Abstract
The life span of individuals that are sero-positive for human immunodeficiency virus (HIV) has greatly improved; however, complications involving the central nervous system (CNS) remain a concern. While HIV does not directly infect neurons, the proteins produced by the virus, including HIV transactivator of transcription (Tat), are released from infected glia; these proteins can be neurotoxic. This neurotoxicity is thought to mediate the pathology underlying HIV-associated neurological impairments. Cocaine abuse is common among HIV infected individuals, and this abuse augments HIV-associated neurological deficits. The brain regions and pathophysiological mechanisms that are dysregulated by both chronic cocaine and Tat are the focus of the current review.
Keywords: addiction, calcium channels, Cav1.2, neuropathogenesis, prefrontal cortex
Introduction
Since the beginning of the acquired immunodeficiency syndrome (AIDS) epidemic, there has been a strong link between human immunodeficiency virus (HIV) and drug abuse [1]. Recreational use of psychostimulants, including cocaine, increases the likelihood that the user will engage in impulsive and unsafe behaviors such as risky sexual practices. These risky behaviors increase the chance of exposure to HIV and other diseases [2]. Cocaine is also used to self-medicate for mental health stressors [3], and recent users (i.e., self-reported to have used cocaine within 6 months) and former users (i.e., self-reported to have ever used cocaine) report more depressive symptoms that non-users [4]. Upon learning that they are HIV+, many individuals experience depression-like symptoms including fear, anguish, and thoughts of suicide [5], and these stressors may serve to promote cocaine abuse. Regardless of the original motive for abusing drugs, chronic exposure to these drugs can exacerbate HIV-related disease progression. For example, cocaine increases viral replication [6] and accelerates loss of CD4+ T-cells [7] in the HIV+ individual. This review focuses on the brain, and mechanisms that may contribute to enhanced vulnerability and/or exacerbated neuropathology that can occur during HIV/AIDS and cocaine abuse co-morbidity.
HIV and cocaine in the brain
HIV-associated neurocognitive disorders (HAND) include impaired concentration, memory deficits, and motor impairment [8,9]. These symptoms reflect neuronal damage subsequent to HIV-infection; however, the virus does not directly infect neurons. Damage is inflicted by HIV-infected monocytes, which readily cross the blood-brain barrier (BBB), [10] and subsequently release toxic viral proteins within the central nervous system (CNS) [11,12]. CNS damage occurs early during infection; HIV-1 RNA can be detected in cerebral spinal fluid (CSF) as early as 8 days after the estimated introduction of the virus into the human host [13]. Combination antiretroviral therapies (cART) do not readily cross the BBB, which reduces efficacy for suppressing HIV infection within the CNS [14]. Accordingly, HAND are reported in up to 50% of HIV+ patients, even with cART-controlled viral replication [8,15–17]. The neurocognitive deficiencies associated with HIV are augmented in those that are cocaine abusers [4,18], as cocaine use disrupts BBB integrity and facilitates HIV penetration into the CNS [19,20]. Thus, in spite of cART-controlled infection, the brain of cocaine-abusing HIV+ individuals may be particularly vulnerable to the destructive features of virus.
Regions of the brain that show an accelerated pathology in the co-morbid condition include the hippocampus, striatum, and frontal cortex (including the prefrontal cortex, PFC) [21]. Cocaine addicts (that were not screened for HIV infection) exhibit hypo-activation of the PFC [22]; atrophy and tissue thinning also occurs in the PFC of HIV+ individuals [23,24]. HIV+ men exhibit deficits in attention and executive function during tasks that involve the PFC [25]. Another study conducted with women revealed that blood oxygen level-dependent (BOLD) functional magnetic resonance imaging (fMRI) during verbal learning tasks correlates with severe cognitive impairment and both current and former cocaine using HIV+ women demonstrate reduced PFC activation compared to HIV+ non-users [26]. These outcomes likely reflect a convergence of neuronal dysregulation within the co-morbid brain. Understanding the common sites of action for, and/or maladaptations to, HIV-1 infection and cocaine abuse within the PFC would help decipher the underpinnings of these co-morbid consequences.
Neurological consequences of Tat
Following entry into the CSF/CNS, HIV-1 spreads from monocytes and macrophages to the astrocytes and the resident immune cells of the brain, microglia [10,27]. Microglia and astrocytes play key roles in brain innate immunity and mediate the neuroinflammatory response to HIV [28,29]. The turnover rate for these cells is relatively slow, which provides a potential resevoir for HIV-1 in the brain [30]. HIV-1 expresses at least nine toxic proteins that can be released from infected immune cells, and some of these proteins impair the biochemical and physiological mechanisms of neurons, resulting in dysregulation [11]. Of these proteins, one that is known to be neurotoxic and has been extensively studied is a transactivator of transcription (Tat). Tat is a small, intrinsically flexible and highly basic protein of ~80–103 amino acids (14–16kD) that can bind with high affinity to a large number of diverse partners (proteins or lipids), to form complexes with a wide range of action [31]. Tat drives replication of the HIV-1 via transactivation of the promoter region of the viral genome [32]. Tat also can inhibit proteolysis [33–35], and even a transient exposure to Tat is sufficient to dysregulate neurons [36]. Moreover, the pathological consequences of Tat can persist long after the protein itself is degraded [37], reflecting the ability of Tat to induce the expression of a large variety of cellular genes, interact with cellular proteins and to promote production of pro-inflammatory cytokines and chemokines [38,39]. These characteristics also underlie the ability of Tat to exert its effects far from its site of release [40]. Extracellular Tat can be transported into neurons via the low density lipoprotein receptor (LRP)-mediated endocytotic pathway [41]. Intracellular Tat can disrupt cytoplasmic Ca2+ concentrations either by inducing release from intracellular stores (i.e., disrupting function of 1,4,5-triphosphate receptors) or by enhancing influx of extracellular Ca2+ via voltage- and ligand-gated channels (i.e., voltage-gated Ca2+ channels and N-methyl-D-aspartate receptors (NMDAR)) on the plasma membrane [42,43]. Thus, Tat-induced consequences (including neurotoxicity) are dependent on extracellular Ca2+ influx [44]. The Tat-induced rise in cytoplasmic Ca2+ is followed by mitochondrial Ca2+ uptake and generation of mitochondrial reactive oxygen species (ROS) [45]. Accumulation of ROS can lead to oxidative stress and induce signaling cascades for neuronal apoptosis [45,46]. In addition to direct effects on neurons, Tat also alters the function of glial cells that then contribute to neuronal pathology. HIV-1 Tat promotes reactive gliosis and secretion of pro-inflammatory cytokines. This action reduces neuronal synaptic densities to simplify synaptic connections, leading to neuronal injury [40,47–54]. Tat concentrations in the CSF of HIV+ patients range from 10–1000nM [55–57]. These concentrations of Tat in vitro can be neurotoxic [36]. Accordingly, Tat-induced neurotoxicity is thought to be a major contributor to HIV-related neuropathogenesis.
Neurological consequences of chronic cocaine
Cocaine binds to monoamine, (i.e., serotonin, dopamine and norepinephrine) transporters on nerve terminals and prevents the re-uptake of the respective neurotransmitters from the synapse. As a result, synaptic concentrations of neurotransmitters are dramatically increased, enhancing the activation of synaptic and peri-synaptic receptors. The chronic presence of cocaine decreases the sensitivity of postsynaptic neurons to monoamines (i.e. dopamine in the nucleus accumbens (NAc)) and increases the sensitivity of dopamine auto-receptors, resulting in a functional loss of dopamine in the synapse [58,59]. During periods of repeated drug exposure, reduced neurotransmitter production, receptor internalization, and synaptic simplification can occur in order to compensate for the exaggerated activation of the mesocorticolimbic system. These maladaptations endure long after terminating cocaine exposure.
The mesocorticolimbic system is critically involved in the motivational aspects of drug abuse including reward, and salience attribution to drug-associated cues. Neurons in the medial prefrontal cortex (mPFC) provide excitatory input to the NAc [60], ventral pallidum (VP) [60,61], and ventral tegmental area (VTA) [60,62], key brain regions involved in addictions. Chronic cocaine increases the excitatory drive from the mPFC and enhances activity of dopaminergic cells, which are linked with reward-associated, cue-elicited behaviors [63–65]. Specifically, projections from the infralimbic (IL) and prelimbic (PrL) subregions of the mPFC terminate within the NAc shell and NAc core, respectively [60]. Activation of the PrL to NAc core projections promotes drug-seeking, whereas activation of IL to NAc shell projections inhibits drug-seeking behaviors [65,66]. Cocaine, but not food, engages projections originating in the PrL mPFC and increases glutamate levels in the NAc core, to stimulate reinstatement of responding in rats [66]. These findings suggest that the cocaine-induced rise in NAc core glutamate levels is abnormal rather than physiological. Rats that self-administer cocaine in the presence of discrete cues show drug-seeking in the presence of these cues during periods of withdrawal [67,68] likely reflecting persistent molecular neuroadaptations in the mPFC [69,70]. Functional imaging performed on abstinent cocaine addicts reveal hyper-reactivity of the mPFC that is strongly correlated with self-reports of craving for cocaine during the presentation of cocaine-associated cues [22,71]. Therefore, mPFC-associated circuitry is thought to underpin the abnormally enhanced reactivity to drug-associated cues experienced by addicted individuals during periods of drug abstinence.
Common pathways for cocaine and Tat effects in mPFC
The high prevalence of cocaine abuse in HIV+ individuals and the ability of cocaine addiction to exacerbate HAND indicates that chronic cocaine and HIV-1 toxic proteins may present overlapping neuropathological events in brain regions that govern HAND and reward-motivated behaviors. Using Tat as a prototype, we overview here current evidence for such an overlap, and explain the mechanisms that contribute to the neuropathological convergence.
Indirect effects of Tat and cocaine on neurons
In a healthy CNS, astrocytes outnumber neurons five-fold. Astrocytes envelope neuronal synapses to aid many essential neuronal functions that are critical for normal synaptic transmission [72], including the maintenance of fluid, ion, pH, and transmitter homeostasis of the synaptic interstitial fluid [73,74]. Insults to the CNS induce reactive astrogliosis, a process which has become a pathological hallmark for numerous neurological diseases [73]. Activated astrocytes lose the capacity to maintain synaptic homeostasis of glutamate, resulting in excessive glutamate-induced calcium influx which when excessive can be neurotoxic [74].
Tat exerts a powerful and persistent influence on astrocytes. In vitro studies revealed that astrogliosis can occur within 7 days of exposure to Tat [50]. Extending these studies to the mammalian brain, we assessed astrogliosis in the mPFC after Tat administration [75]. Glial fibrillary acidic protein (GFAP) is a structural protein in astrocytes, and increased GFAP expression signifies astrogliosis. GFAP expression was increased in the mPFC 14 days after a single intracerebroventricular (i.c.v.) injection of Tat [75]. Tat up-regulates GFAP expression through direct interaction with the enhancer elements activator protein 1 (AP1) and specificity protein 1 (SP1). The transcription factor early growth response-1 (Egr-1) regulates p300 transcription, and deletion of the Egr-1 cis-transacting element within the p300 promoter abolishes Tat-induced GFAP expression [76]. Though the mechanisms are less clear for cocaine, the psychostimulant is known to increase GFAP immunoreactivity in the PFC (and nucleus accumbens) following repeated administration and 3 weeks of withdrawal [77]. Astrogliosis caused by cocaine would diminish the capacity of astrocytes to support neuronal health and in so doing, enhance neuron vulnerability to the detrimental consequences of Tat.
Direct effects of Tat and cocaine on neurons
The ability of cocaine to bind to and block the dopamine transporter (DAT) is well established. Recent studies have revealed that Tat also can bind directly to DAT [78,79]. Tat binding to DAT inhibits the reuptake of dopamine released from activated neurons [78] and promotes DAT internalization [80]. Tat-induced conformational changes in DAT also increase the affinity of cocaine for the transporter protein [78]. Thus, Tat has the capacity to significantly enhance extracellular concentrations of dopamine through multiple mechanisms [81]. Excessive extracellular dopamine over-actives dopaminergic receptors and drives the reward-motivation known to underlie cocaine-mediated behaviors. Recent studies indicate that a similar outcome occurs following exposure to Tat, which can potentiate behaviors mediated by cocaine reward [82].
In the frontal cortex and striatum of rats, acute or repeated administration of cocaine alters mitochondrial complex I subunits (i.e. nicotinamide adenine dinucleotide-4) to decrease oxidative phosphorylation and increase ROS production [83]. These cocaine-mediated effects occur without the loss of neurons. The absence of neuronal death may reflect ROS buffering as cocaine causes a concomitant increase in the production of antioxidant enzymes which prevents the activation of cellular apoptotic signaling cascades [83]. In contrast, Tat induces ROS production that can result in death of cultured neurons; cocaine enhances this effect [84]. These reports indicate that cocaine may enhance the Tat-induced rise in cytoplasmic Ca2+ and mitochondrial Ca2+ uptake leading to the generation of mitochondrial ROS, and that excessive intracellular Ca2+ may be a common mechanism by which cocaine abuse accelerates HIV+ associated neuropathology.
Evaluations of Ca2+ function in neurons indicate that this mechanism is a site of convergence for chronic cocaine and Tat. Removal of extracellular Ca2+ decreases Tat-induced Ca2+ influx and toxicity of cultured rat cortical neurons [44,85]. Enhanced Ca2+ influx is regulated, at least in part, by NMDAR [42,86,87] and voltage-gated Ca2+ channels (likely the L-type Ca2+ channels) [88]. Tat (1–500 nM) dose-dependently induces a fast, transient increase of Ca2+ influx in cultured rat cortical neurons, which is not blocked by inhibition of NMDAR [85]. In contrast, blockade of the L-channels significantly reduces Tat-induced Ca2+ influx and neurotoxicity in cultured human fetal microglia and monocytes [89,90]. Tat increases neuronal excitability in mPFC pyramidal neurons, and this increase can be inhibited by diltiazem, an L-type Ca2+ channel blocker (Fig 2). Physiological low nanomolar concentrations (~15–20nM) of Tat delivered via i.c.v. injection enhances L-channel expression in the mPFC of rats as compared to vehicle injected controls (Fig. 3). The major advantage of using the i.c.v. injection is that a known concentration of Tat can be introduced into the brain, bypassing the BBB. The resulting CSF concentrations in these rats are within the range of the Tat CSF concentrations detected in humans [55–57]. Tat-induced upregulation of these channels could abnormally increase Ca2+ influx and intracellular Ca2+ levels, enhancing reactivity of mPFC pyramidal cells to membrane depolarization. This phenomenon is present for up to 14 days after Tat administration, and it overlaps with the time-course of L-channel plasticity during cocaine withdrawal [91]. After a prolonged (2–3 weeks) withdrawal from repeated, experimenter administered cocaine, surface expression of the L-type Ca2+ channels has been shown to be significantly increased in the mPFC [91].
Fig. 2.

Tat increases evoked firing of rat mPFC pyramidal neurons in a Ca2+-dependent manner. Curves illustrate the number of action potentials (spikes) evoked by depolarizing current pulses at pretreatment baseline (SAL-BlCtr; open circles) is significantly increased following exposure to 40nM of recombinant Tat (SAL-Tat; filled circles) and this effect was blocked by co-perfusion with an antagonist of open state L-Type Ca2+ channels, diltiazem (40nM; SAL-Tat/Dilt; filled squares). Two-way rmANOVA; p<0.01 for treatment history effect, current effect, and the interaction with post hoc Newman-Keuls *p<0.025, **p<0.01. SAL refers to the ip injections of saline that these rats received prior to killing and harvesting the brain slices. Reprinted with permission from Napier et al. JNIP. 1 June 2014-Vol 9-Issue 3-p354–368. Springer US©.
Fig. 3.

Tat increases Cav1.2-α1c immuno-reactive cells in the rat mPFC. Shown is staining from representative brain sections taken 14 days after a single intracerebroventricular injection of Tat (80μg/20μl; bottom) or PBS (20μl; top). Left are brain sections containing the mPFC and MC photographed at 1.25× (scale bar=1mm). Right are representative 20× magnification images of the mPFC and the location for which are indicated with a box on the 1.25× images (scale bar=100μm). Examples of Cav1.2-α1c immuno-reactive cells are indicated by the arrows. Reprinted with permission from Wayman et al. Neuroreport. 3 October 2012-Vol 23-Issue 14-p825–829. Wolters Kluwer Health Lippincott Williams & Wilkins©.
We recently reported that in vitro application of Tat, while recording mPFC pyramidal neurons from rat brain slices, dramatically increases firing and NMDAR-independent Ca2+ influx through L-type Ca2+ channels [88]. Tat enhances Ca2+ influx at concentrations between 10–40nM, the magnitude of which correlates with Tat concentration. Elevated intracellular Ca2+, under normal conditions, will bind to calmodulin to inactivate L-type Ca2+ channels. Recovery from inactivation occurs as the membrane becomes re-polarized or hyperpolarized [92]. Tat can bind calmodulin [93] and inhibit Ca2+-dependent inactivation of the channel, which would allow for excessive influx of Ca2+ into the neuron. A history of repeated, non-contingent cocaine treatment (i.e., cocaine administered by the experimenter) increases neuronal firing that is augmented by 40nM Tat (Fig.4A). Additionally, cocaine treatment enhances Ca2+ influx, and acute exposure to 10nM Tat enhances this effect (Fig.4B). Rats that self-administer or self-titrate their dose of cocaine (similar to the human scenario) exhibit pyramidal neuron firing that is greater than that recorded from saline-yoked (SAL-Yoked) controls [94]. Pyramidal neurons from rats that self-administered cocaine (COC-SA) exhibit Tat-induced increases in firing at much lower concentrations of Tat (e.g., 5nM) [94], and these neurons are more easily driven to over-excitation, wherein firing neurons become inhibited with low Tat concentrations (Fig 5). The neurons recorded in these studies [94] were in cortical layers 5/6, had a pyramidal morphological profile, prominent apical dendrite, and exhibited regular firing patterns. These characteristics are consistent with glutamatergic pyramidal neurons [95]. GABAergic interneurons typically feature multiple small apical dendrites and exhibit irregular firing patterns [96]. While literature on L-type Ca2+ channel expression in prefrontal GABAergic interneurons is limited, hippocampal interneurons have been shown to express L-type Ca2+ channels [97]. However, it is important to note that any effect of Tat-induced upregulation of L-type Ca2+ channels on mPFC interneurons is overridden by the increased excitability of the pyramidal neurons. Thus, L–type Ca2+ channels may be one of the common pathways for cocaine- and Tat-induced neuroadaptations in mPFC pyramidal neurons, and this convergence may explain how cocaine can accelerate the onset and increase the severity of neurocognitive impairment in HIV+ individuals.
Fig. 4.

Repeated non-contingent administration of cocaine to rats enhanced Tat-induced firing and Ca2+ potentials recorded from mPFC pyramidal neurons ex vivo. (A.) Curves illustrate that the number of action potentials (spikes) evoked by depolarizing currents is markedly greater in a mPFC pyramidal neurons from cocaine (COC)-treated rats (open squares) than that recorded from saline (SAL)-treated rats (open circles). BlCtr, baseline control. Bath-applied Tat (40nM) facilitated the evoked firing in both neurons from SAL-treated rats (filled circles) and COC-treated rats (open squares). Data are presented as mean ± S.E.M. Two-way rmANOVA, p<0.05 for treatment histry effect, current effect, and the interaction. Post hoc Newman-Keuls illustrated as follows: *p < 0.025 and **p < 0.01, compared to SAL-BlCtr; #p < 0.025 and ##p < 0.01, compared with SAL-Tat; ^p < 0.025, compared to COC-BlCtr. (B.) At baseline, the duration of Ca2+ potentials were remarkably prolonged in an mPFC pyramidal neuron from a COC-treated rat (COC-BlCtr) as compared to a neuron from a SAL-treated rat (SAL-BlCtr). Bath-applied Tat (10nM) enhanced the Ca2+ potential duration in a neuron from a SAL-treated rat (SAL-Tat) and exaggerated the enhanced Ca2+ potential observed in the neuron from a cocaine-treated rat (COC-Tat). Reprinted with permission from Napier et al. JNIP. 1 June 2014-Vol 9-Issue 3-p354–368. Springer US©.
Fig. 5.

The ability of Tat to excessively activate neurons is greater in rats that self-administer cocaine. Shown is the number of pyramidal neurons that exhibited a Tat-induced “excitotoxicity” as defined by an abnormal action potential profile associated with a reduction in spike number during membrane depolarization. Two cocaine administration profiles were compared, non-contingent and self-administration (SA). Chi-square test revealed that the pyramidal neurons from adolescent rats that received experimenter delivered cocaine were less sensitive to 40nM Tat than adult COC-SA rats were when exposed to 10nM Tat (χ2(1)=8.03, p<0.005). This indicates that SA (and/or the adult brain state) makes mPFC pyramidal neurons more sensitive to the toxic effects of Tat.
The consequences of cocaine and Tat on L-type Ca2+ channels in the mPFC appear to be synergistic. When both cocaine and Tat are present, the effects are greater than additive and this is confirmed by interaction statistics [88]. Cocaine itself is not lethal to neurons, but cocaine does alter neuronal function and responses from glia. The consequences from exposure to low nanomolar concentrations of Tat also occurred in the absence of a significant loss of neurons [75], suggesting that Tat has not yet produced the neurotoxicity that is observed in other studies. However, these same concentrations of Tat (that normally would be subthreshold for evoking neuronal pathology of pathophysiology) may become toxic in the presence of cocaine.
Conclusion
In summary, we extend in vitro work with Tat and cocaine to in vivo work and demonstrate that the changes may be long lasting even if Tat is no longer present. These changes reflect alterations to glial cells, which may set the stage for a hostile environment and increase the vulnerability of neurons to become damaged or dysfunction. We identify that alterations to L-channel expression and function as common mechanism between the pathophysiology arising from exposure to the Tat protein and cocaine. The neuroadaptations that occur in pyramidal neurons within the mPFC as a result of repeated cocaine exposure may potentiate the acute effects of Tat, including upregulated L-channel function and increased neuron excitability. In addition, rats that self-administer cocaine appear to be more sensitive to the enhanced excitability induced by cocaine and the acute effects of Tat. Taken together L-channels are a mechanism that may contribute to an enhanced vulnerability and/or exacerbated neuropathology during HIV/AIDS and cocaine abuse co-morbidity.
Fig. 1.

Tat up-regulates GFAP expression in the rat mPFC. Shown is GFAP staining from representative brain sections taken 14 days after a single intracerebroventricular injection of Tat (80μg/20μl) (bottom) or the PBS vehicle (20μl) (top). Left are brain sections containing the medial prefrontal cortex (mPFC) and motor cortex (MC) photographed at 1.25× (scale bar=1mm). Right are representative 20× magnification images of the mPFC and the location for which are indicated with a box on the 1.25× images (scale bar=100μm). Reprinted with permission from Wayman et al. Neuroreport. 3 October 2012-Vol 23-Issue 14-p825–829. Wolters Kluwer Health Lippincott Williams & Wilkins©.
Acknowledgments
WNW, LC, ALP and TCN participated in the research design; WNW, LC, and ALP acquired the data, WNW and LC performed data analysis, and WNW, LC, ALP, and TCN contributed to the writing of the manuscript. The authors thank Drs. Jingli Zhang and Xiu-Ti Hu for their contributions to some of the studies that are overviewed in this manuscript.
Footnotes
Conflict of Interest
The authors do not have a financial relationship with the funding organizations, and have no conflicts of interest to report. Studies overviewed in this manuscript were supported in part by USPHSG’s DA033206 and DA033882, the Center for Compulsive Behavior and Addiction at Rush University Medical Center, and the Chicago Developmental Center for AIDS Research.
Reference List
- 1.World Health Organization, Joint United Nations Programme on HIV/AIDS. AIDS epidemic update 2009. 2009 [Google Scholar]
- 2.National Institue on Drug Abuse. DrugFacts: HIV/AIDS and Drug Abuse: Intertwined Epidemics. 2012 [Google Scholar]
- 3.Abraham HD, Fava M. Order of onset of substance abuse and depression in a sample of depressed outpatients. Compr Psychiatry. 1999;40:44–50. doi: 10.1016/s0010-440x(99)90076-7. [DOI] [PubMed] [Google Scholar]
- 4.Meyer VJ, Rubin LH, Martin E, Weber KM, Cohen MH, Golub ET, et al. HIV and recent illicit drug use interact to affect verbal memory in women. J Acquir Immune Defic Syndr. 2013;63:67–76. doi: 10.1097/QAI.0b013e318289565c. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Stevens PE, Tighe DB. Trauma of discovery: women’s narratives of being informed they are HIV-infected. AIDS Care. 1997;9:523–538. doi: 10.1080/713613201. [DOI] [PubMed] [Google Scholar]
- 6.Mantri CK, Pandhare DJ, Mantri JV, Dash CC. Cocaine enhances HIV-1 replication in CD4+ T cells by down-regulating MiR-125b. PLoS ONE. 2012;7:e51387. doi: 10.1371/journal.pone.0051387. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Samikkannu T, Rao KV, Arias AY, Kalaichezian A, Sagar V, Yoo C, et al. HIV infection and drugs of abuse: role of acute phase proteins. J Neuroinflammation. 2013;10:113. doi: 10.1186/1742-2094-10-113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Tozzi V, Balestra P, Bellagamba R, Corpolongo A, Salvatori MF, Visco-Comandini U, et al. Persistence of neuropsychologic deficits despite long-term highly active antiretroviral therapy in patients with HIV-related neurocognitive impairment: prevalence and risk factors. J Acquir Immune Defic Syndr. 2007;45:174–182. doi: 10.1097/QAI.0b013e318042e1ee. [DOI] [PubMed] [Google Scholar]
- 9.Antinori A, Arendt G, Becker JT, Brew BJ, Byrd DA, Cherner M, et al. Updated research nosology for HIV-associated neurocognitive disorders. Neurology. 2007;69:1789–1799. doi: 10.1212/01.WNL.0000287431.88658.8b. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Gonzalez-Scarano F, Martin-Garcia J. The neuropathogenesis of AIDS. Nat Rev Immunol. 2005;5:69–81. doi: 10.1038/nri1527. [DOI] [PubMed] [Google Scholar]
- 11.Jones G, Power C. Regulation of neural cell survival by HIV-1 infection. Neurobiol Dis. 2006;21:1–17. doi: 10.1016/j.nbd.2005.07.018. [DOI] [PubMed] [Google Scholar]
- 12.Nath A. Human immunodeficiency virus (HIV) proteins in neuropathogenesis of HIV dementia. J Infect Dis. 2002;186:S193–S198. doi: 10.1086/344528. [DOI] [PubMed] [Google Scholar]
- 13.Valcour V, Chalermchai T, Sailasuta N, Marovich M, Lerdlum S, Suttichom D, et al. Central nervous system viral invasion and inflammation during acute HIV infection. J Infect Dis. 2012;206:275–282. doi: 10.1093/infdis/jis326. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Yilmaz A, Verhofstede C, D’Avolio A, Watson V, Hagberg L, Fuchs D, et al. Treatment intensification has no effect on the HIV-1 central nervous system infection in patients on suppressive antiretroviral therapy. J Acquir Immune Defic Syndr. 2010;55:590–596. doi: 10.1097/QAI.0b013e3181f5b3d1. [DOI] [PubMed] [Google Scholar]
- 15.McArthur JC, Steiner J, Sacktor N, Nath A. Human immunodeficiency virus-associated neurocognitive disorders: Mind the gap. Ann Neurol. 2010;67:699–714. doi: 10.1002/ana.22053. [DOI] [PubMed] [Google Scholar]
- 16.Heaton RK, Clifford DB, Franklin DR, Jr, Woods SP, Ake C, Vaida F, et al. HIV-associated neurocognitive disorders persist in the era of potent antiretroviral therapy: CHARTER Study. Neurology. 2010;75:2087–2096. doi: 10.1212/WNL.0b013e318200d727. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Heaton RK, Franklin DR, Ellis RJ, McCutchan JA, Letendre SL, Leblanc S, et al. HIV-associated neurocognitive disorders before and during the era of combination antiretroviral therapy: differences in rates, nature, and predictors. J Neurovirol. 2011;17:3–16. doi: 10.1007/s13365-010-0006-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Byrd DA, Fellows RP, Morgello S, Franklin D, Heaton RK, Deutsch R, et al. Neurocognitive impact of substance use in HIV infection. J Acquir Immune Defic Syndr. 2011;58:154–162. doi: 10.1097/QAI.0b013e318229ba41. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Fiala M, Singer EJ, Commins D, Mirzapoiazova T, Verin A, Espinosa A, et al. HIV-1 Antigens in Neurons of Cocaine-Abusing Patients. Open Virol J. 2008;2:24–31. doi: 10.2174/1874357900802010024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Fiala M, Eshleman AJ, Cashman J, Lin J, Lossinsky AS, Suarez V, et al. Cocaine increases human immunodeficiency virus type 1 neuroinvasion through remodeling brain microvascular endothelial cells. J Neurovirol. 2005;11:281–291. doi: 10.1080/13550280590952835. [DOI] [PubMed] [Google Scholar]
- 21.Ferris MJ, Mactutus CF, Booze RM. Neurotoxic profiles of HIV, psychostimulant drugs of abuse, and their concerted effect on the brain: Current status of dopamine system vulnerability in NeuroAIDS. Neurosci Biobehav Rev. 2008;32:883–909. doi: 10.1016/j.neubiorev.2008.01.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Goldstein RZ, Volkow ND. Drug addiction and its underlying neurobiological basis: neuroimaging evidence for the involvement of the frontal cortex. Am J Psychiatry. 2002;159:1642–1652. doi: 10.1176/appi.ajp.159.10.1642. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Wiley CA, Masliah E, Morey M, Lemere C, DeTeresa R, Grafe M, et al. Neocortical damage during HIV infection. Ann Neurol. 1991;29:651–657. doi: 10.1002/ana.410290613. [DOI] [PubMed] [Google Scholar]
- 24.Thompson PM, Dutton RA, Hayashi KM, Toga AW, Lopez OL, Aizenstein HJ, et al. Thinning of the cerebral cortex visualized in HIV/AIDS reflects CD4+ T lymphocyte decline. Proc Natl Acad Sci U S A. 2005;102:15647–15652. doi: 10.1073/pnas.0502548102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Su T, Schouten J, Geurtsen GJ, Wit FW, Stolte IG, Prins M, et al. Multivariate normative comparison, a novel method for more reliably detecting cognitive impairment in HIV infection. AIDS. 2015 doi: 10.1097/QAD.0000000000000573. [DOI] [PubMed] [Google Scholar]
- 26.Meyer VJ, Little DM, Fitzgerald DA, Sundermann EE, Rubin LH, Martin EM, et al. Crack cocaine use impairs anterior cingulate and prefrontal cortex function in women with HIV infection. J Neurovirol. 2014;20:352–361. doi: 10.1007/s13365-014-0250-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Hauser KF, El-Hage N, Stiene-Martin A, Maragos WF, Nath A, Persidsky Y, et al. HIV-1 neuropathogenesis: glial mechanisms revealed through substance abuse. J Neurochem. 2007;100:567–586. doi: 10.1111/j.1471-4159.2006.04227.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Hong S, Banks WA. Role of the immune system in HIV-associated neuroinflammation and neurocognitive implications. Brain Behav Immun. 2014 doi: 10.1016/j.bbi.2014.10.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Garden GA. Microglia in human immunodeficiency virus-associated neurodegeneration. Glia. 2002;40:240–251. doi: 10.1002/glia.10155. [DOI] [PubMed] [Google Scholar]
- 30.Lassmann H, Schmied M, Vass K, Hickey WF. Bone marrow derived elements and resident microglia in brain inflammation. Glia. 1993;7:19–24. doi: 10.1002/glia.440070106. [DOI] [PubMed] [Google Scholar]
- 31.Debaisieux S, Rayne F, Yezid H, Beaumelle B. The ins and outs of HIV-1 Tat. Traffic. 2012;13:355–363. doi: 10.1111/j.1600-0854.2011.01286.x. [DOI] [PubMed] [Google Scholar]
- 32.Rosen CA, Terwilliger E, Dayton A, Sodroski JG, Haseltine WA. Intragenic cis-acting art gene-responsive sequences of the human immunodeficiency virus. Proc Natl Acad Sci U S A. 1988;85:2071–2075. doi: 10.1073/pnas.85.7.2071. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Li W, Li G, Steiner J, Nath A. Role of Tat protein in HIV neuropathogenesis. Neurotox Res. 2009;16:205–220. doi: 10.1007/s12640-009-9047-8. [DOI] [PubMed] [Google Scholar]
- 34.Seeger M, Ferrell K, Frank R, Dubiel W. HIV-1 tat inhibits the 20 S proteasome and its 11 S regulator-mediated activation. J Biol Chem. 1997;272:8145–8148. doi: 10.1074/jbc.272.13.8145. [DOI] [PubMed] [Google Scholar]
- 35.Huang X, Seifert U, Salzmann U, Henklein P, Preissner R, Henke W, et al. The RTP site shared by the HIV-1 Tat protein and the 11S regulator subunit alpha is crucial for their effects on proteasome function including antigen processing. J Mol Biol. 2002;323:771–782. doi: 10.1016/s0022-2836(02)00998-1. [DOI] [PubMed] [Google Scholar]
- 36.Agrawal L, Louboutin JP, Strayer DS. Preventing HIV-1 Tat-induced neuronal apoptosis using antioxidant enzymes: mechanistic and therapeutic implications. Virology. 2007;363:462–472. doi: 10.1016/j.virol.2007.02.004. [DOI] [PubMed] [Google Scholar]
- 37.Nath A, Haughey NJ, Jones M, Anderson C, Bell JE, Geiger JD. Synergistic neurotoxicity by human immunodeficiency virus proteins Tat and gp120: protection by memantine. Ann Neurol. 2000;47:186–194. [PubMed] [Google Scholar]
- 38.Li L, Dahiya S, Kortagere S, Aiamkitsumrit B, Cunningham D, Pirrone V, et al. Impact of Tat Genetic Variation on HIV-1 Disease. Adv Virol. 2012;2012:123605. doi: 10.1155/2012/123605. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Nath A, Conant K, Chen P, Scott C, Major E. Transient exposure to HIV-1 Tat protein results in cytokine production in macrophages and astrocytes: A hit and run phenomenon. J Biol Chem. 1999;274:17098–17102. doi: 10.1074/jbc.274.24.17098. [DOI] [PubMed] [Google Scholar]
- 40.Bruce-Keller AJ, Chauhan A, Dimayuga FO, Gee J, Keller JN, Nath A. Synaptic transport of human immunodeficiency virus-Tat protein causes neurotoxicity and gliosis in rat brain. J Neurosci. 2003;23:8417–8422. doi: 10.1523/JNEUROSCI.23-23-08417.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Liu Y, Jones M, Hingtgen CM, Bu G, Laribee N, Tanzi RE, et al. Uptake of HIV-1 tat protein mediated by low-density lipoprotein receptor-related protein disrupts the neuronal metabolic balance of the receptor ligands. Nat Med. 2000;6:1380–1387. doi: 10.1038/82199. [DOI] [PubMed] [Google Scholar]
- 42.Haughey NJ, Nath A, Mattson MP, Slevin JT, Geiger JD. HIV-1 Tat through phosphorylation of NMDA receptors potentiates glutamate excitotoxicity. J Neurochem. 2001;78:457–467. doi: 10.1046/j.1471-4159.2001.00396.x. [DOI] [PubMed] [Google Scholar]
- 43.Haughey NJ, Holden CP, Nath A, Geiger JD. Involvement of inositol 1,4,5-trisphosphate-regulated stores of intracellular calcium in calcium dysregulation and neuron cell death caused by HIV-1 protein tat. J Neurochem. 1999;73:1363–1374. doi: 10.1046/j.1471-4159.1999.0731363.x. [DOI] [PubMed] [Google Scholar]
- 44.Cheng J, Nath A, Knudsen B, Hochman S, Geiger JD, Ma M, et al. Neuronal excitatory properties of human immunodeficiency virus type 1 Tat protein. Neuroscience. 1998;82:97–106. doi: 10.1016/s0306-4522(97)00174-7. [DOI] [PubMed] [Google Scholar]
- 45.Kruman II, Nath A, Mattson MP. HIV-1 protein Tat induces apoptosis of hippocampal neurons by a mechanism involving caspase activation, calcium overload, and oxidative stress. Exp Neurol. 1998;154:276–288. doi: 10.1006/exnr.1998.6958. [DOI] [PubMed] [Google Scholar]
- 46.New DR, Ma M, Epstein LG, Nath A, Gelbard HA. Human immunodeficiency virus type 1 Tat protein induces death by apoptosis in primary human neuron cultures. J Neurovirol. 1997;3:168–173. doi: 10.3109/13550289709015806. [DOI] [PubMed] [Google Scholar]
- 47.Jones M, Olafson K, Del Bigio MR, Peeling J, Nath A. Intraventricular injection of human immunodeficiency virus type 1 (HIV-1) tat protein causes inflammation, gliosis, apoptosis, and ventricular enlargement. J Neuropathol Exp Neurol. 1998;57:563–570. doi: 10.1097/00005072-199806000-00004. [DOI] [PubMed] [Google Scholar]
- 48.Bansal AK, Mactutus CF, Nath A, Maragos W, Hauser KF, Booze RM. Neurotoxicity of HIV-1 proteins gp120 and Tat in the rat striatum. Brain Res. 2000;879:42–49. doi: 10.1016/s0006-8993(00)02725-6. [DOI] [PubMed] [Google Scholar]
- 49.Aksenov MY, Hasselrot U, Bansal AK, Wu G, Nath A, Anderson C, et al. Oxidative damage induced by the injection of HIV-1 Tat protein in the rat striatum. Neurosci Lett. 2001;305:5–8. doi: 10.1016/s0304-3940(01)01786-4. [DOI] [PubMed] [Google Scholar]
- 50.Aksenov MY, Hasselrot U, Wu G, Nath A, Anderson C, Mactutus CF, et al. Temporal relationships between HIV-1 Tat-induced neuronal degeneration, OX-42 immunoreactivity, reactive astrocytosis, and protein oxidation in the rat striatum. Brain Res. 2003;987:1–9. doi: 10.1016/s0006-8993(03)03194-9. [DOI] [PubMed] [Google Scholar]
- 51.Orsini MJ, Debouck CM, Webb CL, Lysko PG. Extracellular human immunodeficiency virus type 1 Tat protein promotes aggregation and adhesion of cerebellar neurons. J Neurosci. 1996;16:2546–2552. doi: 10.1523/JNEUROSCI.16-08-02546.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Chauhan A, Turchan J, Pocernich C, Bruce-Keller A, Roth S, Butterfield DA, et al. Intracellular human immunodeficiency virus Tat expression in astrocytes promotes astrocyte survival but induces potent neurotoxicity at distant sites via axonal transport. J Biol Chem. 2003;278:13512–13519. doi: 10.1074/jbc.M209381200. [DOI] [PubMed] [Google Scholar]
- 53.Eugenin EA, King JE, Nath A, Calderon TM, Zukin RS, Bennett MV, et al. HIV-tat induces formation of an LRP-PSD-95- NMDAR-nNOS complex that promotes apoptosis in neurons and astrocytes. Proc Natl Acad Sci U S A. 2007;104:3438–3443. doi: 10.1073/pnas.0611699104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Hahn YK, Podhaizer EM, Farris SP, Miles MF, Hauser KF, Knapp PE. Effects of chronic HIV-1 Tat exposure in the CNS: heightened vulnerability of males versus females to changes in cell numbers, synaptic integrity, and behavior. Brain Struct Funct. 2013 doi: 10.1007/s00429-013-0676-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Xiao H, Neuveut C, Tiffany HL, Benkirane M, Rich EA, Murphy PM, et al. Selective CXCR4 antagonism by Tat: implications for in vivo expansion of coreceptor use by HIV-1. Proc Natl Acad Sci U S A. 2000;97:11466–11471. doi: 10.1073/pnas.97.21.11466. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Westendorp MO, Frank R, Ochsenbauer C, Stricker K, Dhein J, Walczak H, et al. Sensitization of T cells to CD95-mediated apoptosis by HIV-1 Tat and gp120. Nature. 1995;375:497–500. doi: 10.1038/375497a0. [DOI] [PubMed] [Google Scholar]
- 57.Campbell GR, Loret EP. What does the structure-function relationship of the HIV-1 Tat protein teach us about developing an AIDS vaccine? Retrovirology. 2009;6:50. doi: 10.1186/1742-4690-6-50. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Henry DJ, Greene MA, White FJ. Electrophysiological effects of cocaine in the mesoaccumbens dopamine system: Repeated administration. J Pharmacol Exp Ther. 1989;251:833–839. [PubMed] [Google Scholar]
- 59.Dwoskin LP, Peris J, Yasuda RP, Philpott K, Zahniser NR. Repeated cocaine administration results in supersensitivity of striatal D-2 dopamine autoreceptors to pergolide. Life Sci. 1988;42:255–262. doi: 10.1016/0024-3205(88)90634-0. [DOI] [PubMed] [Google Scholar]
- 60.Sesack SR, Deutch AY, Roth RH, Bunney BS. Topographical organization of the efferent projections of the medial prefrontal cortex in the rat: An anterograde tract-tracing study with phaseolus vulgaris leucoagglutinin. J comp Neurol. 1989;290:213–242. doi: 10.1002/cne.902900205. [DOI] [PubMed] [Google Scholar]
- 61.Fuller TA, Russchen FT, Price JL. Sources of presumptive glutamergic/aspartergic afferents to the rat ventral striatopallidal region. J Comp Neurol. 1987;258:317–338. doi: 10.1002/cne.902580302. [DOI] [PubMed] [Google Scholar]
- 62.Christie MJ, Bridge S, James LB, Beart PM. Excitotoxin lesions suggest an aspartatergic projection from rat medial prefrontal cortex to ventral tegmental area. Brain Res. 1985;333:169–172. doi: 10.1016/0006-8993(85)90140-4. [DOI] [PubMed] [Google Scholar]
- 63.Anderson SM, Famous KR, Sadri-Vakili G, Kumaresan V, Schmidt HD, Bass CE, et al. CaMKII: a biochemical bridge linking accumbens dopamine and glutamate systems in cocaine seeking. Nat Neurosci. 2008;11:344–353. doi: 10.1038/nn2054. [DOI] [PubMed] [Google Scholar]
- 64.Ishikawa A, Ambroggi F, Nicola SM, Fields HL. Dorsomedial prefrontal cortex contribution to behavioral and nucleus accumbens neuronal responses to incentive cues. J Neurosci. 2008;28:5088–5098. doi: 10.1523/JNEUROSCI.0253-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Peters J, Kalivas PW, Quirk GJ. Extinction circuits for fear and addiction overlap in prefrontal cortex. Learn Mem. 2009;16:279–288. doi: 10.1101/lm.1041309. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.McFarland K, Lapish CC, Kalivas PW. Prefrontal glutamate release into the core of the nucleus accumbens mediates cocaine-induced reinstatement of drug-seeking behavior. Journal of Neuroscience. 2003;23:3531–3537. doi: 10.1523/JNEUROSCI.23-08-03531.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Grimm JW, Hope BT, Wise RA, Shaham Y. Neuroadaptation. Incubation of cocaine craving after withdrawal. Nature. 2001;412:141–142. doi: 10.1038/35084134. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Weiss F, Martin-Fardon R, Ciccocioppo R, Kerr TM, Smith DL, Ben-Shahar O. Enduring resistance to extinction of cocaine-seeking behavior induced by drug-related cues. Neuropsychopharmacology. 2001;25:361–372. doi: 10.1016/S0893-133X(01)00238-X. [DOI] [PubMed] [Google Scholar]
- 69.Anastasio NC, Stutz SJ, Fox RG, Sears RM, Emeson RB, DiLeone RJ, et al. Functional status of the serotonin 5-HT2C receptor (5-HT2CR) drives interlocked phenotypes that precipitate relapse-like behaviors in cocaine dependence. Neuropsychopharmacology. 2014;39:370–382. doi: 10.1038/npp.2013.199. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Ciccocioppo R, Sanna PP, Weiss F. Cocaine-predictive stimulus induces drug-seeking behavior and neural activation in limbic brain regions after multiple months of abstinence: reversal by D(1) antagonists. Proc Natl Acad Sci U S A. 2001;98:1976–1981. doi: 10.1073/pnas.98.4.1976. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Childress AR, Mozley PD, McElgin W, Fitzgerald J, Reivich M, O’Brien CP. Limbic activation during cue-induced cocaine craving. Am J Psychiatry. 1999;156:11–18. doi: 10.1176/ajp.156.1.11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Araque A, Carmignoto G, Haydon PG. Dynamic signaling between astrocytes and neurons. Annu Rev Physiol. 2001;63:795–813. doi: 10.1146/annurev.physiol.63.1.795. [DOI] [PubMed] [Google Scholar]
- 73.Sofroniew MV, Vinters HV. Astrocytes: biology and pathology. Acta Neuropathol. 2010;119:7–35. doi: 10.1007/s00401-009-0619-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Anderson CM, Swanson RA. Astrocyte glutamate transport: review of properties, regulation, and physiological functions. Glia. 2000;32:1–14. [PubMed] [Google Scholar]
- 75.Wayman WN, Dodiya HB, Persons AL, Kashanchi F, Kordower JH, Hu X-T, et al. Enduring cortical alterations after a single in vivo treatment of HIV-1 Tat. Neuroreport. 2012;23:825–829. doi: 10.1097/WNR.0b013e3283578050. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Fan Y, Zou W, Green LA, Kim BO, He JJ. Activation of Egr-1 expression in astrocytes by HIV-1 Tat: new insights into astrocyte-mediated Tat neurotoxicity. J Neuroimmune Pharmacol. 2011;6:121–129. doi: 10.1007/s11481-010-9217-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Bowers MS, Kalivas PW. Forebrain astroglial plasticity is induced following withdrawal from repeated cocaine administration. Eur J Neurosci. 2003;17:1273–1278. doi: 10.1046/j.1460-9568.2003.02537.x. [DOI] [PubMed] [Google Scholar]
- 78.Midde NM, Huang X, Gomez AM, Booze RM, Zhan CG, Zhu J. Mutation of tyrosine 470 of human dopamine transporter is critical for HIV-1 Tat-induced inhibition of dopamine transport and transporter conformational transitions. J Neuroimmune Pharmacol. 2013;8:975–987. doi: 10.1007/s11481-013-9464-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Ritz MC, Lamb RJ, Goldberg SR, Kuhar MJ. Cocaine receptors on dopamine transporters are related to self-administration. Science. 1987;237:1219–1223. doi: 10.1126/science.2820058. [DOI] [PubMed] [Google Scholar]
- 80.Midde NM, Gomez AM, Zhu J. HIV-1 Tat protein decreases dopamine transporter cell surface expression and vesicular monoamine transporter-2 function in rat striatal synaptosomes. J Neuroimmune Pharmacol. 2012;7:629–639. doi: 10.1007/s11481-012-9369-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Theodore S, Cass WA, Dwoskin LP, Maragos WF. HIV-1 protein Tat inhibits vesicular monoamine transporter-2 activity in rat striatum. Synapse. 2012;66:755–757. doi: 10.1002/syn.21564. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Paris JJ, Carey AN, Shay CF, Gomes SM, He JJ, McLaughlin JP. Effects of Conditional Central Expression of HIV-1 Tat Protein to Potentiate Cocaine-Mediated Psychostimulation and Reward Among Male Mice. Neuropsychopharmacology. 2014;39:380–388. doi: 10.1038/npp.2013.201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Dietrich JB, Mangeol A, Revel MO, Burgun C, Aunis D, Zwiller J. Acute or repeated cocaine administration generates reactive oxygen species and induces antioxidant enzyme activity in dopaminergic rat brain structures. Neuropharmacology. 2005;48:965–974. doi: 10.1016/j.neuropharm.2005.01.018. [DOI] [PubMed] [Google Scholar]
- 84.Aksenov MY, Aksenova MV, Nath A, Ray PD, Mactutus CF, Booze RM. Cocaine-mediated enhancement of Tat toxicity in rat hippocampal cell cultures: the role of oxidative stress and D1 dopamine receptor. Neurotoxicology. 2006;27:217–228. doi: 10.1016/j.neuro.2005.10.003. [DOI] [PubMed] [Google Scholar]
- 85.Brailoiu GC, Brailoiu E, Chang JK, Dun NJ. Excitatory effects of human immunodeficiency virus 1 Tat on cultured rat cerebral cortical neurons. Neuroscience. 2008;151:701–710. doi: 10.1016/j.neuroscience.2007.11.031. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Self RL, Mulholland PJ, Nath A, Harris BR, Prendergast MA. The human immunodeficiency virus type-1 transcription factor Tat produces elevations in intracellular Ca2+ that require function of an N-methyl-D-aspartate receptor polyamine-sensitive site. Brain Res. 2003;995:39–45. doi: 10.1016/j.brainres.2003.09.052. [DOI] [PubMed] [Google Scholar]
- 87.Chandra T, Maier W, Konig HG, Hirzel K, Kogel D, Schuler T, et al. Molecular interactions of the type 1 human immunodeficiency virus transregulatory protein Tat with N-methyl-d-aspartate receptor subunits. Neuroscience. 2005;134:145–153. doi: 10.1016/j.neuroscience.2005.02.049. [DOI] [PubMed] [Google Scholar]
- 88.Napier TC, Chen L, Kashanchi F, Hu XT. Repeated Cocaine Treatment Enhances HIV-1 Tat-induced Cortical Excitability via Over-activation of L-type Calcium Channels. J Neuroimmune Pharmacol. 2014;9:354–368. doi: 10.1007/s11481-014-9524-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Hegg CC, Hu S, Peterson PK, Thayer SA. Beta-chemokines and human immunodeficiency virus type-1 proteins evoke intracellular calcium increases in human microglia. Neuroscience. 2000;98:191–199. doi: 10.1016/s0306-4522(00)00101-9. [DOI] [PubMed] [Google Scholar]
- 90.Contreras X, Bennasser Y, Chazal N, Moreau M, Leclerc C, Tkaczuk J, et al. Human immunodeficiency virus type 1 Tat protein induces an intracellular calcium increase in human monocytes that requires DHP receptors: involvement in TNF-alpha production. Virology. 2005;332:316–328. doi: 10.1016/j.virol.2004.11.032. [DOI] [PubMed] [Google Scholar]
- 91.Ford KA, Wolf ME, Hu XT. Plasticity of L-type Ca2+ channels after cocaine withdrawal. Synapse. 2009;63:690–697. doi: 10.1002/syn.20651. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Peterson BZ, DeMaria CD, Yue DT. Calmodulin is the Ca2+ sensor for Ca2+-dependent inactivation of 1-type calcium channels. Neuron. 1999;22:549–558. doi: 10.1016/s0896-6273(00)80709-6. [DOI] [PubMed] [Google Scholar]
- 93.McQueen P, Donald LJ, Vo TN, Nguyen DH, Griffiths H, Shojania S, et al. Tat peptide-calmodulin binding studies and bioinformatics of HIV-1 protein-calmodulin interactions. Proteins. 2011;79:2233–2246. doi: 10.1002/prot.23048. [DOI] [PubMed] [Google Scholar]
- 94.Wayman WN, Chen L, Napier TC, Hu XT. Cocaine self-administration enhances excitatory responses of pyramidal neurons in the rat medial prefrontal cortex to HIV-1 Tat. Eur J Neurosci. 2015 doi: 10.1111/ejn.12853. In Press. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Yang CR, Seamans JK, Gorelova N. Electrophysiological and morphological properties of layers V–VI principal pyramidal cells in rat prefrontal cortex in vitro. J Neurosci. 1996;16:1904–1921. doi: 10.1523/JNEUROSCI.16-05-01904.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Markram H, Toledo-Rodriguez M, Wang Y, Gupta A, Silberberg G, Wu C. Interneurons of the neocortical inhibitory system. Nat Rev Neurosci. 2004;5:793–807. doi: 10.1038/nrn1519. [DOI] [PubMed] [Google Scholar]
- 97.Jiang M, Swann JW. A role for L-type calcium channels in the maturation of parvalbumin-containing hippocampal interneurons. Neuroscience. 2005;135:839–850. doi: 10.1016/j.neuroscience.2005.06.073. [DOI] [PubMed] [Google Scholar]
