–403. [PubMed] [Google Scholar]
  • Granado N, Lastres-Becker I, Ares-Santos S, Oliva I, Martin E, Cuadrado A, et al. Nrf2 deficiency potentiates methamphetamine-induced dopaminergic axonal damage and gliosis in the striatum. Glia. 2011;59:1850–1863. [PubMed] [Google Scholar]
  • Gross NB, Duncker PC, Marshall JF. Cortical ionotropic glutamate receptor antagonism protects against methamphetamine-induced striatal neurotoxicity. Neuroscience. 2011;199:272–283. [PMC free article] [PubMed] [Google Scholar]
  • Guilarte TR, Nihei MK, McGlothan JL, Howard AS. Methamphetamine-induced deficits of brain monoaminergic neuronal markers: Distal axotomy or neuronal plasticity. Neuroscience. 2003;122:499–513. [PubMed] [Google Scholar]
  • Gupta V, Gupta A, Jafar T, Gupta V, Agrawal S, Srivastava N, et al. Association of TNF-α promoter gene G-308A polymorphism with metabolic syndrome, insulin resistance, serum TNF-α and leptin levels in Indian adult women. Cytokine. 2012;57:32–36. [PubMed] [Google Scholar]
  • Harms R, Morsey B, Boyer CW, Fox HS, Sarvetnick N. Methamphetamine administration targets multiple immune subsets and induces phenotypic alterations suggestive of immunosuppression. PLoS One. 2012;7:e49897. [PMC free article] [PubMed] [Google Scholar]
  • Hebert MA, O’Callaghan JP. Protein phosphorylation cascades associated with methamphetamine-induced glial activation. Annals of the New York Academy of Sciences. 2000;914:238–262. [PubMed] [Google Scholar]
  • Herrold AA, Voigt RM, Napier TC. mGluR5 is necessary for maintenance of methamphetamine-induced associative learning. European Neuropsychopharmacology. 2013;23:691–696. [PMC free article] [PubMed] [Google Scholar]
  • Hozumi H, Asanuma M, Miyazaki I, Fukuoka S, Kikkawa Y, Kimoto N, et al. Protective effects of interferon-gamma against methamphetamine-induced neurotoxicity. Toxicology Letters. 2008;177:123–129. [PubMed] [Google Scholar]
  • In SW, Son EW, Rhee DK, Pyo S. Modulation of murine macrophage function by methamphetamine. Journal of Toxicology and Environmental Health. Part A. 2004;67:1923–1937. [PubMed] [Google Scholar]
  • In SW, Son EW, Rhee DK, Pyo S. Methamphetamine administration produces immunomodulation in mice. Journal of Toxicology and Environmental Health. 2005;68:2133–2145. [PubMed] [Google Scholar]
  • Jernigan TL, Gamst AC, Archibald SL, Fennema-Notestine C, Mindt MR, Marcotte TD, et al. Effects of methamphetamine dependence and HIV infection on cerebral morphology. The American Journal of Psychiatry. 2005;162:1461–1472. [PubMed] [Google Scholar]
  • Kinsey BM, Jackson DC, Orson FM. Anti-drug vaccines to treat substance abuse. Immunology and Cell Biology. 2009;87:309–314. [PubMed] [Google Scholar]
  • Kosten T, Domingo C, Orson F, Kinsey B. Vaccines against stimulants: Cocaine and MA. British Journal of Clinical Pharmacology. 2014;77:368–374. [PMC free article] [PubMed] [Google Scholar]
  • Kousik SM, Napier TC, Carvey PM. The effects of psychostimulant drugs on blood brain barrier function and neuroinflammation. Frontiers in Pharmacology. 2012;3:121. [PMC free article] [PubMed] [Google Scholar]
  • 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. Annals of the New York Academy of Sciences. 2006;1074:31–41. [PubMed] [Google Scholar]
  • Lau A, Tymianski M. Glutamate receptors, neurotoxicity and neurodegeneration. Pflügers Archiv. 2010;460:525–542. [PubMed] [Google Scholar]
  • LaVoie MJ, Card JP, Hastings TG. Microglial activation precedes dopamine terminal pathology in methamphetamine-induced neurotoxicity. Experimental Neurology. 2004;187:47–57. [PubMed] [Google Scholar]
  • Lee YW, Hennig B, Yao J, Toborek M. Methamphetamine induces AP-1 and NF-kappaB binding and transactivation in human brain endothelial cells. Journal of Neuroscience Research. 2001;66:583–591. [PubMed] [Google Scholar]
  • Letendre SL, Cherner M, Ellis RJ, Marquie-Beck J, Gragg B, Marcotte T, et al. The effects of hepatitis C, HIV, and methamphetamine dependence on neuropsychological performance: Biological correlates of disease. AIDS. 2005;19(Suppl. 3):S72–S78. [PubMed] [Google Scholar]
  • Li QQ, Luo YX, Sun CY, Xue YX, Zhu WL, Shi HS, et al. A morphine/heroin vaccine with new hapten design attenuates behavioral effects in rats. Journal of Neurochemistry. 2011;119:1271–1281. [PubMed] [Google Scholar]
  • Liang H, Wang X, Chen H, Song L, Ye L, Wang SH, et al. Methamphetamine enhances HIV infection of macrophages. The American Journal of Pathology. 2008;172:1617–1624. [PMC free article] [PubMed] [Google Scholar]
  • Lio D, Caruso C, Di Stefano R, Colonna Romano G, Ferraro D, Scola L, et al. IL-10 and TNF-alpha polymorphisms and the recovery from HCV infection. Human Immunology. 2003;64:674–680. [PubMed] [Google Scholar]
  • Loftis JM, Choi D, Hoffman W, Huckans MS. Methamphetamine causes persistent immune dysregulation: A cross-species, translational report. Neurotoxicity Research. 2011;20:59–68. [PMC free article] [PubMed] [Google Scholar]
  • Loftis JM, Huckans M. Substance use disorders: Psychoneuroimmunological mechanisms and new targets for therapy. Pharmacology & Therapeutics. 2013;139:289–300. [PMC free article] [PubMed] [Google Scholar]
  • Loftis JM, Matthews AM, Hauser P. Psychiatric and substance use disorders in individuals with hepatitis C: Epidemiology and management. Drugs. 2006;66:155–174. [PubMed] [Google Scholar]
  • Loftis JM, Turner EH. Novel treatment strategies for depression in patients with hepatitis C. Psychosomatics. 2010;51:357–358. [PubMed] [Google Scholar]
  • Loftis JM, Wilhelm CJ, Vandenbark AA, Huckans M. Partial MHC/neuroantigen peptide constructs: A potential neuroimmune-based treatment for methamphetamine addiction. PLoS One. 2013;8:e56306. [PMC free article] [PubMed] [Google Scholar]
  • Lotrich FE, Loftis JM, Ferrell RE, Rabinovitz M, Hauser P. IL28B polymorphism is associated with both side effects and clearance of hepatitis c during interferon-alpha therapy. Journal of Interferon & Cytokine Research. 2010;31(3):331–336. [PMC free article] [PubMed] [Google Scholar]
  • Lu DY, Yu WH, Yeh WL, Tang CH, Leung YM, Wong KL, et al. Hypoxia-induced matrix metalloproteinase-13 expression in astrocytes enhances permeability of brain endothelial cells. Journal of Cellular Physiology. 2009;220:163–173. [PubMed] [Google Scholar]
  • Mahajan SD, Hu Z, Reynolds JL, Aalinkeel R, Schwartz SA, Nair MP. Methamphetamine modulates gene expression patterns in monocyte derived mature dendritic cells: Implications for HIV-1 pathogenesis. Molecular Diagnosis & Therapy. 2006;10:257–269. [PubMed] [Google Scholar]
  • Martinez LR, Mihu MR, Gacser A, Santambrogio L, Nosanchuk D. Methamphetamine enhances histoplasmosis by immunosuppression of the host. The Journal of Infectious Diseases. 2009;200:131–141. [PubMed] [Google Scholar]
  • Martins T, Baptista S, Goncalves J, Leal E, Milhazes N, Borges F, et al. Methamphetamine transiently increases the blood-brain barrier permeability in the hippocampus: Role of tight junction proteins andmatrixmetalloproteinase-9. Brain Research. 2011;1411:28–40. [PubMed] [Google Scholar]
  • Mizoguchi H, Takuma K, Fukakusa A, Ito Y, Nakatani A, Ibi D, et al. Improvement by minocycline of methamphetamine-induced impairment of recognition memory in mice. Psychopharmacology. 2008;196:233–241. [PubMed] [Google Scholar]
  • Narita M, Miyatake M, Narita M, Shibasaki M, Shindo K, Nakamura A, et al. Direct evidence of astrocytic modulation in the development of rewarding effects induced by drugs of abuse. Neuropsychopharmacology. 2006;31:2476–2488. [PubMed] [Google Scholar]
  • Nath A, Geiger J. Neurobiological aspects of human immunodeficiency virus infection: Neurotoxic mechanisms. Progress in Neurobiology. 1998;54:19–33. [PubMed] [Google Scholar]
  • Nath A, Maragos WF, Avison MJ, Schmitt FA, Berger JR. Acceleration of HIV dementia with methamphetamine and cocaine. Journal of Neurovirology. 2001;7:66–71. [PubMed] [Google Scholar]
  • Nelson DA, Tolbert MD, Singh SJ, Bost KL. Expression of neuronal trace amine-associated receptor (Taar) mRNAs in leukocytes. Journal of Neuroimmunology. 2007;192:21–30. [PMC free article] [PubMed] [Google Scholar]
  • Northrop NA, Smith LP, Yamamoto BK, Eyerman DJ. Regulation of glutamate release by α7 nicotinic receptors: Differential role in methamphetamine-induced damage to dopaminergic and serotonergic terminals. The Journal of Pharmacology and Experimental Therapeutics. 2011;336:900–907. [PMC free article] [PubMed] [Google Scholar]
  • Northrop NA, Yamamoto BK. Persistent neuroinflammatory effects of serial exposure to stress and methamphetamine on the blood-brain barrier. Journal of Neuroimmune Pharmacology. 2012;7:951–968. [PMC free article] [PubMed] [Google Scholar]
  • Orson FM, Kinsey BM, Singh RA, Wu Y, Kosten TR. Vaccines for cocaine abuse. Human Vaccines. 2009;5:194–199. [PMC free article] [PubMed] [Google Scholar]
  • Oxenkrug G, Perianayagam M, Mikolich D, Requintina P, Shick L, Ruthazer R, et al. Interferon-gamma (+874) T/A genotypes and risk of IFN-alpha-induced depression. Journal of Neural Transmission. 2011;118:271–274. [PMC free article] [PubMed] [Google Scholar]
  • Panas MW, Xie Z, Panas HN, Hoener MC, Vallender EJ, Miller GM. Trace amine associated receptor 1 signaling in activated lymphocytes. Journal of Neuroimmune Pharmacology. 2012;7:866–876. [PMC free article] [PubMed] [Google Scholar]
  • Persidsky Y, Ramirez SH, Haorah J, Kanmogne GD. Blood-brain barrier: Structural components and function under physiologic and pathologic conditions. Journal of Neuroimmune Pharmacology. 2006;1:223–236. [PubMed] [Google Scholar]
  • Potula R, Hawkins BJ, Cenna JM, Fan S, Dykstra H, Ramierez SH, et al. Methamphetamine causes mitochondrial oxidative damage in human T lymphocytes leading to functional impairment. Journal of Immunology. 2010;185:2867–2876. [PMC free article] [PubMed] [Google Scholar]
  • Pubill D, Chipana C, Camins A, Pallàs M, Camarasa J, Escubedo E. Free radical production induced by methamphetamine in rat striatal synaptosomes. Toxicology and Applied Pharmacology. 2005;204:57–68. [PubMed] [Google Scholar]
  • Ramirez SH, Potula R, Fan S, Eidem T, Papugani A, Reichenbach N, et al. Methamphetamine disrupts blood-brain barrier function by induction of oxidative stress in brain endothelial cells. Journal of Cerebral Blood Flow and Metabolism. 2009;29:1933–1945. [PMC free article] [PubMed] [Google Scholar]
  • Reynolds JL, Mahajan SD, Aalinkeel R, Nair B, Sykes DE, Schwartz SA. Proteomic analyses of the effects of drugs of abuse on monocyte-derived mature dendritic cells. Immunological Investigations. 2009;38:526–550. [PMC free article] [PubMed] [Google Scholar]
  • Rippeth JD, Heaton RK, Carey CL, Marcotte TD, Moore DJ, Gonzalez R, et al. Methamphetamine dependence increases risk of neuropsychological impairment in HIV infected persons. Journal of the International Neuropsychological Society. 2004;10:1–14. [PubMed] [Google Scholar]
  • Robson MJ, Turner RC, Naser ZJ, McCurdy CR, O’Callaghan JP, Huber JD, et al. SN79, a sigma receptor antagonist, attenuates methamphetamine-induced astrogliosis through a blockade of OSMR/gp130 signaling and STAT3 phosphorylation. Experimental Neurology. 2014;254:180–189. [PMC free article] [PubMed] [Google Scholar]
  • Saito M, Yamaguchi T, Kawata T, Ito H, Kanai T, Terada M, et al. Effects of methamphetamine on cortisone concentration, NK cell activity and mitogen response of T-lymphocytes in female cynomolgus monkeys. Experimental Animals. 2006;55:477–481. [PubMed] [Google Scholar]
  • Schwarz JM, Hutchinson MR, Bilbo SD. Early-life experience decreases drug-induced reinstatement of morphine CPP in adulthood via microglial-specific epigenetic programming of anti-inflammatory IL-10 expression. The Journal of Neuroscience. 2011;31:17835–17847. [PMC free article] [PubMed] [Google Scholar]
  • Sekine Y, Ouchi Y, Sugihara G, Takei N, Yoshikawa E, Nakamura K, et al. Methamphetamine causes microglial activation in the brains of human abusers. The Journal of Neuroscience. 2008;28:5756–5761. [PMC free article] [PubMed] [Google Scholar]
  • Shah A, Silverstein PS, Singh DP, Kumar A. Involvement of metabotropic glutamate receptor 5, AKT/PI3K signaling and NF-κB pathway in methamphetamine-mediated increase in IL-6 and IL-8 expression in astrocytes. Journal of Neuroinflammation. 2012;9:52. [PMC free article] [PubMed] [Google Scholar]
  • Shen XY, Kosten TA, Lopez AY, Kinsey BM, Kosten TR, Orson FM. A vaccine against methamphetamine attenuates its behavioral effects in mice. Drug and Alcohol Dependence. 2013;129:41–48. [PMC free article] [PubMed] [Google Scholar]
  • Shin EJ, Duong CX, Nguyen TX, Bing G, Bach JH, Park DH, et al. PKCδ inhibition enhances tyrosine hydroxylase phosphorylation in mice after methamphetamine treatment. Neurochemistry International. 2011;59:39–50. [PMC free article] [PubMed] [Google Scholar]
  • Shin EJ, Duong CX, Nguyen XK, Li Z, Bing G, Bach JH, et al. Role of oxidative stress in methamphetamine-induced dopaminergic toxicity mediated by protein kinase Cδ Behavioural Brain Research. 2012;232:98–113. [PMC free article] [PubMed] [Google Scholar]
  • Sinha S, Subramanian S, Proctor TM, Kaler LJ, Grafe M, Dahan R, et al. A promising therapeutic approach for multiple sclerosis: Recombinant T-cell receptor ligands modulate experimental autoimmune encephalomyelitis by reducing interleukin-17 production and inhibiting migration of encephalitogenic cells into the CNS. The Journal of Neuroscience. 2007;27:12531–12539. [PMC free article] [PubMed] [Google Scholar]
  • Smith KJ, Butler TR, Prendergast MA. Inhibition of sigma-1 receptor reduces N-methyl-D-aspartate induced neuronal injury in methamphetamine-exposed and-naive hippocampi. Neuroscience Letters. 2010;13(481):144–148. [PMC free article] [PubMed] [Google Scholar]
  • Smith AK, Simon JS, Gustafson EL, Noviello S, Cubells JF, Epstein MP, et al. Association of a polymorphism in the indoleamine- 2,3-dioxygenase gene and interferon-α-induced depression in patients with chronic hepatitis C. Molecular Psychiatry. 2012;17:781–789. [PMC free article] [PubMed] [Google Scholar]
  • Snider SE, Hendrick ES, Beardsley PM. Glial cell modulators attenuate methamphetamine self-administration in the rat. European Journal of Pharmacology. 2013;701:124–130. [PMC free article] [PubMed] [Google Scholar]
  • Snider SE, Vunck SA, van den Oord EJ, Adkins DE, McClay JL, Beardsley PM. The glial cell modulators, ibudilast and its amino analog, AV1013, attenuate methamphetamine locomotor activity and its sensitization in mice. European Journal of Pharmacology. 2012;679:75–80. [PMC free article] [PubMed] [Google Scholar]
  • Sriram K, Benkovic SA, Hebert MA, Miller DB, O’Callaghan JP. Induction of gp130-related cytokines and activation of JAK2/STAT3 pathway in astrocytes precedes up-regulation of glial fibrillary acidic protein in the 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine model of neurodegeneration: Key signaling pathway for astrogliosis in vivo? The Journal of Biological Chemistry. 2004;279:19936–19947. [PubMed] [Google Scholar]
  • Stevens MW, Tawney RL, West CM, Kight AD, Henry RL, Owens SM, et al. Preclinical characterization of an anti-methamphetamine monoclonal antibody for human use. MAbs. 2014;6:547–555. [PMC free article] [PubMed] [Google Scholar]
  • Subramanian S, Zhang B, Kosaka Y, Burrows GG, Grafe MR, Vandenbark AA, et al. Recombinant T cell receptor ligand treats experimental stroke. Stroke. 2009;40:2539–2545. [PMC free article] [PubMed] [Google Scholar]
  • Sulzer D, Chen TK, Lau YY, Kristensen H, Rayport S, Ewing A. Amphetamine redistributes dopamine from synaptic vesicles to the cytosol and promotes reverse transport. The Journal of Neuroscience. 1995;15:4102–4108. [PMC free article] [PubMed] [Google Scholar]
  • Suzuki S, Kawamata J, Matsushita T, Matsumura A, Hisahara S, Takata K, et al. 3-[(2,4-Dimethoxy)benzylidene]-anabaseine dihydrochloride protects against 6-hydroxydopamine-induced parkinsonian neurodegeneration through α7 nicotinic acetylcholine receptor stimulation in rats. Journal of Neuroscience Research. 2013;91:462–471. [PubMed] [Google Scholar]
  • Talloczy Z, Martinez J, Joset D, Ray Y, Gacser A, Toussi S, et al. Methamphetamine inhibits antigen processing, presentation, and phagocytosis. PLoS Pathogens. 2008;4:e28. [PMC free article] [PubMed] [Google Scholar]
  • Tamura S, Morikawa Y, Senba E. Localization of oncostatin M receptor beta in adult and developing CNS. Neuroscience. 2003;119:991–997. [PubMed] [Google Scholar]
  • Taylor MJ, Letendre SL, Schweinsburg BC, Alhassoon OM, Brown GG, Gongvatana A, et al. Hepatitis Cvirus infection is associated with reduced white matter N-acetylaspartate in abstinent methamphetamine users. Journal of the International Neuropsychological Society. 2004;10:110–113. [PubMed] [Google Scholar]
  • Theodore S, Cass WA, Maragos WF. Involvement of cytokines in human immunodeficiency virus-1 protein Tat and methamphetamine interactions in the striatum. Experimental Neurology. 2006;199:490–498. [PubMed] [Google Scholar]
  • Theodore S, Cass WA, Nath A, Steiner J, Young K, Maragos WF. Inhibition of tumor necrosis factor-alpha signaling prevents human immunodeficiency virus-1 protein Tat and methamphetamine interaction. Neurobiology of Disease. 2006;23:663–668. [PubMed] [Google Scholar]
  • Theodore S, Stolberg S, Cass WA, Maragos WF. Human immunodeficiency virus-1 protein tat and methamphetamine interactions. Annals of the New York Academy of Sciences. 2006;1074:178–190. [PubMed] [Google Scholar]
  • Thio CL, Goedert JJ, Mosbruger T, Vlahov D, Strathdee SA, O’Brien SJ, et al. An analysis of tumor necrosis factor alpha gene polymorphisms and haplotypes with natural clearance of hepatitis C virus infection. Genes and Immunity. 2004;5:294–300. [PubMed] [Google Scholar]
  • Thomas DM, Angoa Pérez M, Francescutti-Verbeem DM, Shah MM, Kuhn DM. The role of endogenous serotonin in methamphetamine-induced neurotoxicity to dopamine nerve endings of the striatum. Journal of Neurochemistry. 2010;115:595–605. [PMC free article] [PubMed] [Google Scholar]
  • Thomas DM, Francescutti-Verbeem DM, Kuhn DM. Methamphetamine-induced neurotoxicity and microglial activation are not mediated by fractalkine receptor signaling. Journal of Neurochemistry. 2008a;106:696–705. [PMC free article] [PubMed] [Google Scholar]
  • Thomas DM, Francescutti-Verbeem DM, Kuhn DM. The newly synthesized pool of dopamine determines the severity of methamphetamine-induced neurotoxicity. Journal of Neurochemistry. 2008b;105:605–616. [PMC free article] [PubMed] [Google Scholar]
  • Thomas DM, Francescutti-Verbeem DM, Liu X, Kuhn DM. Identification of differentially regulated transcripts in mouse striatum following methamphetamine treatment–an oligonucleotide microarray approach. Journal of Neurochemistry. 2004;88:380–393. [PubMed] [Google Scholar]
  • Thomas DM, Walker PD, Benjamins JA, Geddes TJ, Kuhn DM. Methamphetamine neurotoxicity in dopamine nerve endings of the striatum is associated with microglial activation. The Journal of Pharmacology and Experimental Therapeutics. 2004;311:1–7. [PubMed] [Google Scholar]
  • Vaarmann A, Gandhi S, Abramov AY. Dopamine induces Ca2+ signaling in astrocytes through reactive oxygen species generated by monoamine oxidase. The Journal of Biological Chemistry. 2010;285:25018–25023. [PMC free article] [PubMed] [Google Scholar]
  • Vandenbark AA, Rich C, Mooney J, Zamora A, Wang C, Huan J, et al. Recombinant TCR ligand induces tolerance to myelin oligodendrocyte glycoprotein 35–55 peptide and reverses clinical and histological signs of chronic experimental autoimmune encephalomyelitis in HLA-DR2 transgenic mice. Journal of Immunology. 2003;171:127–133. [PubMed] [Google Scholar]
  • Volkow ND, Chang L, Wang GJ, Fowler JS, Franceschi D, Sedler M, et al. Loss of dopamine transporters in methamphetamine abusers recovers with protracted abstinence. The Journal of Neuroscience. 2001;21:9414–9418. [PMC free article] [PubMed] [Google Scholar]
  • Wang C, Gold BG, Kaler LJ, Yu X, Afentoulis ME, Burrows GG, et al. Antigen-specific therapy promotes repair of myelin and axonal damage in established EAE. Journal of Neurochemistry. 2006;98:1817–1827. [PMC free article] [PubMed] [Google Scholar]
  • Wang Q, Shin EJ, Nguyen XK, Li Q, Bach JH, Bing G, et al. Endogenous dynorphin protects against neurotoxin-elicited nigrostriatal dopaminergic neuron damage and motor deficits in mice. Journal of Neuroinflammation. 2012;9:124. [PMC free article] [PubMed] [Google Scholar]
  • Wasik AM, Millan MJ, Scanlan T, Barnes NM. Gordon evidence for functional trace amine associated receptor-1 in normal and malignant B cells. Leukemia Research. 2012;36:245–249. [PMC free article] [PubMed] [Google Scholar]
  • Weber E, Morgan EE, Iudicello JE, Blackstone K, Grant I, Ellis RJ, et al. Substance use is a risk factor for neurocognitive deficits and neuropsychiatric distress in acute and early HIV infection. Journal of Neurovirology. 2013;19:65–74. [PMC free article] [PubMed] [Google Scholar]
  • Wey SP, Wu HY, Chang FC, Jan TR. Methamphetamine and diazepam suppress antigen-specific cytokine expression and antibody production in ovalbuminsensitized BALB/c mice. Toxicology Letters. 2008;181:157–162. [PubMed] [Google Scholar]
  • Whitney NP, Eidem TM, Peng H, Huang Y, Zheng JC. Inflammation mediates varying effects in neurogenesis: Relevance to the pathogenesis of brain injury and neurodegenerative disorders. Journal of Neurochemistry. 2009;108:1343–1359. [PMC free article] [PubMed] [Google Scholar]
  • Wilhelm CJ, Hashimoto JG, Roberts ML, Loftis JM, Wiren KM. Distinct glucocorticoid signaling associated with ethanol-induced inflammotoxicity. 37th annual research society on alcoholism (RSA) scientific meeting. 2014 Jun; 2014 (abstract accepted). [Google Scholar]
  • Wilhelm CJ, Johnson RA, Lysko PG, Eshleman AJ, Janowsky A. Effects of methamphetamine and lobeline on vesicular monoamine and dopamine transporter-mediated dopamine release in a cotransfected model system. The Journal of Pharmacology and Experimental Therapeutics. 2004;310:1142–1151. [PubMed] [Google Scholar]
  • Wires ES, Alvarez D, Dobrowolski C, Wang Y, Morales M, Karn J, et al. Methamphetamine activates nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) and induces human immunodeficiency virus (HIV) transcription in human microglial cells. Journal of Neurovirology. 2012;18:400–410. [PMC free article] [PubMed] [Google Scholar]
  • Yamamoto BK, Moszczynska A, Gudelsky GA. Amphetamine toxicities: Classical and emerging mechanisms. Annals of the New York Academy of Sciences. 2010;1187:101–121. [PMC free article] [PubMed] [Google Scholar]
  • Yamamoto BK, Raudensky J. The role of oxidative stress, metabolic compromise, and inflammation in neuronal injury produced by amphetamine-related drugs of abuse. Journal of Neuroimmune Pharmacology. 2008;3:203–217. [PMC free article] [PubMed] [Google Scholar]
  • Yan Y, Nitta A, Koseki T, Yamada K, Nabeshima T. Dissociable role of tumor necrosis factor alpha gene deletion in methamphetamine self-administration and cue-induced relapsing behavior in mice. Psychopharmacology. 2012;221:427–436. [PubMed] [Google Scholar]
  • Ye L, Peng JS, Wang X, Wang YJ, Luo GX, Ho WZ. Methamphetamine enhances Hepatitis C virus replication in human hepatocytes. Journal of Viral Hepatitis. 2008;15:261–270. [PMC free article] [PubMed] [Google Scholar]
  • Yu Q, Zhang D, Walston M, Zhang J, Liu Y, Watson RR. Chronic methamphetamine exposure alters immune function in normal and retrovirus-infected mice. International Immunopharmacology. 2002;2:951–962. [PubMed] [Google Scholar]
  • Zhang XQ, Cui Y, Cui Y, Chen Y, Na XD, Chen FY, et al. Activation of p38 signaling in the microglia in the nucleus accumbens contributes to the acquisition and maintenance of morphine-induced conditioned place preference. Brain, Behavior, and Immunity. 2012;26:318–325. [PubMed] [Google Scholar]
  • Zhang L, Kitaichi K, Fujimoto Y, Nakayama H, Shimizu E, Iyo M, et al. Protective effects of minocycline on behavioral changes and neurotoxicity in mice after administration of methamphetamine. Progress in Neuro-Psychopharmacology & Biological Psychiatry. 2006;30:1381–1393. [PubMed] [Google Scholar]
  • 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. [PMC free article] [PubMed] [Google Scholar]
  • Zorick T, Nestor L, Miotto K, Sugar C, Hellemann G, Scanlon G, et al. Withdrawal symptoms in abstinent methamphetamine-dependent subjects. Addiction. 2010;105:1809–1818. [PMC free article] [PubMed] [Google Scholar]
statistics
Morty Proxy This is a proxified and sanitized view of the page, visit original site.