ArticlePDF AvailableLiterature Review

Abstract and Figures

Over the past two decades methamphetamine (MA) abuse has seen a dramatic increase. The abuse of MA is particularly high in groups that are at higher risk for HIV-1 infection, especially men who have sex with men (MSM). This review is focused on MA toxicity in the CNS as well as in the periphery. In the CNS, MA toxicity is comprised of numerous effects, including, but not limited to, oxidative stress produced by dysregulation of the dopaminergic system, hyperthermia, apoptosis, and neuroinflammation. Multiple lines of evidence demonstrate that these effects exacerbate the neurodegenerative damage caused by CNS infection of HIV perhaps because both MA and HIV target the frontostriatal regions of the brain. MA has also been demonstrated to increase viral load in the CNS of SIV-infected macaques. Using transgenic animal models, as well as cultured cells, the HIV proteins Tat and gp120 have been demonstrated to have neurotoxic properties that are aggravated by MA. In addition, MA has been shown to exhibit detrimental effects on the blood-brain barrier (BBB) that have the potential to increase the probability of CNS infection by HIV. Although the effects of MA in the periphery have not been as extensively studied as have the effects on the CNS, recent reports demonstrate the potential effects of MA on HIV infection in the periphery including increased expression of HIV co-receptors and increased expression of inflammatory cytokines.
Content may be subject to copyright.
REVIEW
Methamphetamine toxicity and its implications
during HIV-1 infection
Peter S. Silverstein &Ankit Shah &Raeesa Gupte &
Xun Liu &Robert W. Piepho &Santosh Kumar &
Anil Kumar
Received: 23 May 2011 / Accepted: 22 June 2011 / Published online: 23 July 2011
#Journal of NeuroVirology, Inc. 2011
Abstract Over the past two decades methamphetamine
(MA) abuse has seen a dramatic increase. The abuse of MA
is particularly high in groups that are at higher risk for HIV-1
infection, especially men who have sex with men (MSM).
This review is focused on MA toxicity in the CNS as well as
in the periphery. In the CNS, MA toxicity is comprised of
numerous effects, including, but not limited to, oxidative
stress produced by dysregulation of the dopaminergic system,
hyperthermia, apoptosis, and neuroinflammation. Multiple
lines of evidence demonstrate that these effects exacerbate the
neurodegenerative damage caused by CNS infection of HIV
perhaps because both MA and HIV target the frontostriatal
regions of the brain. MA has also been demonstrated to
increase viral load in the CNS of SIV-infected macaques.
Using transgenic animal models, as well as cultured cells, the
HIV proteins Tat and gp120 have been demonstrated to have
neurotoxic properties that are aggravated by MA. In addition,
MA has been shown to exhibit detrimental effects on the
bloodbrain barrier (BBB) that have the potential to increase
the probability of CNS infection by HIV. Although the effects
of MA in the periphery have not been as extensively studied
as have the effects on the CNS, recent reports demonstrate the
potential effects of MA on HIV infection in the periphery
including increased expression of HIV co-receptors and
increased expression of inflammatory cytokines.
Keywords Methamphetamine .HIV-1 .Dopamine .CNS .
Immune system
Through the 1990s and well into the first decade of the
twenty-first century, methamphetamine (MA) abuse has
been responsible for an increase in admissions to publicly
funded drug treatment programs (Gonzales et al. 2010). In
2005, it was estimated that 10.4 million people of age
12 years and over had used MA at least once in their life,
and over half a million people reported current use of the
drug (Office of Applied Studies 2007). Furthermore,
hospital emergency department visits related to MA abuse
exhibited a dramatic increase between 1995 and 2002. The
problem of MA abuse is not limited to the USA, since
significant use is also reported in Eastern Europe and
Southeast Asia (Degenhardt et al. 2010). The abuse of MA
is particularly high among men who have sex with men
(MSM; Gonzales et al. 2009). In this group, MA use is
associated with an increase in high-risk sexual activities
such as decreased condom use and increased numbers of
sexual partners (Gonzales et al. 2010; Shoptaw and Reback
2007). Studies have indicated that the HIV incidence is
doubled or tripled in MSM who use amphetamines
compared with MSM who do not use drugs (Buchacz et
al. 2005). In addition, resistance to antiretroviral drugs has
been shown to be positively correlated with MA abuse in
MSM (Gorbach et al. 2008). Although research studies can
be confounded by factors such as the abuse of multiple
drugs (i.e., polydrug abuse) and differential levels of sexual
activity and exposure, the study by Gorbach et al.
accounted for some of these factors and still demonstrated
a significant correlation between MA abuse and the
acquisition of drug-resistant virus.
This review focuses on MA toxicity in both the CNS and
the periphery, especially in the context of HIV-1 infection.
The mechanisms responsible for the toxicity, as well as the
mechanisms involved in mediating interactions between the
effects of MA and HIV infection will be discussed.
P. S. Silverstein (*):A. Shah :R. Gupte :X. Liu :
R. W. Piepho :S. Kumar :A. Kumar
Division of Pharmacology and Toxicology, School of Pharmacy,
University of Missouri-Kansas City,
Kansas City, MO 64108, USA
e-mail: Silversteinp@umkc.edu
J. Neurovirol. (2011) 17:401415
DOI 10.1007/s13365-011-0043-4
Furthermore, evidence that MA abuse affects the outcome
of HIV-1 infection will be reviewed.
The dopaminergic system and methamphetamine
MA is very similar in chemical structure to the neurotrans-
mitter dopamine, and this is thought to be the basis for
many of the effects of this drug (Riddle et al. 2006; Sulzer
et al. 2005). Early work that demonstrated the role of
dopamine in MA-mediated effects utilized various methods
of regulating dopamine synthesis, metabolism and disposi-
tion, followed by a determination of the effects of these
agents on MA-mediated neurotoxicity. In the neostriatum of
MA-treated rats, it was reported that the activity of tyrosine
hydroxylase (TH), the rate-limiting enzyme in dopamine
biosynthesis, was inhibited (Gibb and Kogan 1979). It was
also demonstrated that a-methyl-r-tyrosine, an inhibitor of
TH, abrogates the MA-induced decrease in TH activity in
the neostriatum. The protective effect of α-methyl-r-
tyrosine was abrogated by administration of L-DOPA,
which served to restore cytoplasmic dopamine levels. The
depressive effects of MA are even more pronounced on
tryptophan hydroxylase (TPH), the enzyme responsible for
serotonin synthesis (Hotchkiss and Gibb 1980; Schmidt et
al. 1985). Experiments utilizing the administration of a-
methyl-r-tyrosine and L-DOPA strongly suggested that the
effects of MA on TPH were mediated by the dopaminergic
system (Hotchkiss and Gibb 1980; Schmidt et al. 1985).
The effects of MA on the serotonergic system are blocked
by compounds that inhibit the effects of MA on the
dopaminergic system (e.g., α-methyl-r-tyrosine). However,
agents that specifically target the serotonergic system (e.g.,
5-HT uptake inhibitors) have no influence on MA-induced
effects on the dopaminergic system. This suggests the
overriding importance of dopamine in regulating both
systems in response to MA.
Components of the dopaminergic system
Dopamine receptors
Dopamine receptors comprise a family of G protein-
coupled receptors that are prominent in the CNS, as well
as in certain cell types and tissues in the periphery. In
addition to the CNS, dopamine receptors have also been
reported to be expressed in the cardiovascular and renal
systems, as well as in the retina and adrenal glands (Ozono
et al. 1997; Pivonello et al. 2004). In the cardiovascular
system, dopamine affects contractility (Ruffolo and Messick
1985) and vasodilation (Munch et al. 1991), it controls renal
filtration (Olsen 1998), whereas in the retina it is responsible
for circadian rhythm and retinal development (Witkovsky
2004). Dopamine receptors are also expressed on various
immune cells in the body. D
1
D
5
receptor expression on
human lymphocytes has been demonstrated through the use
of radioligand binding (Amenta et al. 1999; Ricci and
Amenta 1994; Ricci et al. 1998; Ricci et al. 1997; Ricci et al.
1999).
The availability of antagonists for the dopamine recep-
tors D
1
and D
2
has facilitated the identification of these
proteins as key players in the effects of MA. Sonsalla et al.
(Sonsalla et al. 1986) demonstrated that both a D
1
antagonist, SCH23390, and a D
2
antagonist, sulpiride,
decreased the MA-induced effects on the striatal dopami-
nergic parameters. However, only the D
1
antagonist was
effective at reducing the effects of MA on the striatal
serotonergic parameters. The ability of the D
1
antagonist
SCH23390 to protect against MA-induced neurotoxicity
has also demonstrated (Angulo et al. 2004; Xu et al. 2005).
The report by Xu et al. also showed that raclopride, a D
2
antagonist, as well as SCH23390, ameliorated MA-induced
astrogliosis, depletion of the dopamine transporter (DAT),
and cell death. Jayanthi et al. (Jayanthi et al. 2005)
demonstrated the induction of the FasL-Fas death pathway
by MA treatment. Treatment with SCH23390 reduced the
level of TUNEL staining and blocked the MA-induced
cleavage of caspase 8. Taken together, these results strongly
suggest a role for both D
1
and D
2
receptors in the
modulation of MA-induced neurotoxicity.
Dopamine transporters
Dopamine transporters and their regulation will obviously
have profound effects upon the disposition of dopamine.
The effects of MA on dopamine transporters have been
examined by numerous investigators. Early work by
Wagner et al. (Wagner et al. 1980) showed that multiple
doses of MA were followed by striatal dopamine depletion
and loss of dopamine uptake. Other work that demonstrated
the importance of the DAT in mediating MA toxicity was
performed by Schmidt and Gibb (Schmidt and Gibb 1985).
These investigators demonstrated that amfonelic acid, an
inhibitor of DAT, prevented the MA-induced reduction of
TH activity in rats. Similarly, Marek et al. demonstrated that
dopamine uptake inhibitors, including amfonelic acid, were
able to ameliorate the neurotoxic effects of MA (Marek et
al. 1990). Additional confirmation of the role of DAT in
MA-induced toxic effects was provided using knock-out
mice that lacked DAT expression (Fumagalli et al. 1998). In
homozygous KO mice, the lack of DAT expression
abrogated the MA-induced depletion of striatal DA levels.
Animals heterozygous for DAT expression displayed levels
of striatal DA that were intermediate between the wild-type
and DAT knock-out animals. Similar differences between
wild-type and knock-out mice were observed in terms of
402 J. Neurovirol. (2011) 17:401415
levels of MA-induced astrogliosis and oxygen radical
production. Taken together, these results confirm a central
role for DA and DAT in various MA-induced effects.
Vesicular monoamine transporter 2 (VMAT-2) is another
transporter that is important in terms of DA disposition
[reviewed in (Fleckenstein et al. 2009)]. In addition to
expression in neurons, VMAT-2 is expressed in β-cells of
the pancreas (Saisho et al. 2008) and mononuclear cells
(Anlauf et al. 2006; Tang et al. 2003). The function of this
transporter is to package monoamine transmitters such as
DA into synaptic vesicles in neurons. Mice that are
heterozygous for VMAT-2 expression were utilized to
demonstrate that altered vesicular transport of DA resulted
in increased neurotoxicity associated with altered distribu-
tion of DA and its metabolites (Fumagalli et al. 1999). MA-
treatment of mice was demonstrated to reduce binding of a
VMAT-2 ligand to its transporter (Hogan et al. 2000). The
role of VMAT-2 in DA transport and MA-associated
toxicity was confirmed by (Brown et al. 2002)who
demonstrated that the DA transport inhibitor methylpheni-
date (MPD) caused a redistribution of VMAT-2 from the
fraction associated with the plasmalemmal membrane to the
fraction associated with the cytoplasmic vesicles. Multiple
administrations of MA have been demonstrated to cause a
decrease in VMAT-2 immunoreactivity in the vesicular
subcellular fraction, while causing little change in the
plasmalemmal membrane or whole synaptosomal fractions
(Sandoval et al. 2003). Treatment with MPD after MA
treatment attenuated the MA-induced decrease in vesicular
DA uptake and vesicular DA content without altering the
total striatal dopamine content (Sandoval et al. 2003).
Further evidence for the importance of VMAT-2 in
modulating MA toxicity comes from a report that utilized
lobeline, an alkaloid that had been shown to inhibit MA-
induced behavioral characteristics (Eyerman and Yamamoto
2005). These investigators found that although lobeline did
not affect the MA-induced increase in extracellular striatal
DA, MA-induced hyperthermia and the loss of VMAT-2
immunoreactivity were ameliorated by the alkaloid. As
described above, MA, through its effects on DAT and
VMAT, has a major impact on the expression of transporters
modulating dopamine disposition. It is undoubtedly
through this mechanism that MA exerts at least a portion
of its effects on the CNS.
As can be seen from the brief review above, the
dopaminergic system plays a key role in mediating MA-
induced neurotoxicity. The effects of MA can be modulated
by alteration of DA levels, either through affecting DA
metabolism or it can be mediated through DAT or VMAT-2.
D
1
or D
2
receptors are also capable of affecting MA-
induced effects. Taken together, these results strongly argue
for a central role of the dopaminergic system in mediating
MA toxicity.
Methamphetamine-associated toxicity
The toxicity of MA has been studied in rodents, monkeys,
and humans at behavioral, cellular, and molecular levels.
Studies in rodent model systems
Rodent models have been used to study the effects of MA
at the cellular and molecular levels. In addition, these
models have been used to study MA-mediated hyperthermia
and oxidative stress.
a. Effects of Methamphetamine at the Cellular and
Molecular Levels. Siegel and colleagues have shown that
long-term MA exposure increases the expression of
muscarinic acetylcholine receptors in the hippocampus,
resulting in impaired novel location recognition in female
but not in male mice (Siegel et al. 2010). In contrast to
adult mice, exposure of adolescent mice to MA results in
impaired novel object recognition; the level of impairment
observed is equal in both male and female mice (Siegel et
al. 2011). In addition, MA exposure of adolescent mice
does not affect anxiety-like behavior, sensorimotor gating,
and contextual and cued fear conditioning. As seen in
adolescent mice and unlike that observed in adult mice,
juvenile mice do not show a gender difference in terms of
MA-induced neurotoxicity or increase in body temperature
(Dluzen et al. 2010).
Using in vitro cell cultures as well as a mouse model,
Narita and colleagues demonstrated that MA induces long-
lasting activation of astrocytes in the cingulate cortex and
nucleus accumbens through the protein kinase C pathway
(Narita et al. 2005). In MA-treated mice, the ratio of
activated microglia to non-activated microglia increases
from day 1 to day 7, as opposed to the control groups in
which the ratio is below 0.15. MA-induced glial activation
has been suggested to occur through inflammatory cytokines
based on the fact MA induces IL6 in the hippocampus and
striatum and it induces TNF-αin the hippocampus and frontal
cortex (Goncalves et al. 2008).
Genc and colleagues have shown that MA causes
cytotoxicity in rat oligodendrocyte cultures through the
induction of apoptotic cell death, which is evident from the
expression of pro-apoptotic proteins (Genc et al. 2003).
Exposure of cultured adult rat ventricular cardiomyocytes
(RVC) to acute MA treatment increases the size of RVC,
while chronic MA treatment decreases its size (Ruffolo and
Messick 1985). Furthermore, while acute MA exposure
increases microtubule (MT) assembly, chronic MA exposure
causes a reduction in MT assembly.
Overall, these studies have demonstrated that MA can have
a range of toxic effects on cells by induction of pathways that
include activation, cytokine induction, and apoptosis.
J. Neurovirol. (2011) 17:401415 403
b. The Effects of Methamphetamine on Oxidative Stress.
Using a rat model, Acikgoz et al. demonstrated that acute
repeated administration, as well as chronic administration,
of MA causes an increase in SOD activity along with an
increase lipid peroxidation in the striatum (Acikgoz et al.
1998). Fluorescein derivatives, along with cell cultures
derived from VMAT-2 wild-type and mutant mice, were
used to demonstrate that MA exposure caused an increase in
ROS (Larsen et al. 2002). It was also demonstrated that the
loss of VMAT-2 expression was associated with higher levels
of ROS and increased levels of cysteinyl-DA, a metabolite
associated with oxidized DA. Using a rat model of MA-
induced neurotoxicity, it was demonstrated that MA admin-
istration at toxic levels resulted in an increase in protein-
cysteinyl DA as a result of an increase in the oxidative
metabolites of DA (LaVoie and Hastings 1999). Using mice
subjected to repeated doses of MA, Jayanthi et al.
demonstrated that levels of glutathione peroxidase, catalase,
and Cu/Zn-SOD were decreased in the striatum of these
animals (Jayanthi et al. 1998). Significantly, this correlated
with an increase in the products of lipid peroxidation. MA
has also been shown to have a similar effect on most, but not
all, peroxiredoxins in rat brain (Chen et al. 2007). In a model
similar to that of Jayanthi et al., repeated injections of MA
produced elevated levels of protein carbonyls and thiobarbi-
turic acid reactive substances, both of which are markers of
oxidative stress, in brain (Gluck et al. 2001). Proteomic
analysis has also been utilized to confirm the upregulation
of enzymatic markers of oxidative stress in MA-treated rats
(Iwazaki et al. 2006). In another study, MA-induced toxicity
was found to be associated with an increase in protein bound
quinone and increased expression of quinone reductase
(Miyazaki et al. 2006). Furthermore, pre-treatment of animals
with an inducer of quinone reductase, which is known to be
protective against quinine-induced toxicity, protected against
MA-induced toxicity.
c. The Effects of Methamphetamine on Hyperthermia and
Microglial Activation. Several studies on MA-induced
hyperthermia in a rat model have been reported by Kiyatkin
and colleagues. MA induces both brain and body hyper-
thermia but brain hyperthermia is much stronger and more
rapid than body hyperthermia (Brown et al. 2003). Unlike
body hyperthermia, brain hyperthermia is dramatically
enhanced at warm ambient temperatures, often resulting in
lethality in mice (Brown and Kiyatkin 2005). In general,
MA-induced brain hyperthermia causes damage to brain
cells, including neurons, glia, epithelial, and endothelial
cells (Kiyatkin 2005; Kiyatkin 2010; Kiyatkin et al. 2007).
MA-induced brain hyperthermia also causes acute glial
activation and edema. Furthermore, chronic MA-induced
brain and body hyperthermia, as well as acute MA
intoxication-induced brain hyperthermia have been corre-
lated with increased permeability of the BBB (Kiyatkin et
al. 2007; Sharma and Kiyatkin 2009). An earlier study
demonstrated that MA-induced hyperthermia induces heat
shock protein (HSP) (Kuperman et al. 1997), which is
consistent with the general phenomenon that hyperthermia
can induce the expression of heat shock proteins (HSPs). A
recent study confirms that acute MA intoxication in rats
causes induction of wide-spread HSP expression in neural
and glial cells, as well as in the cortex, hippocampus,
thalamus, and hypothalamus (Kiyatkin and Sharma 2011).
The induction of HSPs in these cells correlates with brain
hyperthermia, permeability of BBB, acute glial activation,
and brain edema. Although the induction of HSP is an
adaptive mechanism to counteract hyperthermia, it does not
counteract the damaging effects of oxidative stress, hyper-
thermia, and edema in rats.
In addition to hyperthermia, MA has been shown to
induce microglial activation (Kuhn et al. 2008; Thomas et
al. 2004a; Thomas et al. 2004b). Microglial activation was
determined by staining brain sections with isolectin B
4
, and
activation was found to be independent of hyperthermia.
Agents that changed dopamine disposition, such as L-DOPA
and reserpine, enhanced the effects of MA on microglial
activation but had no effects on microglial activation by
themselves. It is of particular interest that reserpine caused
hypothermia in mice, while L-DOPA caused hyperthermia.
However, attenuation of MA-induced microglial activation
by minocycline that resulted in reduction of IL-1a and IL-6
levels did not afford neuroprotection (Sriram et al. 2006).
Thus, although microglial activation is not dependent upon
hyperthermia, microglial activation alone could not account
for the observed neurotoxicity.
Studies in monkey model systems
Since acute high MA dosing regimens can lead to
considerable toxicity and even death in experimental
animals, a non-lethal chronic MA administration procedure
for the rhesus macaque that utilizes an escalating dose
protocol has been developed by Madden and colleagues
(Madden et al. 2005). This regimen produces several
behavioral and physiological effects, including decreased
food intake and increased cortisol excretion, which are
similar to MA-induced effects in humans. In vervet
monkeys, MA exposure has been correlated with the
oxidative stress that occurs during aging (Melega et al.
2007). They showed that, after 1 month of MA treatment,
there is an increase in iron levels in the substantia nigra pars
reticulata and the globus pallidus, along with a concurrent
increase in ferritin-immunoreactivity and a decrease in
tyrosine hydroxylase-immunoreactivity in the substantia nigra.
While the increase in tyrosine hydroxylase-immunoreactivity
is observed after 1.5 years of simulated MA abuse, iron levels
404 J. Neurovirol. (2011) 17:401415
of the adult MA-exposed animals (age 59 years) are
comparable with those of drug-naive, aged animals (19
22 years). In a subsequent study, these investigators demon-
strated that multiple doses of MA administered to socially
housed vervet monkeys cause a progressive increase in
abnormal behavior and a decrease in social behavior (Melega
et al. 2008). In this study, the vervet monkeys exhibited an
increase in anxiety on no injectiondays and a decrease in
aggression was observed throughout the study. Finally, since
some behavioral and pharmacological patterns of chronic MA
abuse and schizophrenia are similar, a primate animal model
of schizophrenia has been established using chronic phency-
clidine (PCP) monkeys. An acute MA injection to the
chronically treated PCP monkeys exacerbated the behavioral
effects of PCP, suggesting that these monkeys can be used as
a primate model of schizophrenia (Mao et al. 2008).
Clinical studies
Clinical studies have been used to study the effects of MA
at behavioral, cellular, and molecular levels. In addition,
clinical investigations have also focused on MA-mediated
oxidative stress.
a. The Effects of Methamphetamine at the Cellular and
Molecular Levels. Chung and colleagues have shown that
the chronic use of MA in humans decreases the cerebral
blood flow in subcortical and dorsal cortical brain regions
(Chung et al. 2010). However, its binge use is associated
with severe neurotoxicity to the monoaminergic neurotrans-
mitter system as a result of long-term changes in both
global and regional blood flows. This produces a pattern of
hypoperfusion, which resembles the pattern of atypical
Parkinsons disease. Thus, binge use of MA has been
employed as an experimental model of Parkinsons disease
in animals (Garcia de Yebenes et al. 2000; Romero et al.
2006).
To understand MA-induced neuronal changes in humans,
several imaging techniques have been utilized as described
below. Iyo and colleagues used single photon emission
computed tomography, magnetic resonance spectroscopy,
and positron emission tomography to investigate MA-
mediated psychosis (Iyo et al. 2004). In MA users, these
studies have shown (1) a high incidence of multiple patchy
deficits in cerebral blood flow, (2) a significantly reduced
ratio of creatine plus phosphocreatine/choline-containing
compounds in the brain, and (3) a decrease in the density of
dopamine transporter in the nucleus accumbens and
caudate/putamen. These effects correlate with the duration
of MA use and the severity of residual psychotic symptoms.
High-resolution genetic resonance imaging (Gorbach et al.)
and surface-based computational image analyses in MA
abusers have revealed severe gray-matter deficits in the
cingulate, limbic, and paralimbic cortices, as well as a
decrease in hippocampal volumes and white-matter hyper-
trophy (Thompson et al. 2004). In addition, MA abuse
causes a selective cerebral deterioration resulting in
impaired memory and damage to the medial temporal lobe
and cingulate-limbic cortex. Furthermore, findings from
functional magnetic resonance imaging (fMRI) suggest a
relationship between decision-making dysfunction and
neural activation in different prefrontal areas (Paulus et al.
2002). MA abusers show a decrease in the activation of
dorsolateral prefrontal cortex and fail to activate the
ventromedial cortex during the two-choice prediction task
compared with the two-choice response task.
A recent finding from a study that focused on prenatal
exposure to MA using neuroimaging suggests that MA
exposure in utero is toxic to dopamine-rich basal ganglia
regions (Roussotte et al. 2010). High levels of MA
exposure during pregnancy have been associated with
increased lethargy and physiological stress; first trimester
produces more stress, while third trimester results in
lethargy, poorer quality of movement, and hypotonicity
(LaGasse et al. 2011; Smith et al. 2008).
b. The Effects of Methamphetamine on Oxidative Stress.
MA-induced oxidative stress has also been studied using
human cell cultures. Cubells and colleagues have demon-
strated MA-induced neurite damage and ROS production in
neuronal cultures using differential interference contrast and
fluorescence techniques, respectively (Cubells et al. 1994).
MA exposure has also been shown to increase the
permeability of brain microvascular endothelial cell
(BMVEC) monolayers by decreasing the expression of tight
junction (TJ) proteins and increasing ROS formation
(Ramirez et al. 2009). It has also been shown that MA
modulates TJ expression, leading to decreased transendothe-
lial resistance and enhanced transendothelial migration of
immunocompetent cells across the BBB (Mahajan et al.
2008). Furthermore, N27 dopaminergic neuronal cells have
been used to demonstrate the role of cathepsin-D in MA-
induced autophagy and apoptosis as a result of increased
oxidative stress (Kanthasamy et al. 2006).
c. The Effects of Methamphetamine on Behavior. Al-
though acute use of MA is known to elevate energy and
alertness, its chronic use is associated with increases in
psychosis, anxiety, and depression(Glasner-Edwards et al.
2010; Gonzales et al. 2009; Gonzales et al. 2010; Rawson
et al. 2002;Zwebenetal.2004). Furthermore, MA
intoxication is associated with violent, agitated, and suicidal
behaviors (Newton et al. 2004). The severity of MA-mediated
effects is related to the quantity and frequency of MA
administration. The route of administration as well as
individual genetic differences has an effect upon behavioral
J. Neurovirol. (2011) 17:401415 405
outcomes. The symptoms of MA-induced psychosis are
similar to the symptoms of schizophrenia, such as paranoid
ideation, delusions, and auditory and visual hallucinations
(Zweben et al. 2004). Although in most users, psychosis
occurs temporarily and is typically abolished within a week
of abstinence, it may persist for several months in a small
fraction of users.
Taken together (overview presented in upper half of
Fig. 1and Table 1), the studies described above demon-
strate that MA exposure has a wide range of toxic effects in
both the CNS and the periphery. These toxic effects include
brain and body hyperthermia, induction of apoptotic path-
ways, increased levels of markers of oxidative stress, as
well as behavioral effects (Krasnova and Cadet 2009). Even
in the absence of additional agents, the toxic effects of MA
are rather significant. In the context of viral infection, they
will show themselves to be even more deleterious.
Methamphetamine and HIV
Effects of methamphetamine on the CNS
Globally, at least 33.3 million people are estimated to be
living with HIV as of 2009. At least 2030% of the patients
infected with HIV-1 will eventually be diagnosed with HIV-
associated dementia (HAD; McArthur et al. 1993; Nath et
al. 2000; Navia et al. 1986a; Navia et al. 1986b). The
neurotoxic effects of HIV-1 are primarily attributed to its
ability to readily penetrate into the central nervous system
(CNS) early during the course of infection. Deficiency in
the functionality of dopaminergic neurons has been ob-
served to be associated with early stage HIV-1 infection
(Berger et al. 1994). Although the introduction of highly
active antiretroviral therapy (HAART) has significantly
reduced the incidence of HAD (Clifford 2008), milder
neurotoxicity, including minor cognitive motor disorders
and HIV-associated neurodegenerative disorders (HAND)
have increased in incidence (Antinori et al. 2007). Many
anti-retroviral drugs fail to penetrate the blood brain barrier
(BBB), thus making it difficult to treat HAND patients
(Thomas 2004). HIV-associated neurotoxicity is primarily
thought to be mediated by the neurotoxins released from
infected cells, mostly resident microglia, after migration of
the infected cells through the BBB (Gendelman and
Meltzer 1989; Meltzer and Gendelman 1992; Meltzer et
al. 1990). The frontostriatal regions of the brain are highly
vulnerable to this so-called Trojan Horsemechanism by
which HIV-1 penetrates the CNS (Itoh et al. 2000; Reyes et
al. 1991). MA also targets these frontostriatal regions by
increasing DA and glutamate transmission, which further
leads to neuronal damage and cell death (Davidson et al.
2001; Langford et al. 2003; Stephans and Yamamoto 1994;
Wilson et al. 1996). Multiple models for MA-mediated
neurotoxicity have been proposed (Cadet and Krasnova
2007; Reiner et al. 2009). However, MA-mediated neuronal
damage is chiefly attributed to depletion of dopamine and
5-HT (Cadet et al. 1994; Wagner et al. 1980), dopamine
transporters (DAT) (Xu et al. 2005), and vesicular mono-
amine oxidase (Mao et al.) in the corpus striatum (Frey et
al. 1997).
The effects of methamphetamine and viral proteins on CNS
toxicity
In an early study, it was demonstrated that treatment with
MA and Tat increased neuronal cell death when human fetal
neurons were exposed to these agents in culture (Magnuson
et al. 1995). Based upon their earlier studies, along with
other relevant data, Nath et al. proposed that dopaminergic
MA
Non-HIV
HIV
Impaired novel location recognition
Activated microglia and apoptotic cell death
Oxidative stress, temperature, and BBB permeability
Abnormal behavior, anxiety, and oxidative stress
Social behavior
Rodent
Monkey
Human Psychosis, anxiety, and depression
Lethargy and physiological stress during pregnancy
Oxidative stress and BBB permeability
Immune
system
Effects
CNS
Effects
Viral replication
HIV-1 co-receptor expression
Altered cytokine secretion
Impaired antigen presentation
Oxidative stress
Dopamine levels
Neuronal apoptosis
Behavioral sensitization
Oxidative stress
Altered cytokine expression
BBB permeability
Fig. 1 Schematic of an over-
view of the effects of MA and
MA in the context of HIV
infection. The upper portion of
the figure focuses on the effects
of MA in rodent and monkey
model systems, as well as those
results derived from clinical
studies on humans. The bottom
portion of the figure focuses on
the effects of MA in the context
of HIV infection
406 J. Neurovirol. (2011) 17:401415
Table 1 Overview of the effects of MA and MA +HIV/Tat or gp120 in different model systems
Treatment Tissue/cell type Effect Reference
MA Rat neostiatum Tyrosine hydroxylase Gibb and Kogan 1979
Rat neostiatum Tryptophan hydroxylase Hotchkiss and Gibb 1980
Rat caudate DAT ;DAWagner et al. 1980
Rat body, rat brain HyperthermiaBrown et al. 2003
Brown and Kiyatkin 2005
Sharma and Kiyatkin 2009
Rat striatum Altered VMAT-2 localization Brown et al. 2002
Sandoval et al. 2003
VMAT-2Eyerman and Yamamoto 2005
Oxidative stress markers LaVoie and Hastings 1999
Oxidative metabolites
Quinone levels
Oxidative stress markers Iwazaki et al. 2006
Rat neurons Neurite degenerationCubells et al. 1994
Rat neural and glial cells HSP expression Kiyatkin and Sharma 2011
Mouse neurons Apoptosis Jayanthi et al. 2005
Mouse hippocampus Muscarinic acetylcholine receptor Siegel et al. 2010
Mouse hippocampus Impaired novel location recognition Siegel et al. 2011
Mouse hippocampus IL-6 ; TNF-αGoncalves et al. 2008
Mouse striatum/hippocampus DA levels ; 5-HT levels Fumagalli et al. 1998
Mouse striatum Altered VMAT-2 ligand binding Hogan et al. 2000
Mouse striatum Microglial activation Thomas et al. 2004a
Kuhn et al. 2008
Mouse neuron Lipid peroxidation Jayanthi et al. 1998
Mouse HSP expression Kuperman et al. 1997
Vervet monkey Oxidative stress Melega et al. 2007
Food intake Melega et al. 2008
Social behavior
Anxiety
Rhesus macaques Food intake Madden et al. 2005
Cortisol excretion
Human brain Cerebral blood flow Chung et al. 2010
DAT Iyo et al. 2004
Neural activation Paulus et al. 2002
Hippocampal volumeThompson et al. 2004
Human dendritic cells TNF-α, IL-1β, CCR5 , IL-8 Mahajan et al. 2006
p38 MAPK phosphorylation
PI3K phosphorylation
CXCR4 , CCR5 Nair et al. 2006
p38 MAPK phosphorylation
MIP-1α, MIP-1β, RANTES Nair and Saiyed 2010
Human MDM CCR5 , IFN-αLiang et al. 2008
Mixed neuron/astrocyte cultures MMPs Conant et al. 2004
Human brain Endothelial cells Tight junction proteins Ramirez et al. 2009
Human Psychosis , anxiety , depression Rawson et al. 2002
Zweben et al. 2004
Depression Glasner-Edwards et al. 2010
Psychosis Gonzales et al. 2009
Anxiety , depression Gonzales et al. 2010
MA+HIV HIV-1 transgenic rats Behavioral sensitization Liu et al. 2009
J. Neurovirol. (2011) 17:401415 407
activation-mediated depletion in dopamine levels impaired
the function of the DA transporter and that the resultant
alterations in DA reuptake (Nath et al. 2000)were
responsible for the toxic effects of MA and HIV-1 on
dopaminergic neurons. Later, various MRS studies (Chang
et al. 2005;Schweinsburgetal.2005)showedthatMA
abuse by HIV-positive individuals aggravated damage in the
brain in terms of N-acetyl aspartate reduction.
Multiple studies have been undertaken that focus on the
molecular mechanisms involved in the cross-talk between
the viral proteins and MA. Studies by Maragos et al.
revealed altered dopamine levels due to the combined
effects of MA and HIV-1 Tat (Maragos et al. 2002). Using
SpragueDawley rats treated with threshold doses of Tat
and MA, they demonstrated greater depletion in the striatal
DA levels of rats treated with both Tat and MA when
compared with the depletion of DA levels upon treatment
with either MA or HIV Tat. Using neuronal cultures, they
also showed that MA and Tat treatment resulted in higher
levels of cell death and mitochondrial dysfunction as
compared with either agent alone. They also showed that
Tat and MA together can cause further decreases in the
overflow of dopamine as compared with either treatment
alone in the striatal regions of the rats. This suggests that
both the DA levels and the dynamics of DA release in the
striatum are affected by the interaction of MA with HIV-1
proteins. These alterations of the DA levels and activation
could be responsible for basal ganglia dysfunction in MA-
abusing HIV-infected patients (Cass et al. 2003). Turchan et
al. demonstrated synergistic toxicity of gp120/Tat with MA
that resulted in neuronal cell death and alteration of
mitochondrial membrane potential (Turchan et al. 2001).
These findings suggested the possibility that oxidative
stress may play a role in the synergy between MA and HIV.
Increased oxidative stress is found to be associated with
HIV-associated neuroinflammation. Banerjee et al. showed
that intrastriatal MA injection in mice resulted in synergistic
interactions between MA and HIV that were mediated
through oxidative stress (Banerjee et al. 2010). Mice treated
with MA following gp120 or Tat injections showed high
levels of oxidative stress markers such as malonyl dialde-
hyde (MDA) and protein carbonylation along with higher
lipid peroxidation in the brain. In addition, the presence of
both agents resulted in levels of antioxidant enzymes like
GSH and GPx that were significantly decreased when
compared with either treatment alone. The involvement of
oxidative stress was demonstrated through the use of the
antioxidant N-acetyl cysteine amide, which prevented the
disruption of mitochondrial potential that was caused by
MA+Tat or MA+gp120. Flora et al. highlighted the role of
redox-sensitive pathways in the combined effects of MA
and HIV-1 Tat (Flora et al. 2003; Flora et al. 2002).
Interestingly, the intrahippocampal injection of Tat and MA
in mice showed increased activity of transcription factors
associated with oxidative stress particularly in the cortical,
striatal, and hippocampal regions of the brain. The
transcription factors NFκB, AP-I, and CREB showed
increased DNA binding activity in the hippocampal and
cortical regions of mice treated with Tat and MA as
compared with either substance alone. The increase in the
Table 1 (continued)
Treatment Tissue/cell type Effect Reference
Human immature dendritic cells Adhesion protein (galectin-1, filamin 1 )Reynolds et al. 2007
Human PBMC Peroxiredoxin6 , HSP70p5 ,vimentinReynolds et al. 2009
Human MDM Viral replication Liang et al. 2008
T cells Viral replication Toussi et al. 2009
Dendritic cells Viral replication Reynolds et al. 2007
MA+SIV Macaque Viral load in brain Marcondes et al. 2010
SOD , GST Pendyala et al. 2011
MA+gp120 Mouse brain Oxidative stress markers Banerjee et al. 2010
Protein carbonylation
Transgenic mice Behavioral changes Roberts et al. 2010
Human BMEC Z0-1 , claudin 3/5 , JAM-2 Mahajan et al. 2008
Human fetal brain cells Cell death Turchan et al. 2001
Alteration of mitochondrial membrane
MA+Tat Rat striatum MCP-1 ,IL-1α, IL-1β, TNF-αTheodore et al. 2006a,b
Rat brain Altered dopamine levels Maragos et al. 2002
Mouse striatum TNF-α, lipid peroxidation , AP-1 Flora et al. 2002
Oxidative stress Flora et al. 2003
Mixed neuron/astrocyte cultures MMP Conant et al. 2004
408 J. Neurovirol. (2011) 17:401415
DNA binding activity of the transcription factors further led
to increased expression of IL-1β, TNF-α, and ICAM-1,
particularly in mouse striatum. In a later report, Langford et al.
extended these findings and showed that the combination of
HIV-1 Tat and MA can induce oxidative stress and alter
mitochondrial membrane calcium potentials, which can
further result in neuronal cell death (Langford et al. 2004).
High levels of various inflammatory cytokines are
associated with the toxicities observed in neuroinflamma-
tion. In particular, increased expression of TNF-αis
positively correlated with HAD (Glass et al. 1993). Flora
et al. showed increased expression of TNF-αin the brains
of mice treated with intrahippocampal Tat injections
following IP MA administration. In addition to TNF-α
induction in various regions of brain like frontal cortex,
corpus striatum, hippocampus, and cerebellum, elevated
levels of IL-1βand ICAM-1 were observed in the same
regions. The increase in these genes was found to be
associated with increased oxidative stress signaling. Be-
cause TNF-αand IL-1βalso act as pro-inflammatory
cytokines, the toxicities produced by MA and Tat together
could prove to be a double-edged swordof inflammation
and oxidative stress (Flora et al. 2003). Theodore et al.
(Theodore et al. 2006a) confirmed the previous findings by
using TNF-αR1 and TNF-αR2 double knockout mice. In
the DKO mice, Tat+ MA failed to deplete the DA levels as
compared with the depletion observed in Tat+ MA-treated
WT animals. The induction of TNF-αwas also found to
result in increased hippocampal neuron loss. Increased
levels of the pro-inflammatory cytokines MCP-1, TIMP-1,
and IL-1αwere found in a cytokine array prepared from rat
striatum treated with MA and Tat as compared with either
treatment alone. Furthermore, MCP-1 KO mice did not
show the depletion of DA observed in the combined
treatments as compared with the treatments by a single
agent (Theodore et al. 2006a,2006b). Together all these
findings provide strong evidence for the role of cytokines in
mediating the interactions observed between MA and HIV
infection in terms of increasing neurodegeneration.
Bloodbrain barrier integrity is essential for maintenance
of brain homeostasis. Tight junction proteins are a critical
component responsible for maintaining the high level of
impermeability of the BBB. Mahajan et al. showed that
HIV-1 gp120 and MA synergistically disrupt the BBB and
deplete various tight junction proteins such as ZO-1, JAM-2,
and Claudin-3/5 (Mahajan et al. 2008). Furthermore, studies
by Banerjee et al. showed that mice treated with MA and
viral proteins like gp120 or Tat had decreased levels of TJ
proteins like ZO-1 and occludin. Treatment with antioxidants
also demonstrated restoration of levels of TJ proteins. This
observation confirmed the role of oxidative stress in the loss
of BBB integrity Banerjee (Banerjee et al. 2010). Studies by
Conant et al. also demonstrated synergy between MA and
Tat in the induction of MMP levels in the striatum (Conant et
al. 2004).
Behavioral effects of methamphetamine in transgenic
rodents
Various transgenic rodent models have been developed that
simulate conditions of HIV/AIDS. Using HIV transgenic
rats, it was shown that MA increases behavioral sensitiza-
tion in these animals (Liu et al. 2009). HIV-1 transgenic rats
treated with MA showed increased behavioral sensitization
in terms of rearing and head movements when compared
with control transgenics that were not treated with MA. It
was also shown that D1R expression was higher in the
transgenic rats treated with MA. This cohort also showed
lower brain to body weight ratios, suggestive of brain
atrophy. Another study utilized a transgenic mouse model
expressing gp120 to demonstrate that MA-induced stereotypic
behavior and locomotion are significantly increased in HIV
transgenic mice (Roberts et al. 2010). This report, in
conjunction with the prior study, indicated possible behavioral
alterations that underscore the complexity associated with the
aggravating effects of MA abuse on HIV-associated CNS
toxicity. Taken together, these findings provide strong
evidence of increased CNS impairment in HIV-infected
individuals consuming illicit drugs.
The effects of methamphetamine on viral replication
One of the major contributing factors to HIV disease
progression due to MA is the increase in viral load due to
MA exposure. MA causes a dysregulation of dopamine
disposition, which has been shown to enhance viral
replication and also activate latent virus in T lymphocytes
(Rohr et al. 1999). This suggests the possibility that MA
may be able to increase HIV replication in the CNS.
Recently, the effect of MA administration on brain viral
load using macaques was determined. Rhesus macaques,
when infected with simian immunodeficiency virus (SIV),
can serve as a model of HIV infection in humans. Although
MA administration in monkeys produced no change in the
plasma viral load, the brain viral load was significantly
higher (Marcondes et al. 2010). The increased activation of
microglia and astrocytes in the brain demonstrates the toxic
potential of MA in HIV-1 infected individuals. Activation
of NK cells in the periphery and the expression of co-
receptor CCR5 on brain macrophages were also observed to
increase. MA was also shown to increase CD14+/CD16+
macrophages in brains of HIV-1-infected animals, and these
macrophages are known targets for SIV/HIV infection in
the brain. The MA-mediated increase in macrophage
activation and brain viral load suggests that MA may
exacerbate the CNS effects of HIV infection.
J. Neurovirol. (2011) 17:401415 409
There is evidence to suggest that MA use may result in
increased viral load in the periphery. Treatment of human
monocyte-derived macrophages with MA was able to
potentiate HIV reverse transcriptase activity in a dose
dependent manner, and the effects of MA could be
abrogated by blocking D1 receptors expressed on macro-
phages (Liang et al. 2008). Increased HIV replication was
also observed in immature dendritic cells treated with MA
prior to HIV-1 infection (Reynolds et al. 2007). Another
study examined both the in vitro and in vivo effects of MA
on HIV replication. In vitro, HIV replication was significantly
increased in monocytes and CD4
+
T cells treated with MA.
Viremia was also increased in vivo in mice transgenic for the
HIV provirus and human cyclin T1 as determined by p24
antigen production in splenocytes as well as viral RNA copy
numbers in serum. These effects were mediated by translo-
cation of NFКB p65 subunit into the nucleus and subsequent
transcription from the HIV-1 LTR (Toussi et al. 2009).
Because of the potential for infected monocytes to cross the
BBB, an increased viral load in the periphery may result in
higher viral loads in the CNS.
The effects of methamphetamine and HIV on the immune
system
HIV initially infects cells of the immune system and
subsequently invades and compromises various other
systems of the body. It is known to modulate various
immune functions such as activation of T cells and NK cells
as well as to disrupt cytokine balance. Because of its action
as a psychostimulant, the effects of MA in the context of
HIV infection have been primarily studied in the CNS.
However, dopamine receptors and transporters, which are
reported to mediate the effects of MA, are also expressed in
the periphery on various immune cells. It is therefore
relevant, and important, to study the effects of MA in the
immune system of HIV-infected individuals.
As mentioned previously, MA has been shown to
increase HIV replication in various immune cells such as
dendritic cells, monocytes, and CD4
+
T-cells (Reynolds et
al. 2007; Toussi et al. 2009). Several studies have also
documented the ability of MA to modulate other immune
functions. MA has been demonstrated to upregulate the
expression of the HIV co-receptor CCR5 on macrophages
while simultaneously suppressing the expression of the
anti-viral cytokine IFN-α(Liang et al. 2008). Microarray
studies on dendritic cells differentiated from normal human
PBMC and treated with MA showed altered gene expression
patterns. The expression levels of the pro-inflammatory
cytokines TNF-α,IL-1β, and IL-8 were upregulated, as was
the expression of CCR5. Phosphorylation of the signal
transduction molecule p38-MAPK was increased while
PI3K phosphorylation was decreased (Mahajan et al. 2006).
This study was followed up by performing proteomic
analysis using PBMC isolated from HIV-1 patients and
exposed to MA for 24 h. MA decreased expression of
HSP70p5 and peroxiredoxin 6 while increasing the expres-
sion of vimentin in these cells. HSP70 prevents vpr-induced
cell cycle arrest, whereas peroxiredoxins are antioxidants
that inhibit HIV-1 infection. Vimentin, on the other hand,
facilitates spread of HIV to adjacent cells (Reynolds et al.
2009). Proteomic analysis of MA-treated immature dendritic
cells infected with HIV-1 showed increased expression of
proteins that promote HIV adhesion, entry, and replication
such as galectin-1, PDI, filamin 1, and talin 1 (Reynolds et
al. 2007). MA may therefore act as a co-factor that promotes
HIV pathogenesis by increasing the susceptibility of cells to
viral invasion, activation of HIV transcriptional mechanisms,
and T cell depletion through apoptosis.
Dendritic cells (DC) are among the first lines of defense
against invading pathogens and consequently are among the
initial targets of HIV infection. MA treatment reduces the
expression of the mature DC marker CD83 that plays a role
in antigen presentation and T cell activation. It also
decreased secretion of the chemokines MIP-1α, MIP-1β,
and RANTES which can bind to the CCR5 co-receptor and
prevent entry of virus into cells (Nair and Saiyed 2010).
MA and gp120 synergistically upregulate DC ICAM-3 by
binding the non-integrin DC-SIGN. DC-SIGN is known to
promote HIV infection in the absence of CD4 or the HIV
co-receptors. In addition, MA also caused a dose-dependent
increase in the HIV-1 co-receptors CXCR4 and CCR5.
These effects were mediated by interaction of MA with D1
dopamine receptor and the phosphorylation of p38 MAPK
(Nair et al. 2006). Talloczy et. al investigated the effects of
pharmacologically relevant concentrations of MA on
antigen processing, presentation, and phagocytosis in
murine dendritic cells and macrophages (Talloczy et al.
2008). It was observed that MA induced alkalization in
acidic organelles and thereby impaired dendritic cell function
involving lysosomal degradation of foreign proteins. MA also
inhibited macrophage phagocytic function while promoting
fungal replication in macrophages. Therefore, the ability of
MA to disrupt the pH gradient in these cells was responsible
for loss of their respective functions.
A number of studies have also been carried out on non-
human primates, which serve as excellent model systems to
study HIV. Chronically SIV-infected rhesus macaques showed
changes in virushost interaction due to MA exposure.
Though plasma viral loads were not elevated, significant
changes were observed in immune cells. NK cell activation
was prominent in brain, blood, and lymphoid organs as
determined by degranulation and cytokine expression on the
cell surface (Marcondes et al. 2010). Oxidative stress is also
believed to play a pivotal role in chronic SIV and MA co-
morbidity. Proteomic plasma analyses of chronically infected
410 J. Neurovirol. (2011) 17:401415
macaques that were administered MA revealed significantly
elevated levels of the enzymes superoxide dismutase as well
as glutathione-S-transferase (Pendyala et al. 2011). This
suggests the utilization of these compensatory mechanisms
to combat oxidative stress.
Thus, MA has been implicated in exacerbating HIV-
induced effects in the CNS and periphery through number
of mechanisms promoting HIV replication and infectivity,
altering expression of important immune components,
impairing antigen presentation and elevating oxidative
stress (summarized in Table 1, Fig. 1).
Summary and future directions
A review of the literature shows that the chemical similarity
between MA and dopamine is thought to be the basis for
the toxic effects of this drug. Early work demonstrated that
treatment of rodents with MA results in altered expression
of many of the enzymes involved in dopamine biosynthesis.
Further work demonstrated that dopamine receptors and
dopamine transporters are key players in mediating the
effects of MA. The ability of MA to affect the function of
DAT and VMAT-2 causes an aberrant distribution of
dopamine and its metabolites. The altered distribution of
dopamine not only affects signaling, but it can also produce
oxidative stress. MA-abuse in humans has been demon-
strated to cause altered cerebral blood flow and severe gray
matter deficits in several regions of the brain. The use of
rodent models has also facilitated identification of brain
hyperthermia as one of the deleterious effects on the CNS
associated with MA abuse.
The frontostriatal regions of the brain that are most
susceptible to the deleterious effects of MA are also one of
the initial targets of HIV-1 infection. The HIV-1 viral proteins
Tat and gp120 interact with MA synergistically to increase
neuronal cell death, oxidative stress, and inflammatory
cytokine production by cells of the CNS. MA has been shown
to decrease tight junction proteins in the BBB such as ZO-1
and claudin-3/5. This may facilitateHIV-1 penetration into the
CNS. Using a macaque model of HIV-1 infection, one group
demonstrated that MA treatment of infected animals resulted
in an increase in the viral load in CNS.
More recently, MA has been shown to have the
potential to affect HIV infection in the periphery.
Treatment of human MDMs infected with HIV showed
increased levels of viral replication. HIV infection may
be exacerbated by the increased levels of inflammatory
cytokines and chemokines or increased levels of the
CCR5 co-receptor seen in MA-treated macrophages and
dendritic cells. Such increases may potentiate viral
replication in the periphery and thus increase the
potential for CNS infection.
Although the biological effects of MA abuse have been
extensively studied during the past two decades, much
remains to be explored regarding the effects of MA abuse
on HIV-1 infection. Although there have been some reports
regarding the effect of HIV and its associated proteins on
DAT, the effect of viral infection on the expression and
function of VMAT-2 is unknown. Because recent work has
demonstrated that both of these transporters affect dopa-
mine disposition, and both are also affected by MA, this
represents a major gap in our understanding of HIV-MA
interactions in the CNS. Another unanswered question
regards the potential role of dopamine receptors in affecting
HIV replication in microglial cells. These receptors have
already been demonstrated to increase HIV replication in
human MDMs treated with MA. Although the data is
limited, it has already been demonstrated that MA treatment
of macaques increases the CNS viral load. This raises the
question as to whether the primary mechanism is an
increase in the permeability of the BBB or an increase in
HIV replication in microglial cells. The effects of MA are
obviously not limited to the CNS, and the findings that
have shown an MA-induced increase in HIV-1 replication
in human MDMs, and increased co-receptor expression on
dendritic cells certainly suggest that further investigation of
the effects of MA on HIV pathogenesis in the periphery are
warranted. The next several years should yield some
interesting results regarding MAHIV interactions.
Acknowledgments The preparation of this review was supported by
funding from National Institute on Drug Abuse (DA025528 and
DA025011).
Conflicts of interest The authors declare no conflicts of interest.
References
Acikgoz O, Gonenc S, Kayatekin BM, Uysal N, Pekcetin C, Semin I,
Gure A (1998) Methamphetamine causes lipid peroxidation and
an increase in superoxide dismutase activity in the rat striatum.
Brain Res 813:200202
Amenta F, Bronzetti E, Felici L, Ricci A, Tayebati SK (1999)
Dopamine D2-like receptors on human peripheral blood lym-
phocytes: a radioligand binding assay and immunocytochemical
study. J Auton Pharmacol 19:151159
Angulo JA, Angulo N, Yu J (2004) Antagonists of the neurokinin-1 or
dopamine D1 receptors confer protection from methamphetamine
on dopamine terminals of the mouse striatum. Ann N Y Acad Sci
1025:171180
Anlauf M, Schafer MK, Schwark T, von Wurmb-Schwark N, Brand V,
Sipos B, Horny HP, Parwaresch R, Hartschuh W, Eiden LE,
Kloppel G, Weihe E (2006) Vesicular monoamine transporter 2
(VMAT2) expression in hematopoietic cells and in patients with
systemic mastocytosis. J Histochem Cytochem 54:201213
Antinori A, Arendt G, Becker JT, Brew BJ, Byrd DA, Cherner M,
Clifford DB, Cinque P, Epstein LG, Goodkin K, Gisslen M, Grant
I, Heaton RK, Joseph J, Marder K, Marra CM, McArthur JC,
Nunn M, Price RW, Pulliam L, Robertson KR, Sacktor N, Valcour
J. Neurovirol. (2011) 17:401415 411
V, Wojna VE (2007) Updated research nosology for HIV-
associated neurocognitive disorders. Neurology 69:17891799
Banerjee A, Zhang X, Manda KR, Banks WA, Ercal N (2010) HIV
proteins (gp120 and Tat) and methamphetamine in oxidative
stress-induced damage in the brain: potential role of the thiol
antioxidant N-acetylcysteine amide. Free Radic Biol Med
48:13881398
Berger JR, Kumar M, Kumar A, Fernandez JB, Levin B (1994)
Cerebrospinal fluid dopamine in HIV-1 infection. AIDS 8:6771
Brown PL, Kiyatkin EA (2005) Fatal intra-brain heat accumulation
induced by meth-amphetamine at normothermic conditions in
rats. Int J Neuroprotect Neuroregeneration 1:8690
Brown JM, Riddle EL, Sandoval V, Weston RK, Hanson JE, Crosby
MJ, Ugarte YV, Gibb JW, Hanson GR, Fleckenstein AE (2002) A
single methamphetamine administration rapidly decreases vesicular
dopamine uptake. J Pharmacol Exp Ther 302:497501
Brown PL, Wise RA, Kiyatkin EA (2003) Brain hyperthermia is
induced by methamphetamine and exacerbated by social interac-
tion. J Neurosci 23:39243929
Buchacz K, McFarland W, Kellogg TA, Loeb L, Holmberg SD, Dilley
J, Klausner JD (2005) Amphetamine use is associated with
increased HIV incidence among men who have sex with men in
San Francisco. AIDS 19:14231424
Cadet JL, Krasnova IN (2007) Interactions of HIV and methamphet-
amine: cellular and molecular mechanisms of toxicity potentia-
tion. Neurotox Res 12:181204
Cadet JL, Sheng P, Ali S, Rothman R, Carlson E, Epstein C (1994)
Attenuation of methamphetamine-induced neurotoxicity in copper/
zinc superoxide dismutase transgenic mice. J Neurochem 62:380
383
Cass WA, Harned ME, Peters LE, Nath A, Maragos WF (2003) HIV-1
protein Tat potentiation of methamphetamine-induced decreases
in evoked overflow of dopamine in the striatum of the rat. Brain
Res 984:133142
Chang L, Ernst T, Speck O, Grob CS (2005) Additive effects of HIV
and chronic methamphetamine use on brain metabolite abnor-
malities. Am J Psychiatry 162:361369
Chen HM, Lee YC, Huang CL, Liu HK, Liao WC, Lai WL, Lin YR,
Huang NK (2007) Methamphetamine downregulates peroxire-
doxins in rat pheochromocytoma cells. Biochem Biophys Res
Commun 354:96101
Chung YA, Peterson BS, Yoon SJ, Cho SN, Chai S, Jeong J, Kim DJ
(2010) In vivo evidence for long-term CNS toxicity, associated
with chronic binge use of methamphetamine. Drug Alcohol
Depend 111:155160
Clifford DB (2008) HIV-associated neurocognitive disease continues
in the antiretroviral era. Top HIV Med 16:9498
Conant K, St Hillaire C, Anderson C, Galey D, Wang J, Nath A (2004)
Human immunodeficiency virus type 1 Tat and methamphet-
amine affect the release and activation of matrix-degrading
proteinases. J Neurovirol 10:2128
Cubells JF, Rayport S, Rajendran G, Sulzer D (1994) Methamphet-
amine neurotoxicity involves vacuolation of endocytic organelles
and dopamine-dependent intracellular oxidative stress. J Neurosci
14:22602271
Davidson C, Gow AJ, Lee TH, Ellinwood EH (2001) Methamphet-
amine neurotoxicity: necrotic and apoptotic mechanisms and
relevance to human abuse and treatment. Brain Res Brain Res
Rev 36:122
Degenhardt L, Mathers B, Guarinieri M, Panda S, Phillips B,
Strathdee SA, Tyndall M, Wiessing L, Wodak A, Howard J
(2010) Meth/amphetamine use and associated HIV: Implications
for global policy and public health. Int J Drug Policy 21:347358
Dluzen DE, McDermott JL, Darvesh AS (2010) Relationships among
gender, age, time, and temperature in methamphetamine-induced
striatal dopaminergic neurotoxicity. Neuroscience 167:985993
Eyerman DJ, Yamamoto BK (2005) Lobeline attenuates
methamphetamine-induced changes in vesicular monoamine
transporter 2 immunoreactivity and monoamine depletions in
the striatum. J Pharmacol Exp Ther 312:160169
Fleckenstein AE, Volz TJ, Hanson GR (2009) Psychostimulant-
induced alterations in vesicular monoamine transporter-2 function:
neurotoxic and therapeutic implications. Neuropharmacology 56
(Suppl 1):133138
Flora G, Lee YW, Nath A, Maragos W, Hennig B, Toborek M (2002)
Methamphetamine-induced TNF-alpha gene expression and
activation of AP-1 in discrete regions of mouse brain: potential
role of reactive oxygen intermediates and lipid peroxidation.
Neuromolecular Med 2:7185
Flora G, Lee YW, Nath A, Hennig B, Maragos W, Toborek M (2003)
Methamphetamine potentiates HIV-1 Tat protein-mediated activa-
tion of redox-sensitive pathways in discrete regions of the brain. Exp
Neurol 179:6070
Frey K, Kilbourn M, Robinson T (1997) Reduced striatal vesicular
monoamine transporters after neurotoxic but not after
behaviorally-sensitizing doses of methamphetamine. Eur J
Pharmacol 334:273279
Fumagalli F, Gainetdinov RR, Valenzano KJ, Caron MG (1998) Role
of dopamine transporter in methamphetamine-induced neurotoxicity:
evidence from mice lacking the transporter. J Neurosci 18:4861
4869
Fumagalli F, Gainetdinov RR, Wang YM, Valenzano KJ, Miller GW,
Caron MG (1999) Increased methamphetamine neurotoxicity in
heterozygous vesicular monoamine transporter 2 knock-out mice.
J Neurosci 19:24242431
Garcia de Yebenes J, Yebenes J, Mena MA (2000) Neurotrophic
factors in neurodegenerative disorders: model of Parkinsons
disease. Neurotox Res 2:115137
Genc K, Genc S, Kizildag S, Sonmez U, Yilmaz O, Tugyan K, Ergur
B, Sonmez A, Buldan Z (2003) Methamphetamine induces
oligodendroglial cell death in vitro. Brain Res 982:125130
Gendelman HE, Meltzer MS (1989) Mononuclear phagocytes and the
human immunodeficiency virus. Curr Opin Immunol 2:414419
Gibb JW, Kogan FJ (1979) Influence of dopamine synthesis on
methamphetamine-induced changes in striatal and adrenal tyrosine
hydroxylase activity. Naunyn Schmiedebergs Arch Pharmacol
310:185187
Glasner-Edwards S, Mooney LJ, Marinelli-Casey P, Hillhouse M, Ang
A, Rawson RA (2010) Psychopathology in methamphetamine-
dependent adults 3 years after treatment. Drug Alcohol Rev
29:1220
Glass JD, Wesselingh SL, Selnes OA, McArthur JC (1993) Clinical-
neuropathologic correlation in HIV-associated dementia. Neurology
43:22302237
Gluck MR, Moy LY, Jayatilleke E, Hogan KA, Manzino L, Sonsalla
PK (2001) Parallel increases in lipid and protein oxidative
markers in several mouse brain regions after methamphetamine
treatment. J Neurochem 79:152160
Goncalves J, Martins T, Ferreira R, Milhazes N, Borges F, Ribeiro CF,
Malva JO, Macedo TR, Silva AP (2008) Methamphetamine-
induced early increase of IL-6 and TNF-alpha mRNA expression
in the mouse brain. Ann N Y Acad Sci 1139:103111
Gonzales R, Ang A, Marinelli-Casey P, Glik DC, Iguchi MY, Rawson
RA (2009) Health-related quality of life trajectories of
methamphetamine-dependent individuals as a function of treat-
ment completion and continued care over a 1-year period. J Subst
Abuse Treat 37:353361
Gonzales R, Mooney L, Rawson RA (2010) The methamphetamine
problem in the United States. Annu Rev Public Health 31:385
398
Gorbach PM, Drumright LN, Javanbakht M, Pond SL, Woelk CH,
Daar ES, Little SJ (2008) Antiretroviral drug resistance and risk
412 J. Neurovirol. (2011) 17:401415
behavior among recently HIV-infected men who have sex with
men. J Acquir Immune Defic Syndr 47:639643
Hogan KA, Staal RG, Sonsalla PK (2000) Analysis of VMAT2
binding after methamphetamine or MPTP treatment: disparity
between homogenates and vesicle preparations. J Neurochem
74:22172220
Hotchkiss AJ, Gibb JW (1980) Long-term effects of multiple doses of
methamphetamine on tryptophan hydroxylase and tyrosine
hydroxylase activity in rat brain. J Pharmacol Exp Ther
214:257262
Itoh K, Mehraein P, Weis S (2000) Neuronal damage of the substantia
nigra in HIV-1 infected brains. Acta Neuropathol 99:376384
Iwazaki T, McGregor IS, Matsumoto I (2006) Protein expression
profile in the striatum of acute methamphetamine-treated rats.
Brain Res 1097:1925
Iyo M, Sekine Y, Mori N (2004) Neuromechanism of developing
methamphetamine psychosis: a neuroimaging study. Ann N Y
Acad Sci 1025:288295
Jayanthi S, Ladenheim B, Cadet JL (1998) Methamphetamine-induced
changes in antioxidant enzymes and lipid peroxidation in copper/
zinc-superoxide dismutase transgenic mice. Ann N Y Acad Sci
844:92102
Jayanthi S, Deng X, Ladenheim B, McCoy MT, Cluster A, Cai NS,
Cadet JL (2005) Calcineurin/NFAT-induced up-regulation of the
Fas ligand/Fas death pathway is involved in methamphetamine-
induced neuronal apoptosis. Proc Natl Acad Sci U S A 102:868
873
Kanthasamy A, Anantharam V, Ali SF, Kanthasamy AG (2006)
Methamphetamine induces autophagy and apoptosis in a mesence-
phalic dopaminergic neuronal culture model: role of cathepsin-D in
methamphetamine-induced apoptotic cell death. Ann N Y Acad Sci
1074:234244
Kiyatkin EA (2005) Brain hyperthermia as physiological and
pathological phenomena. Brain Res Brain Res Rev 50:2756
Kiyatkin EA (2010) Brain temperature homeostasis: physiological
fluctuations and pathological shifts. Front Biosci 15:7392
Kiyatkin EA, Sharma HS (2011) Expression of heat shock protein
(HSP 72 kDa) during acute methamphetamine intoxication depends
on brain hyperthermia: neurotoxicity or neuroprotection? J Neural
Transm 118:4760
Kiyatkin EA, Brown PL, Sharma HS (2007) Brain edema and
breakdown of the blood-brain barrier during methamphetamine
intoxication: critical role of brain hyperthermia. Eur J Neurosci
26:12421253
Krasnova IN, Cadet JL (2009) Methamphetamine toxicity and
messengers of death. Brain Res Rev 60:379407
Kuhn DM, Francescutti-Verbeem DM, Thomas DM (2008) Dopamine
disposition in the presynaptic process regulates the severity of
methamphetamine-induced neurotoxicity. Ann N Y Acad Sci
1139:118126
Kuperman DI, Freyaldenhoven TE, Schmued LC, Ali SF (1997)
Methamphetamine-induced hyperthermia in mice: examination of
dopamine depletion and heat-shock protein induction. Brain Res
771:221227
LaGasse LL, Wouldes T, Newman E, Smith LM, Shah RZ, Derauf C,
Huestis MA, Arria AM, Della Grotta S, Wilcox T, Lester BM
(2011) Prenatal methamphetamine exposure and neonatal neuro-
behavioral outcome in the USA and New Zealand. Neurotoxicol
Teratol 33:166175
Langford D, Adame A, Grigorian A, Grant I, McCutchan JA, Ellis RJ,
Marcotte TD, Masliah E (2003) Patterns of selective neuronal
damage in methamphetamine-user AIDS patients. J Acquir
Immune Defic Syndr 34:467474
Langford D, Grigorian A, Hurford R, Adame A, Crews L, Masliah E
(2004) The role of mitochondrial alterations in the combined
toxic effects of human immunodeficiency virus Tat protein and
methamphetamine on calbindin positive-neurons. J Neurovirol
10:327337
Larsen KE, Fon EA, Hastings TG, Edwards RH, Sulzer D (2002)
Methamphetamine-induced degeneration of dopaminergic neurons
involves autophagy and upregulation of dopamine synthesis. J
Neurosci 22:89518960
LaVoie MJ, Hastings TG (1999) Dopamine quinone formation and
protein modification associated with the striatal neurotoxicity of
methamphetamine: evidence against a role for extracellular
dopamine. J Neurosci 19:14841491
Liang H, Wang X, Chen H, Song L, Ye L, Wang SH, Wang YJ, Zhou
L, Ho WZ (2008) Methamphetamine enhances HIV infection of
macrophages. Am J Pathol 172:16171624
Liu X, Chang L, Vigorito M, Kass M, Li H, Chang SL (2009)
Methamphetamine-induced behavioral sensitization is enhanced
in the HIV-1 transgenic rat. J Neuroimmune Pharmacol 4:309
316
Madden LJ, Flynn CT, Zandonatti MA, May M, Parsons LH, Katner
SN, Henriksen SJ, Fox HS (2005) Modeling human metham-
phetamine exposure in nonhuman primates: chronic dosing in the
rhesus macaque leads to behavioral and physiological abnormalities.
Neuropsychopharmacology 30:350359
Magnuson DS, Knudsen BE, Geiger JD, Brownstone RM, Nath A
(1995) Human immunodeficiency virus type 1 tat activates non-
N-methyl-D-aspartate excitatory amino acid receptors and causes
neurotoxicity. Ann Neurol 37:373380
Mahajan SD, Hu Z, Reynolds JL, Aalinkeel R, Schwartz SA, Nair MP
(2006) Methamphetamine modulates gene expression patterns in
monocyte derived mature dendritic cells: implications for HIV-1
pathogenesis. Mol Diagn Ther 10:257269
Mahajan SD, Aalinkeel R, Sykes DE, Reynolds JL, Bindukumar B,
Adal A, Qi M, Toh J, Xu G, Prasad PN, Schwartz SA (2008)
Methamphetamine alters blood brain barrier permeability via the
modulation of tight junction expression: Implication for HIV-1
neuropathogenesis in the context of drug abuse. Brain Res
1203:133148
Mao CV, Hori E, Maior RS, Ono T, Nishijo H (2008) A primate model
of schizophrenia using chronic PCP treatment. Rev Neurosci
19:8389
Maragos WF, Young KL, Turchan JT, Guseva M, Pauly JR, Nath A,
Cass WA (2002) Human immunodeficiency virus-1 Tat protein
and methamphetamine interact synergistically to impair striatal
dopaminergic function. J Neurochem 83:955963
Marcondes MC, Flynn C, Watry DD, Zandonatti M, Fox HS (2010)
Methamphetamine increases brain viral load and activates natural
killer cells in simian immunodeficiency virus-infected monkeys.
Am J Pathol 177:355361
Marek GJ, Vosmer G, Seiden LS (1990) Dopamine uptake inhibitors
block long-term neurotoxic effects of methamphetamine upon
dopaminergic neurons. Brain Res 513:274279
McArthur JC, Hoover DR, Bacellar H, Miller EN, Cohen BA, Becker
JT, Graham NM, McArthur JH, Selnes OA, Jacobson LP et al
(1993) Dementia in AIDS patients: incidence and risk factors.
Multicenter AIDS Cohort Study. Neurology 43:22452252
Melega WP, Lacan G, Harvey DC, Way BM (2007) Methamphetamine
increases basal ganglia iron to levels observed in aging. Neuro-
report 18:17411745
Melega WP, Jorgensen MJ, Lacan G, Way BM, Pham J, Morton G,
Cho AK, Fairbanks LA (2008) Long-term methamphetamine
administration in the vervet monkey models aspects of a human
exposure: brain neurotoxicity and behavioral profiles. Neuro-
psychopharmacology 33:14411452
Meltzer MS, Gendelman HE (1992) Mononuclear phagocytes as
targets, tissue reservoirs, and immunoregulatory cells in human
immunodeficiency virus disease. Curr Top Microbiol Immunol
181:239263
J. Neurovirol. (2011) 17:401415 413
Meltzer MS, Skillman DR, Gomatos PJ, Kalter DC, Gendelman HE
(1990) Role of mononuclear phagocytes in the pathogenesis of
human immunodeficiency virus infection. Annu Rev Immunol
8:169194
Miyazaki I, Asanuma M, Diaz-Corrales FJ, Fukuda M, Kitaichi K,
Miyoshi K, Ogawa N (2006) Methamphetamine-induced dopa-
minergic neurotoxicity is regulated by quinone-formation-related
molecules. FASEB J 20:571573
Munch G, Raether A, Schoffel E, Illes P (1991) Postsynaptic
dopamine DA1- and DA2-receptors in jejunal arteries of rabbits.
J Cardiovasc Pharmacol 18:468471
Nair MP, Saiyed ZM (2010) Effect of methamphetamine on expression
of HIV coreceptors and CC-chemokines by dendritic cells. Life
Sci 88:987994
Nair MP, Mahajan S, Sykes D, Bapardekar MV, Reynolds JL (2006)
Methamphetamine modulates DC-SIGN expression by mature
dendritic cells. J Neuroimmune Pharmacol 1:296304
Narita M, Miyatake M, Shibasaki M, Tsuda M, Koizumi S, Yajima Y,
Inoue K, Suzuki T (2005) Long-lasting change in brain dynamics
induced by methamphetamine: enhancement of protein kinase C-
dependent astrocytic response and behavioral sensitization. J
Neurochem 93:13831392
Nath A, Anderson C, Jones M, Maragos W, Booze R, Mactutus C,
Bell J, Hauser KF, Mattson M (2000) Neurotoxicity and
dysfunction of dopaminergic systems associated with AIDS
dementia. J Psychopharmacol 14:222227
Navia BA, Cho ES, Petito CK, Price RW (1986a) The AIDS dementia
complex: II. Neuropathology. Ann Neurol 19:525535
Navia BA, Jordan BD, Price RW (1986b) The AIDS dementia
complex: I. Clinical features. Ann Neurol 19:517524
Newton TF, Kalechstein AD, Hardy DJ, Cook IA, Nestor L, Ling W,
Leuchter AF (2004) Association between quantitative EEG and
neurocognition in methamphetamine-dependent volunteers. Clin
Neurophysiol 115:194198
Office of Applied Studies S (2007) Methamphetamine Abuse. The
NSDUH Report
Olsen NV (1998) Effects of dopamine on renal haemodynamics
tubular function and sodium excretion in normal humans. Dan
Med Bull 45:282297
Ozono R, OConnell DP, Wang ZQ, Moore AF, Sanada H, Felder
RA, Carey RM (1997) Localization of the dopamine D1
receptor protein in the human heart and kidney. Hypertension
30:725729
Paulus MP, Hozack NE, Zauscher BE, Frank L, Brown GG, Braff DL,
Schuckit MA (2002) Behavioral and functional neuroimaging
evidence for prefrontal dysfunction in methamphetamine-
dependent subjects. Neuropsychopharmacology 26:5363
Pendyala G, Trauger SA, Siuzdak G, Fox HS (2011) Short
communication: quantitative proteomic plasma profiling reveals
activation of host defense to oxidative stress in chronic SIV and
methamphetamine comorbidity. AIDS Res Hum Retroviruses
27:179182
Pivonello R, Ferone D, de Herder WW, de Krijger RR, Waaijers M,
Mooij DM, van Koetsveld PM, Barreca A, De Caro ML,
Lombardi G, Colao A, Lamberts SW, Hofland LJ (2004)
Dopamine receptor expression and function in human normal
adrenal gland and adrenal tumors. J Clin Endocrinol Metab
89:44934502
Ramirez SH, Potula R, Fan S, Eidem T, Papugani A, Reichenbach N,
Dykstra H, Weksler BB, Romero IA, Couraud PO, Persidsky Y
(2009) Methamphetamine disrupts blood-brain barrier function
by induction of oxidative stress in brain endothelial cells. J Cereb
Blood Flow Metab 29:19331945
Rawson RA, Gonzales R, Brethen P (2002) Treatment of metham-
phetamine use disorders: an update. J Subst Abuse Treat 23:145
150
Reiner BC, Keblesh JP, Xiong H (2009) Methamphetamine abuse,
HIV infection, and neurotoxicity. Int J Physiol Pathophysiol
Pharmacol 1:162179
Reyes MG, Faraldi F, Senseng CS, Flowers C, Fariello R (1991)
Nigral degeneration in acquired immune deficiency syndrome
(AIDS). Acta Neuropathol 82:3944
Reynolds JL, Mahajan SD, Sykes DE, Schwartz SA, Nair MP (2007)
Proteomic analyses of methamphetamine (METH)-induced dif-
ferential protein expression by immature dendritic cells (IDC).
Biochim Biophys Acta 1774:433442
Reynolds JL, Mahajan SD, Aalinkeel R, Nair B, Sykes DE, Agosto-
Mujica A, Hsiao CB, Schwartz SA (2009) Modulation of the
proteome of peripheral blood mononuclear cells from HIV-1-
infected patients by drugs of abuse. J Clin Immunol 29:646656
Ricci A, Amenta F (1994) Dopamine D5 receptors in human
peripheral blood lymphocytes: a radioligand binding study. J
Neuroimmunol 53:17
Ricci A, Bronzetti E, Felici L, Tayebati SK, Amenta F (1997)
Dopamine D4 receptor in human peripheral blood lymphocytes:
a radioligand binding assay study. Neurosci Lett 229:130134
Ricci A, Bronzetti E, Felici L, Greco S, Amenta F (1998) Labeling of
dopamine D3 and D4 receptor subtypes in human peripheral blood
lymphocytes with [3H]7-OH-DPAT:a combined radioligand binding
assay and immunochemical study. J Neuroimmunol 92:191195
Ricci A, Bronzetti E, Mignini F, Tayebati SK, Zaccheo D, Amenta F
(1999) Dopamine D1-like receptor subtypes in human peripheral
blood lymphocytes. J Neuroimmunol 96:234240
Riddle EL, Fleckenstein AE, Hanson GR (2006) Mechanisms of
methamphetamine-induced dopaminergic neurotoxicity. AAPS J
8:E413E418
Roberts AJ, Maung R, Sejbuk NE, Ake C, Kaul M (2010) Alteration
of methamphetamine-induced stereotypic behaviour in transgenic
mice expressing HIV-1 envelope protein gp120. J Neurosci
Methods 186:222225
Rohr O, Sawaya BE, Lecestre D, Aunis D, Schaeffer E (1999)
Dopamine stimulates expression of the human immunodeficiency
virus type 1 via NF-kappaB in cells of the immune system.
Nucleic Acids Res 27:32913299
Romero CA, Bustamante DA, Zapata-Torres G, Goiny M, Cassels B,
Herrera-Marschitz M (2006) Neurochemical and behavioural
characterisation of alkoxyamphetamine derivatives in rats. Neuro-
tox Res 10:1122
Roussotte F, Soderberg L, Sowell E (2010) Structural, metabolic, and
functional brain abnormalities as a result of prenatal exposure to
drugs of abuse: evidence from neuroimaging. Neuropsychol Rev
20:376397
Ruffolo RR Jr, Messick K (1985) Effects of dopamine, (+/)-dobutamine
and the (+)- and ()-enantiomers of dobutamine on cardiac function
in pithed rats. J Pharmacol Exp Ther 235:558565
Saisho Y, Harris PE, Butler AE, Galasso R, Gurlo T, Rizza RA, Butler
PC (2008) Relationship between pancreatic vesicular monoamine
transporter 2 (VMAT2) and insulin expression in human
pancreas. J Mol Histol 39:543551
Sandoval V, Riddle EL, Hanson GR, Fleckenstein AE (2003)
Methylphenidate alters vesicular monoamine transport and
prevents methamphetamine-induced dopaminergic deficits. J
Pharmacol Exp Ther 304:11811187
Schmidt CJ, Gibb JW (1985) Role of the dopamine uptake carrier in
the neurochemical response to methamphetamine: effects of
amfonelic acid. Eur J Pharmacol 109:7380
Schmidt CJ, Ritter JK, Sonsalla PK, Hanson GR, Gibb JW (1985)
Role of dopamine in the neurotoxic effects of methamphetamine.
J Pharmacol Exp Ther 233:539544
Schweinsburg BC, Taylor MJ, Alhassoon OM, Gonzalez R, Brown GG,
Ellis RJ, Letendre S, Videen JS, McCutchan JA, Patterson TL, Grant
I (2005) Brain mitochondrial injury in human immunodeficiency
414 J. Neurovirol. (2011) 17:401415
virus-seropositive (HIV+) individuals taking nucleoside reverse
transcriptase inhibitors. J Neurovirol 11:356364
Sharma HS, Kiyatkin EA (2009) Rapid morphological brain abnor-
malities during acute methamphetamine intoxication in the rat: an
experimental study using light and electron microscopy. J Chem
Neuroanat 37:1832
Shoptaw S, Reback CJ (2007) Methamphetamine use and infectious
disease-related behaviors in men who have sex with men:
implications for interventions. Addiction 102(Suppl 1):130135
Siegel JA, Craytor MJ, Raber J (2010) Long-term effects of
methamphetamine exposure on cognitive function and muscarinic
acetylcholine receptor levels in mice. Behav Pharmacol 21:602614
Siegel JA, Park BS, Raber J (2011) Long-term effects of neonatal
methamphetamine exposure on cognitive function in adolescent
mice. Behav Brain Res 219:159164
Smith LM, Lagasse LL, Derauf C, Grant P, Shah R, Arria A, Huestis
M, Haning W, Strauss A, Della Grotta S, Fallone M, Liu J, Lester
BM (2008) Prenatal methamphetamine use and neonatal neuro-
behavioral outcome. Neurotoxicol Teratol 30:2028
Sonsalla PK, Gibb JW, Hanson GR (1986) Roles of D1 and D2 dopamine
receptor subtypes in mediating the methamphetamine-induced
changes in monoamine systems. J Pharmacol Exp Ther 238:932937
Sriram K, Miller DB, OCallaghan JP (2006) Minocycline attenuates
microglial activation but fails to mitigate striatal dopaminergic
neurotoxicity: role of tumor necrosis factor-alpha. J Neurochem
96:706718
Stephans SE, Yamamoto BK (1994) Methamphetamine-induced
neurotoxicity: roles for glutamate and dopamine efflux. Synapse
17:203209
Sulzer D, Sonders MS, Poulsen NW, Galli A (2005) Mechanisms of
neurotransmitter release by amphetamines: a review. Prog Neuro-
biol 75:406433
Talloczy Z, Martinez J, Joset D, Ray Y, Gacser A, Toussi S,
Mizushima N, Nosanchuk JD, Goldstein H, Loike J, Sulzer D,
Santambrogio L (2008) Methamphetamine inhibits antigen
processing, presentation, and phagocytosis. PLoS Pathog 4:e28
Tang Y, Nee AC, Lu A, Ran R, Sharp FR (2003) Blood genomic
expression profile for neuronal injury. J Cereb Blood Flow Metab
23:310319
Theodore S, Cass WA, Maragos WF (2006a) Involvement of cytokines
in human immunodeficiency virus-1 protein Tat and metham-
phetamine interactions in the striatum. Exp Neurol 199:490498
Theodore S, Cass WA, Nath A, Steiner J, Young K, Maragos WF
(2006b) Inhibition of tumor necrosis factor-alpha signaling
prevents human immunodeficiency virus-1 protein Tat and
methamphetamine interaction. Neurobiol Dis 23:663668
Thomas SA (2004) Anti-HIV drug distribution to the central nervous
system. Curr Pharm Des 10:13131324
Thomas DM, Dowgiert J, Geddes TJ, Francescutti-Verbeem D, Liu X,
Kuhn DM (2004a) Microglial activation is a pharmacologically
specific marker for the neurotoxic amphetamines. Neurosci Lett
367:349354
Thomas DM, Walker PD, Benjamins JA, Geddes TJ, Kuhn DM
(2004b) Methamphetamine neurotoxicity in dopamine nerve
endings of the striatum is associated with microglial activation.
J Pharmacol Exp Ther 311:17
Thompson PM, Hayashi KM, Simon SL, Geaga JA, Hong MS, Sui Y,
Lee JY, Toga AW, Ling W, London ED (2004) Structural
abnormalities in the brains of human subjects who use metham-
phetamine. J Neurosci 24:60286036
Toussi SS, Joseph A, Zheng JH, Dutta M, Santambrogio L, Goldstein
H (2009) Short communication: methamphetamine treatment
increases in vitro and in vivo HIV replication. AIDS Res Hum
Retroviruses 25:11171121
Turchan J, Anderson C, Hauser KF, Sun Q, Zhang J, Liu Y, Wise PM,
Kruman I, Maragos W, Mattson MP, Booze R, Nath A (2001)
Estrogen protects against the synergistic toxicity by HIV
proteins, methamphetamine and cocaine. BMC Neurosci 2:3
Wagner GC, Ricaurte GA, Seiden LS, Schuster CR, Miller RJ, Westley
J (1980) Long-lasting depletions of striatal dopamine and loss of
dopamine uptake sites following repeated administration of
methamphetamine. Brain Res 181:151160
Wilson JM, Kalasinsky KS, Levey AI, Bergeron C, Reiber G,
Anthony RM, Schmunk GA, Shannak K, Haycock JW, Kish SJ
(1996) Striatal dopamine nerve terminal markers in human,
chronic methamphetamine users. Nat Med 2:699703
Witkovsky P (2004) Dopamine and retinal function. Doc Ophthalmol
108:1740
Xu W, Zhu JP, Angulo JA (2005) Induction of striatal pre- and
postsynaptic damage by methamphetamine requires the dopamine
receptors. Synapse 58:110121
Zweben JE, Cohen JB, Christian D, Galloway GP, Salinardi M, Parent
D, Iguchi M (2004) Psychiatric symptoms in methamphetamine
users. Am J Addict 13:181190
J. Neurovirol. (2011) 17:401415 415
... Ferroptosis, characterized by iron accumulation and lipid peroxidation, is a form of cell death related to lipid ROS that could be inhibited by the lipid peroxidation inhibitor ferrostatin-1 (Fer-1, a synthetic antioxidant) (Hayes and Dinkova-Kostova 2014). In general, ferroptosis is manifested by decreased levels of glutathione (GSH) peroxidase 4 (GPX4) and an increased level of p53 (Silverstein et al. 2011). GPX4 is an important enzyme that exerts antioxidant functions in concert with GSH, and the inactivation of GPX4 leads to lipid peroxidation and ferroptosis (Liu et al. 2013). ...
... Activation of p53 declines cystine uptake, which in turn limits the production of glutathione (GSH) and further increases the susceptibility of cells to ferroptosis (Thangaraj et al. 2021). Ferroptosis is reported to account for brain injury, neuroinflammation, and neurodegenerative diseases (Silverstein et al. 2011;Ingold et al. 2018;Song and Long 2020). Studies have verified that ferroptosis is a significant contributor to drug-induced toxic nerve injury, including METH, morphine, and virus-induced brain damage (Friedmann Angeli et al. 2014;Kang et al. 2019). ...
Article
Full-text available
Methamphetamine (METH) and HIV-1 lead to oxidative stress and their combined effect increases the risk of HIV-associated neurocognitive disorder (HAND), which may be related to the synergistic ferroptotic impairment in microglia. Ferroptosis is a redox imbalance cell damage associated with iron overload that is linked to the pathogenic processes of METH and HIV-1. NRF2 is an antioxidant transcription factor that plays a protective role in METH and HIV-1-induced neurotoxicity, but its mechanism has not been fully elucidated. To explore the role of ferroptosis in METH abuse and HIV-1 infection and the potential role of NRF2 in this process, we conducted METH and HIV-1 Tat exposure models using the BV2 microglia cells. We found that METH and HIV-1 Tat reduced the expression of ferroptotic protein GPX4 and the cell viability and enhanced the expression of P53 and the level of ferrous iron, while the above indices were significantly improved with pretreatment of ferrostatin-1. In addition, NRF2 knockdown accelerated METH and HIV-1 Tat-induced BV2 cell ferroptosis accompanied by decreased expression of SLC7A11. On the contrary, NRF2 stimulation significantly increased the expression of SLC7A11 and attenuated ferroptosis in cells. In summary, our study indicates that METH and HIV-1 Tat synergistically cause BV2 cell ferroptosis, while NRF2 antagonizes BV2 cell ferroptotic damage induced by METH and HIV-1 Tat through regulation of SLC7A11. Overall, this study provides potential therapeutic strategies for the treatment of neurotoxicity caused by METH and HIV-1 Tat, providing a theoretical basis and new targets for the treatment of HIV-infected drug abusers.
... Cocaine and METH initiate the transcriptional initiation of HIV genes that are crucial for the replication of viral proteins [69]. Cocaine also promotes the infiltration of HIV-1 viral proteins into brain tissue, while METH, in conjunction with the HIV viral protein gp120, reduces the expression of claudin-3, -5, ZO-1, and occludin [29,70]. ...
Article
Full-text available
‘Drug abuse’ has been recognized as one of the most pressing epidemics in contemporary society. Traditional research has primarily focused on understanding how drugs induce neurotoxicity or degeneration within the central nervous system (CNS) and influence systems related to reward, motivation, and cravings. However, recent investigations have increasingly shifted their attention toward the detrimental consequences of drug abuse on the blood–brain barrier (BBB). The BBB is a structural component situated in brain vessels, responsible for separating brain tissue from external substances to maintain brain homeostasis. The BBB’s function is governed by cellular interactions involving various elements of the ‘neurovascular unit (NVU),’ such as neurons, endothelial cells, astrocytes, pericytes, and microglia. Disruption of the NVU is closely linked to serious neurodegeneration. This review provides a comprehensive overview of the harmful effects of psychostimulant drugs on the BBB, highlighting the mechanisms through which drugs can damage the NVU. Additionally, the review proposes novel therapeutic targets aimed at protecting the BBB. By understanding the intricate relationships between drug abuse, BBB integrity, and NVU function, researchers and clinicians may uncover new strategies to mitigate the damaging impact of drug abuse on brain health.
... Additionally, reports have shown that increased CRP and SAA in HIV-1 infection contribute to immune suppression and progression [62][63][64]. Furthermore, it has also been reported that drugs of abuse (methamphetamine, cocaine, alcohol) and HIV-1 infection have an increased risk of developing chronic neuroinflammation, enhanced immune dysfunction (predisposition to HIV-1 infection), and cognitive impairment [65][66][67][68][69][70]. This has been demonstrated in studies on various CNS cells (neurons, glial cells, and pericytes) exposed to HIV-1/HIV-1 proteins and drugs of abuse, which led to increased expression of proinflammatory cytokines (such as TNFα, IL1β, and IL6) in both the intracellular and extracellular compartments [24,25,30,38,65,[71][72][73]. ...
Article
Full-text available
Chronic low-grade inflammation remains an essential feature of HIV-1 infection under combined antiretroviral therapy (cART) and contributes to the accelerated cognitive defects and aging in HIV-1 infected populations, indicating cART limitations in suppressing viremia. Interestingly, ~50% of the HIV-1 infected population on cART that develops cognitive defects is complicated by drug abuse, involving the activation of cells in the central nervous system (CNS) and neurotoxin release, altogether leading to neuroinflammation. Neuroinflammation is the hallmark feature of many neurodegenerative disorders, including HIV-1-associated neurocognitive disorders (HAND). Impaired autophagy has been identified as one of the underlying mechanisms of HAND in treated HIV-1-infected people that also abuse drugs. Several lines of evidence suggest that autophagy regulates CNS cells’ responses and maintains cellular hemostasis. The impairment of autophagy is associated with low-grade chronic inflammation and immune senescence, a known characteristic of pathological aging. Therefore, autophagy impairment due to CNS cells, such as neurons, microglia, astrocytes, and pericytes exposure to HIV-1/HIV-1 proteins, cART, and drug abuse could have combined toxicity, resulting in increased neuroinflammation, which ultimately leads to accelerated aging, referred to as neuroinflammaging. In this review, we focus on the potential role of autophagy in the mechanism of neuroinflammaging in the context of HIV-1 and drug abuse.
... HIV is known to cause DA dysregulation though the release of neurotoxic viral proteins on DA neurons (Bennett et al., 1995;Itoh et al., 2000), and this DA dysregulation is associated with neurocognitive deficits . Thus, METH use can be particularly detrimental to brain and neurocognitive health in HIV+ individuals due to the compounding effects of METH and HIV on the DAergic system and the related mechanisms of oxidative stress, neuroinflammation, and blood brain barrier permeability Silverstein et al., 2011). Consistent with our own findings in METH-individuals without HIV disease , previous studies reported a positive effect of the Met-allele on PFC-dependent neurocognition in HIV+ individuals (Bousman et al., 2010;Saloner et al., 2019;Sundermann et al., 2015); however, this effect was absent in HIV+ men with METH dependence (Bousman et al., 2010). ...
... 13,15 Potentially additive effects of HIV and METH on dopaminergic dysregulation has received considerable attention, yet the combined effects of HIV and METH on norepinephrine (NE) in the CNS are poorly characterized. 16,17 NE is primarily produced in the locus coeruleus, where it is diffusely projected across the brain to perform its neuromodulatory actions. 18 Consequently, the NE system supports a host of cognitive processes, including attention, learning, memory, and decision-making. ...
Article
Full-text available
Objective: HIV disease and methamphetamine (METH) dependence share overlapping mechanisms of neurotoxicity that preferentially compromise monoamine-rich frontostriatal circuitry. However, norepinephrine (NE) function is poorly understood in HIV and METH dependence. We evaluated associations between cerebrospinal fluid (CSF) NE and HIV, METH dependence, and neurocognition. Methods: Participants included 125 adults, stratified by HIV serostatus (HIV+/HIV-) and recent METH dependence (METH+/METH-), who underwent comprehensive neurocognitive testing and lumbar puncture. CSF NE was assayed using high-performance liquid chromatography. Multivariable regression modelled NE as a function of HIV, METH, and their interaction, adjusting for demographic and clinical factors. Pearson's correlations examined relationships between NE and demographically-adjusted neurocognitive domain scores. Results: HIV significantly interacted with METH (p<.001) such that compared to HIV-/METH-, CSF NE was markedly elevated in the single risk-groups (HIV+/METH- : d=0.96; HIV-/METH+: d=0.79) and modestly elevated in the dual-risk group (HIV+/METH+: d=0.48). This interaction remained significant after adjustment for lifetime depression, antidepressant use, and race/ethnicity. In the full sample, higher NE levels significantly correlated with worse global function (r=-0.19), learning (r=-0.23), and delayed recall (r=-0.18). Similar relationships between higher NE and worse neurocognition were detected in the METH- groups (i.e., HIV-/METH- and HIV+/METH-) and in the virally-suppressed persons HIV+ subgroup, but not in the METH+ groups (i.e., HIV-/METH+, HIV+/METH+). Discussion: HIV and METH independently, but not additively, relate to noradrenergic excess in the CNS, and perturbations to noradrenergic function may represent a pathophysiological mechanism of HIV-related neurocognitive dysfunction. Consistent with prior reports that noradrenergic excess compromises hippocampal and prefrontal function, higher NE related to worse neurocognition, even among successfully-treated PWH. Pharmacological and psychosocial interventions that stabilize NE function may improve neurocognition in PWH.
... 15 Use of such stimulants can contribute to unsuppressed VL and coinfections due to disengagement from HIV care, reduced ART adherence, and direct toxic effects on the immune system. 10,[16][17][18] Depression has been identified as the most common mental health issue among PLWH at rates ranging from 30% to 42% and is associated with reduced ART adherence, poorer quality of life, unsuppressed VL, and more rapid disease progression, all of which result in negative clinical outcomes. [19][20][21][22] Adequate management of HIV requires a holistic approach to care that addresses both HIV directly and the comorbid conditions that perpetuate disease progression, 12,23,24 and longitudinal evaluation is necessary to quantify how treatment progression is impacted by multiple comorbidities. ...
Article
Background: In March of 2013, the Los Angeles County (LAC) Division of HIV and STD Programs implemented a clinic-based Medical Care Coordination (MCC) Program to increase viral suppression (VS) (,200 c/mL) among people living with HIV (PLWH) at high risk for poor health outcomes. Objective: This study aimed to estimate trajectories of VS and to assess whether these trajectories differed by stimulant use, housing instability, and depressive symptom severity as reported by PLWH participating in MCC. Methods: Data represent 6408 PLWH in LAC receiving services from the MCC Program and were obtained from LAC HIV surveillance data matched to behavioral assessments obtained across 35 Ryan White Program clinics participating in MCC. Piecewise mixed-effects logistic regression with a random intercept estimated probabilities of VS from 12 months before MCC enrollment through 36 months after enrollment, accounting for time by covariate interactions for 3 comorbid conditions: housing instability, stimulant use, and depressive symptoms. Results: The overall probability of VS increased from 0.35 to 0.77 within the first 6 months in the MCC Program, and this probability was maintained up to 36 months after enrollment. Those who reported housing instability, stimulant use, or multiple comorbid conditions did not achieve the same probability of VS by 36 months as those with none of those comorbidities. Conclusions: Findings suggest that MCC improved the probability of VS for all patient groups regardless of the presence of comorbidities. However, those with comorbid conditions will still require increased support from patient-centered programs to address disparities in VS.
... TAT plays a key role in the dysfunction of the dopamine system associated with HIV disease, and when combined with methamphetamine, has both synergistic or additive effects in dopaminergic function [7,11]. The synergistic effects of TAT and methamphetamine on dopamine transmission suggest that subjects with dual insults may have more severe dependence. ...
Article
Full-text available
Background Methamphetamine abuse and human immunodeficiency virus (HIV) are common comorbidities. HIV-associated proteins, such as the regulatory protein TAT, may contribute to brain reward dysfunction, inducing an altered sensitivity to methamphetamine reward and/or withdrawal in this population. Objective These studies examined the combined effects of TAT protein expression and, chronic and binge methamphetamine regimens on brain reward function. Methods Transgenic mice with inducible brain expression of the TAT protein were exposed to either saline, a chronic, or a binge methamphetamine regimen. TAT expression was induced via doxycycline treatment during the last week of methamphetamine exposure. Brain reward function was assessed daily throughout the regimens, using the intracranial self-stimulation procedure, and after a subsequent acute methamphetamine challenge. Results Both methamphetamine regimens induced withdrawal-related decreases in reward function. TAT expression substantially, but not significantly, increased the withdrawal associated with exposure to the binge regimen compared with the chronic regimen, but did not alter the response to acute methamphetamine challenge. TAT expression also led to persistent changes in adenosine 2B receptor expression in the caudate putamen, regardless of methamphetamine exposure. These results suggest that TAT expression may differentially affect brain reward function, dependent on the pattern of methamphetamine exposure. Conclusion However, the subtle effects observed in these studies highlight that longer-term TAT expression, or its induction at earlier stages of methamphetamine exposure, may be more consequential at inducing behavioral and neurochemical effects.
Article
Military personnel are often exposed to hot environments either for combat operations or peacekeeping missions. Hot environment is a severe stressful situation leading to profound hyperthermia, fatigue and neurological impairments. To avoid stressful environment, some people frequently use methamphetamine (METH) or other psychostimulants to feel comfortable under adverse situations. Our studies show that heat stress alone induces breakdown of the blood-brain barrier (BBB) and edema formation associated with reduced cerebral blood flow (CBF). On the other hand, METH alone induces hyperthermia and neurotoxicity. These effects of METH are exacerbated at high ambient temperatures as seen with greater breakdown of the BBB and brain pathology. Thus, a combination of METH use at hot environment may further enhance the brain damage-associated behavioral dysfunctions. METH is well known to induce severe oxidative stress leading to brain pathology. In this investigation, METH intoxication at hot environment was examined on brain pathology and to explore suitable strategies to induce neuroprotection. Accordingly, TiO2-nanowired delivery of H-290/51 (150 mg/kg, i.p.), a potent chain-breaking antioxidant in combination with mesenchymal stem cells (MSCs), is investigated in attenuating METH-induced brain damage at hot environment in model experiments. Our results show that nanodelivery of H-290/51 with MSCs significantly enhanced CBF and reduced BBB breakdown, edema formation and brain pathology following METH exposure at hot environment. These observations are the first to point out that METH exacerbated brain pathology at hot environment probably due to enhanced oxidative stress, and MSCs attenuate these adverse effects, not reported earlier.
Article
Full-text available
Methamphetamine use has become a rampant public health issue that not only causes devastating consequences to the user but also poses a burden to surrounding communities. A spectrum of ophthalmic sequelae is associated with methamphetamine use and includes episcleritis, scleritis, corneal ulceration, panophthalmitis, endophthalmitis, retinal vasculitis, and retinopathy. In many instances, prompt recognition of the condition and associated infectious process and early initiation of antimicrobial therapy are crucial steps to preventing vision loss. In this review, we summarize the reported ocular complications that may result from methamphetamine use in addition to several postulated mechanisms regarding the ocular toxicity of methamphetamine. The increasing prevalence of methamphetamine use as a public health threat highlights the need for continued investigation of this ophthalmologic issue.
Article
Full-text available
Osteopontin (OPN) also known by its official gene designation secreted phosphoprotein‐1 (SPP1) is a fascinating, multifunctional protein expressed in a number of cell types that functions not only in intercellular communication, but also in the extracellular matrix (ECM). OPN/SPP1 possesses cytokine, chemokine, and signal transduction functions by virtue of modular structural motifs that provide interaction surfaces for integrins and CD44‐variant receptors. In humans, there are three experimentally verified splice variants of OPN/SPP1 and CD44’s ten exons are also alternatively spiced in a cell/tissue‐specific manner, although very little is known about how this is regulated in the central nervous system (CNS). Post‐translational modifications of phosphorylation, glycosylation, and localized cleavage by specific proteases in the cells and tissues where OPN/SPP1 functions, provides additional layers of specificity. However, the former make elucidating the exact molecular mechanisms of OPN/SPP1 function more complex. Flexibility in OPN/SPP1 structure and its engagement with integrins having the ability to transmit signals in inside‐out and outside‐in direction, is likely why OPN/SPP1 can serve as an early detector of inflammation and ongoing tissue damage in response to cancer, stroke, traumatic brain injury, pathogenic infection, and neurodegeneration, processes that impair tissue homeostasis. This review will focus on what is currently known about OPN/SPP1 function in the brain.
Article
Full-text available
Hyperthermia is a symptom of methamphetamine (METH) intoxication and a factor implicated in neurotoxicity during chronic METH use. To characterize the thermic response to METH, it was injected once daily into rats at increasing doses (0, 1, 3, and 9 mg/kg, s.c.) while brain [nucleus accumbens (NAcc), hippocampus] and body (deep temporal muscle) temperatures were continuously monitored. METH produced dose-dependent hyperthermia, with brain structures (especially the NAcc) showing a more rapid and pronounced temperature increase than the muscle. At the highest dose, brain and body temperatures increased 3.5– 4.0°C above basal levels and remained elevated for 3–5 hr. Stressful and other high-activity situations such as interaction with a conspecific female are also known to induce a significant hyperthermic response in the rat. A combination of social interaction and METH administration was tested for additive effects. Male rats were exposed daily to a conspecific female for a total of 120 min, and METH was injected at the same doses 30 min after the initial contact with the female. An initial hyperthermic response ( 1.5°C) to social interaction was followed by a large and prolonged hyperthermic response (3.5–5.0°C, 5–7 hr at 9 mg/kg) to METH, which was again stronger in brain structures (especially in the NAcc) than in the muscle. Although the combined effect of the hyperthermic events was not additive, METH administration during social interaction produced stronger and longer-lasting increases in brain and body temperature than that induced by drug alone, heating the brain in some animals near its biological limit ( 41°C).
Article
Full-text available
The role of the dopamine transporter (DAT) in mediating the neurotoxic effects of methamphetamine (METH) was tested in mice lacking DAT. Dopamine (DA) and serotonin (5-HT) content, glial fibrillary acidic protein (GFAP) expression, and free radical formation were assessed as markers of METH neurotoxicity in the striatum and/or hippocampus of wild-type, heterozygote, and homozygote (DAT -/-) mice. Four injections of METH (15 mg/kg, s.c.), each given 2 hr apart, produced 80 and 30% decreases in striatal DA and 5-HT levels, respectively, in wildtype animals 2 d after administration. In addition, GFAP mRNA and protein expression levels, extracellular DA levels, and free radical formation were increased markedly. Hippocampal 5-HT content was decreased significantly as well (43%). Conversely, no significant changes were observed in total DA content, GFAP expression, extracellular DA levels, or free radical formation in the striatum of DAT -/- mice after METH administration. However, modest decreases were observed in striatal and hippocampal 5-HT levels (10 and 17%, respectively). These observations demonstrate that DAT is required for, and DA is an essential mediator of, METH-induced striatal dopaminergic neurotoxicity, whereas serotonergic deficits are only partially dependent on DAT.
Article
Research into methamphetamine-induced neurotoxicity has experienced a resurgence in recent years. This is due to (1) greater understanding of the mechanisms underlying methamphetamine neurotoxicity, (2) its usefulness as a model for Parkinson's disease and (3) an increased abuse of the substance, especially in the American Mid-West and Japan. It is suggested that the commonly used experimental one-day methamphetamine dosing regimen better models the acute overdose pathologies seen in humans, whereas chronic models are needed to accurately model human long-term abuse. Further, we suggest that these two dosing regimens will result in quite different neurochemical, neuropathological and behavioral outcomes. The relative importance of the dopamine transporter and vesicular monoamine transporter knockout is discussed and insights into oxidative mechanisms are described from observations of nNOS knockout and SOD overexpression. This review not only describes the neuropathologies associated with methamphetamine in rodents, non-human primates and human abusers, but also focuses on the more recent literature associated with reactive oxygen and nitrogen species and their contribution to neuronal death via necrosis and/or apoptosis. The effect of methamphetamine on the mitochondrial membrane potential and electron transport chain and subsequent apoptotic cascades are also emphasized. Finally, we describe potential treatments for methamphetamine abusers with reference to the time after withdrawal. We suggest that potential treatments cart be divided into three categories; (1) the prevention of neurotoxicity if recidivism occurs, (2) amelioration of apoptotic cascades that may occur even in the withdrawal period and (3) treatment of the atypical depression associated with withdrawal.
Article
H]Dihydrotetrabenazine ([3H]DTBZ), a specific ligand for the vesicular monoamine transporter (VMAT2), has been used to characterize the integrity of monoaminergic nerve terminals in experimental animals and humans. The purpose of the present studies was to compare the loss of VMAT2 binding with the loss of other neurochemical markers of the dopamine (DA) nerve terminals in mice treated with neurotoxic doses of methamphetamine (METH) or MPTP. Profound decreases (≥70%) in DA content, tyrosine hydroxylase activity, and [3H]carbomethoxy-3-(4-fluorophenyl)tropane binding to the DA transporter were observed in striatal homogenates at both 1 and 6 days after exposure to the neurotoxins. It is surprising that no significant loss of [3H]DTBZ binding in the homogenates was observed at 1 day after exposure, although a significant loss (-50%) was apparent 6 days later. However, in isolated vesicle preparations, [3H]DTBZ binding and active [3H]DA uptake were markedly reduced (>70%) at 1 day. These observations indicate that vesicle function is compromised at an early time point after exposure to neurotoxic insult. Furthermore, the changes in [3H]DTBZ binding in homogenates may not be a sensitive indicator of early damage to synaptic vesicles, although homogenate binding reliably identifies a loss of VMAT2 at later times.