Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Feb 16.
Published in final edited form as: Handb Exp Pharmacol. 2018;248:397–431. doi: 10.1007/164_2017_77

Cytokines in the CNS

Roberto Marisa 1,*, Patel Reesha R 1, Bajo Michal 1
PMCID: PMC7886012  NIHMSID: NIHMS1642420  PMID: 29236160

Abstract

The innate immune system plays a critical role in the ethanol-induced neuroimmune response in the brain. Ethanol initiates the innate immune response via activation of the innate immune receptors Toll-like receptors (TLRs; e.g., TLR4, TLR3, TLR7) and NOD-like receptors (inflammasome NLRs) leading to a release of a plethora of chemokines and cytokines and development of the innate immune response. Cytokines and chemokines can have pro- or anti-inflammatory properties through which they regulate the immune response. In this chapter, we will focus on key cytokines (e.g., IL-1, IL-6, TNF-α) and chemokines (e.g., MCP-1/CCL2) that mediate the ethanol-induced neuroimmune responses. In this regard, we will use IL-1β, as an example cytokine, to discuss the neuromodulatory properties of cytokines on cellular properties and synaptic transmission. We will discuss their involvement through the set of evidence: 1) changes in gene and protein expression following ethanol exposure, 2) association of gene polymorphism (humans) and alterations in gene expression (animal models) with increased alcohol intake, and 3) modulation of alcohol-related behaviors by transgenic or pharmacological manipulations of the chemokine and cytokine systems. Over the last years, our understanding of the molecular mechanisms mediating cytokine- and chemokine-dependent regulation of immune responses has advanced tremendously, and we review evidence pointing to cytokines and chemokines serving as neuromodulators and regulators of neurotransmission.

Introduction

Innate immunity is the first line of defense against an immune challenge (e.g. infection, toxin and trauma), and the response is characterized by limited specificity and a lack of memory. Regardless of the type of stimulus, the neuroimmune response involves activation of receptors of the innate immune system and release of inflammatory mediators. Inflammatory mediators comprise a heterogeneous group of factors, including cytokines, prostaglandins, free radicals, complement system, acute phase proteins, and neurotransmitters. These mediators regulate diverse aspects of the immune response including its intensity and duration. In general, the immune response/inflammation encompasses innate and adaptive immune responses that work together through direct cell contacts and through interactions involving chemical mediators (e.g. cytokines, antibodies). Contrary to the innate immunity, the adaptive immune response is very specific, develops slowly, and shows a memory (repeated challenge with the same microbe induces a faster and stronger response) [1, 2]. This chapter focuses on a group of inflammatory mediators – cytokines and chemokines, and their role in the ethanol-induced neuroimmune response and adaptive changes in the brain.

Cytokines

Cytokines are a group of more than 300 soluble glycoproteins that are produced by cells in response to immunological stimuli (microbes, toxins, tissue damage, etc.). Cytokines are characterized by pleiotropic, redundant, synergistic and antagonistic effects, and play a crucial role in regulation of the innate and adaptive immune response. The term ‘cytokine’ encompasses several classes of proteins including: interleukins, chemokines, tumor necrosis factor, interferons, and growth factors [1, 2]. Members from each of these cytokine subgroups are involved in ethanol-induced pathology in the central nervous system (CNS) [3, 4].

Interleukins (ILs) mediate signaling between cells of the immune system. ILs are produced by a variety of cells and are involved in regulation of cell growth, differentiation, and motility of immune cells [5]. Chemokines, such as MCP-1/CCL2 (Monocyte chemotactic protein 1/Chemokine ligand 2) are involved in leukocyte trafficking under both homeostatic and inflammatory conditions [6]. The tumor necrosis factor (TNF) family are characterized by their critical role in the inflammatory responses as well as in homeostatic processes [7, 8]. Interferons (IFNs) are pro-inflammatory molecules that are essential for innate and adaptive immunity and provide critical protection during early stages of viral, bacterial, or pathogen infections [9]. Growth factors include members of colony stimulating factors (CSF) which mediate development, differentiation, and expansion of cells of the myeloid series, and transforming growth factor β (TGFβ) which inhibits activation of macrophages, growth of B and T cells, and is cytotoxic [1, 2].

The diversity of cytokine molecules reflects their broad functional roles in regulation of the immune responses and homeostatic processes in the peripheral organs as well as in the CNS. During the immune response, cytokines act in concert via complex interactions to regulate gene expression, cytokine release, induction and termination of cytokine activity. While these interactions are fine-tuned under physiological conditions, their imbalance often leads to the development of pathological immune responses that are associated with numerous disorders.

Cytokines in the CNS

In the CNS, cytokines are produced locally, primarily by glial cells, but all CNS cell types are capable of synthesizing cytokines [10]. Under physiological conditions, some cytokines are produced constitutively at relatively low levels. However, cytokine levels are significantly increased after various CNS and PNS (peripheral nervous system) injuries, seizures or infections [11]. While activation of cytokine signaling in glial cells is crucial for the immune response [12, 13], cytokine signaling in neurons induces rapid and often persistent changes in excitability and/or presynaptic neurotransmitter release [11]. In addition to the local production of cytokines, cytokines are also transported across the blood-brain barrier (BBB) from the periphery via active and passive transporter systems [1416] and are produced by CNS-invading leukocytes [17]. Activation of cytokine signaling in endothelial cells of the BBB mediate the recruitment of circulating leukocytes and in some cases induces breakdown of tight junctions resulting in a leaky, permeable BBB [1823]. Additionally, peripheral cytokines can communicate with the CNS by acting on vagal afferent inputs, which modulate cholinergic signaling in the brain [24, 25].

Beyond regulating the immune response, cytokines in the CNS are also involved in regulation of homeostasis of the nervous system [10]. Cytokines play a critical role in synaptic pruning during development, synapse removal, neurogenesis, and modulation of synaptic transmission in the brain [11, 2631]. Thus, dysregulation of cytokines, for instance by ethanol exposure, has a complex impact on brain physiology and can cause long-lasting neuroadaptive changes [4]. The overall effect of cytokines on neurons and glia is dependent on several factors including interaction with other cytokines, age, sex/gender, brain region, type of stimulus, and previous history of immune challenges [3239].

Neuroimmune system and Alcohol Used Disorders

The neuroimmune system and ethanol have a complex reciprocal interaction, wherein the neuroimmune system modulates the effects of ethanol on synaptic transmission, ethanol drinking, and alcohol-related behaviors, and ethanol modulates the activity of the neuroimmune system. There are several lines of evidence supporting this bidirectional interaction. Genetic predisposition to increased ethanol/alcohol drinking is associated with polymorphisms in neuroimmune genes and altered gene expression of cytokines in humans [4042] and rodents [43, 44]. Moreover, transgenic and pharmacological manipulation of cytokine signaling alters ethanol drinking, ethanol-related behaviors, and the molecular and cellular effects of ethanol in the CNS [4453]. Reciprocally, ethanol exposure induces acute and chronic changes in brain cytokine production, making these interactions very complex. The severity and duration of the neuroimmune response represented by a particular cytokine profile vary with the type of ethanol exposure/drinking (e.g., binge consumption). Table 1 summarizes the acute and chronic ethanol-induced dysregulation of the cytokine production in the animal models and humans. Alcohol use disorder (AUD) is associated with a chronic neuroimmune response and persistently altered neuroimmune gene expression [54]. Human and animal studies suggest that key mediators of the ethanol-induced neuroimmune response and neuroadaptive changes in the CNS include interleukins IL-1β, IL-6, IL-10, chemokine MCP-1/CCL2, and TNF-α. This chapter will highlight our current understanding of the role of these cytokines in AUDs.

Table 1.

Ethanol-induced changes in the brain cytokine gene expression and protein levels in the animal models and humans.

Animal models (rodents)
Cytokine Ethanol Treatment Brain region Δ mRNA levels Δ Protein levels
IL-1β Acute Whole brain = [106, 110, 242]
↓ [242]
↑ [106]
= [110]
↓ [242]
Cortex ↑ [243]
= [35, 243245]
= [244, 246]
Hypothalamus (PVN) ↑ [242]
= [167, 242, 247]
↓ [247]
= [246]
Hippocampus ↑ [242]
= [167, 242, 247]
↓ [247]
= [246]
Cerebellum = [242]
Amygdala = [167, 247]
Prolonged/Binge Whole brain = [248] [110] = [110, 248]
Cortex ↑ [35]
= [225]
↑ [249251]
Hippocampus ↑ [225] ↑ [251]
Cerebellum ↑ [39, 225]
Striatum/NAc ↑ [249]
Chronic Cortex ↑ [244, 252]
= [244]
↓ [223]
↑ [252254]
= [244, 252]
Hypothalamus (PVN) ↓ [242]
Hippocampus = [242] ↑ [254, 255]
Cerebellum ↑ [53] ↑ [53]
Amygdala ↓ [242]
Striatum/NAc ↑ [226]
IL-6 Acute Cortex = [245] = [245]
Hypothalamus (PVN) ↑ [242, 247]
= [167, 242]
Hippocampus ↑ [242, 247]
= [167, 242, 247]
Cerebellum ↑ [242]
= [242]
↓ [242]
Amygdala ↑ [247]
= [167]
Prolonged/Binge Whole brain ↑ [248] ↑ [248]
Cortex = [221, 222] ↑ [178]
= [178]
Hypothalamus (PVN) ↑ [255]
Hippocampus = [221, 222] = [178]
Cerebellum = [222]
Chronic Cortex ↑ [253]
Hypothalamus (PVN) = [242] ↑ [256]
Hippocampus = [242] ↑ [253]
Amygdala ↑ [242]
IL-10 Acute Whole brain = [110]
Hippocampus ↑ [176]
Prolonged/Binge Whole brain = [110] ↓ [110]
Cortex = [38, 178]
Hippocampus ↑ [178]
= [38, 178]
Cerebellum = [39]
Chronic Cortex ↓ [253]
Hippocampus ↓ [253]
IL-1Ra Chronic Cerebellum ↑ [53]
TNF-α Acute Whole brain ↑ [106, 110] ↑ [106]
= [110]
Cortex ↑ [243, 245]
= [243245]
↑ [244, 257]
= [246]
Hypothalamus (PVN) ↑ [242]
= [167, 242]
↓ [242, 247]
Hippocampus ↑ [242]
= [167, 242]
↓ [242, 247]
= [246]
Cerebellum = [242]
↓ [242]
Amygdala = [167, 247]
↓ [247]
Prolonged/Binge Whole brain ↑ [110, 248]
= [110]
↑ [110, 248]
= [110]
Cortex
= [221, 222, 225]
↑ [35, 250, 251]
= [38, 178]
Hypothalamus (PVN) ↑ [255]
= [258]
Hippocampus ↑ [225]
= [221, 222]
↑ [38, 250, 251]
= [38, 178, 259]
Cerebellum ↑ [39, 225]
= [39, 222]
Striatum/NAc ↑ [35]
Chronic Cortex ↑ [223, 244, 252]
= [244]
↑ [252254]
= [244]
Hypothalamus (PVN) = [242] ↑ [256]
Hippocampus ↓ [242] ↑ [253, 254]
Cerebellum ↑ [53] ↑ [53]↓
Amygdala = [242]
Striatum/NAc ↑ [226]
= [226]
MCP-1/CCL2 Acute Whole brain ↑ [110] ↑ [260]
= [110, 260]
Cortex ↑ [245]
=[35] [222, 244, 245]
↑ [249]
= [244]
Hippocampus ↑ [222]
Cerebellum ↑[222]
Prolonged/Binge Whole brain ↑ [110, 248] ↑ [110, 248]
= [110]
Cortex ↑ [35, 221, 225]
= [221]
↑ [221]
Hippocampus ↑ [221, 225]
= [221]
↑ [221]
Cerebellum ↑ [225]
Striatum/NAc ↑ [249]
Chronic Cortex ↑ [223, 244]
= [244]

= [244]
Cerebellum ↑ [53] ↑ [53]
Striatum/NAc ↑ [226]
Humans
Cytokine Ethanol Anamnesis Brain region Δ mRNA levels Δ Protein levels
IL-1β Alcoholics –acutely exposed to EtOH CF = [219]
Alcoholics (postmortem) Hippocampus ↑ [74]
IL-6 Alcoholics + hepatic encephalopathy Superior frontal gyrus =[261]
Precentral gyrus = [261]
IL-10 Alcoholics + hepatic encephalopathy Superior frontal gyrus =[261]
Precentral gyrus = [261]
TNF-α Alcoholics –acutely exposed to EtOH CF = [219]
↓ [219]
MCP-1/CCL2 Alcoholics (postmortem) VTA ↑ [218]
Substantia nigra ↑ [218]
Hippocampus ↑ [218]
Amygdala ↑ [218]
Alcoholics –acutely exposed to EtOH CF ↑ [219]

Table 1. We define “Acute treatment” as a single administration or continuous application for less than 24 hrs. The “Prolonged/Binge treatment” includes several binge models and repeated ethanol exposure for less than 2 weeks, and “Chronic treatment” corresponds to ethanol treatments exceeding 2 weeks. The primary reasons for the discrepancies in the direction of the ethanol effects on a particular cytokine among and within the studies include age (e.g. adolescent vs adult), sex, ethanol treatment/administration (e.g. intraperitoneal vs intragastric application, or continuous vs intermittent treatment), and posttreatment time of the tissue collection (e.g. 1 day vs 28 days). The references in red correspond to the mouse studies and references in black to the studies in rats. The term “Cortex” encompasses findings from the studies on the neocortex, frontal cortex, mPFC, entorhinal and temporal cortex. We mark the direction of the ethanol effects on the cytokines (Δ mRNA/protein levels) as ↑ - increase, = - no change, and ↓ - decrease in the mRNA or protein levels. Abbreviations: CF – cerebrospinal fluid, VTA – ventral tegmental area, PVN – Paraventricular nucleus of the hypothalamus, NAc – Nucleus accumbens.

The interleukin 1 family

The interleukin 1 (IL-1) family is a group of 11 cytokines that initiate and regulate inflammatory responses [55]. IL-1α/β and its cognate IL-1 receptor type 1 (IL-1R1) are expressed throughout the brain [5667] and are synthesized in both neurons [12] and glial cells [68, 69]. Specifically, IL-1R1 is enriched in post-synaptic compartments in rat hippocampus and cortex [70, 71]. In general, IL-1α is produced constitutively, whereas IL-1β synthesis is induced and requires activation of the inflammasome pathway. The inflammasome is a multiprotein complex mainly functioning as a platform for the activation of inflammatory caspases to produce pro-inflammatory cytokines (IL-1β and IL-18) and as a trigger for the release of proteins involved in coordination of cell proliferation and tissue repair. First, an initial immune stimulus induces gene expression and protein synthesis of the inactive proIL-1β. The release of an active IL-1β requires a second stimulus that activates the inflammasome, which leads to cleavage of the proIL-1β by caspase 1 [72, 73]. Notably, activation of the inflammasome pathway, particularly NLRP3/ASC inflammasome, plays a critical role in regulation of the alcohol-induced neuroimmune response [53, 7477].

The pro-inflammatory activity of IL-1α and IL-1β are mediated by downstream signaling of IL-1R1. IL-1α/β binds to the extracellular domain of IL-1R1 leading to the recruitment of accessory proteins (e.g. the co-receptor IL-1R1 accessory protein (IL-1RAcP)), formation of a receptor heterodimeric complex (comprised of IL-1α/β, IL-1R1, and IL-1RAcP), and assemblage with the intracellular adaptor protein MyD88. Downstream of IL-1R1, many intracellular signaling pathways are activated such as: NF-κB, c-Jun N-terminal kinase, and p38 MAPK. Additionally, transcription factors, which induce gene expression of the inflammatory mediators including IL-1α/β, are also activated [78]. Importantly, IL-1α/β-IL-1R1 signaling is regulated by an endogenous antagonist (IL-1Ra) and a decoy receptor (IL-1R2). IL-1Ra (IL-1 receptor antagonist) competes with IL-1α/β for binding sites on IL-1R1, and IL-1α/β binds to the decoy receptor IL-1R2, which does not assemble into the IL-1R1/IL-1RAcp/MyD88 complex [79, 80]. In both cases, the inhibition of IL-1α/β-mediated response is carried out by preventing the activation of downstream IL-1R1 signaling [79].

Based on the available literature, ethanol does not induces changes in the IL-1α levels in the brain [53]. While this does not exclude the possibility of IL-1α playing a role in the ethanol induced neuroimmune response, IL-1α does not appear to play a critical role in ethanol effects in the CNS [45, 81]. Here, we will focus on IL-1β which play a critical role in the ethanol-induced neuroimmune response in the CNS [82].

The mechanisms by which IL-1β exerts its effects can be broadly categorized into two branches: 1) primarily immune cell mediated effects and 2) direct neuronal effects. The first branch of the IL-1β mechanisms encompasses various processes of the neuroimmune response (e.g., free radical generation, activation of glial cells) in which the IL-1β serves as a key regulator. The second branch includes the direct regulation of homeostasis in the CNS by IL-1β and IL-1β−dependent modulation of synaptic transmission. Importantly, the individual mechanisms from both branches do not act independently, but rather work in parallel, influencing the actions of each other.

The IL-1β system modulates the functional activity of neurons in a cell- and brain region-specific manner including: excitability, neurotransmitter receptors, neurotransmitter release, and synaptic plasticity. For example, IL-1β directly modulates voltage-gated ion channels [11]; it increases firing in Purkinje cells [83], decreases firing in dorsal raphe nucleus serotonergic neurons [84, 85], and has dual effects in orbitofrontal cortex neurons [86]. In the hippocampus (including hippocampal neuronal cell cultures), IL-1β increases the membrane expression of GABA (γ-Aminobutyric acid) receptors [87, 88] and IL-1R1 at synaptic sites, where IL-1R1 colocalizes and binds to the GluR2B subunit of NMDA (N-methyl D-aspartic acid) receptors [70]. IL-1β-IL-1R can increase NMDA receptor phosphorylation (e.g., GLuR2B subunit) leading to an increase in NMDA mediated calcium (Ca2+) flux, excitability and excitotoxicity [89]. The dual effects of IL-1β on presynaptic GABA and glutamate release as well as postsynaptic inhibitory and excitatory activity are specific to neuronal type and brain region [81, 9097]. Moreover, IL-1β inhibits synaptic plasticity in CA1 and dentate gyrus neurons of the hippocampus [97101].

IL-1β and Alcohol

There are several lines of evidence supporting the critical role of IL-1β in the neuropathogenesis and behavioral changes associated with alcohol dependence. In humans, polymorphisms in Il1rn and Il1b, the genes encoding IL-1Ra and IL-1β, respectively, are associated with a susceptibility to alcoholism in Spanish men [40]. Similarly, mice with a genetic predisposition to high alcohol consumption show altered expression of several genes of the IL-1/IL-1R system, including Ilf5, Ilf6, Ilf8, Irak4, and Il1rn. All of these genes, except Irak4, are also located within QTLs (quantitative trait locus) for human alcoholism susceptibility and are considered as candidate genes for alcohol drinking [43]. ILf5, ILf6 and ILf8 are ligands for IL-1R2 [102]. Irak4 encodes the protein IRAK4 (IL-1 receptor-associated kinase 4), which plays a key role in the activation of NF-κB signaling [103]. Interestingly, high alcohol preferring (HAP) mice have altered levels of several genes involved in the NF-κB pathway (Casp8, Fadd, Ikbkb, Ikbkg, Map3k1, Map3k7, Tradd), through which IL-1α/β -IL-1R1 mediate its biological action [43]. Follow-up behavioral studies show the involvement of some of these genes in alcohol drinking and preference. Il1rn encodes the IL-1Ra protein that is an endogenous competitive antagonist of IL-1R1. Il1rn knockout mice exhibit a reduction in alcohol drinking and preference [49], increased ethanol clearance and decreased ethanol-induced conditioned taste aversion, increased sensitivity to the sedative/hypnotic effects of ethanol and flurazepam, and reduced severity of acute ethanol withdrawal. Pretreatment with exogenous IL-1Ra (Kineret) reverses some of the behavioral phenotypes of Il1rn KO mice, specifically it reduces the ethanol- and flurazepam-induced sedation and restores the severity of acute ethanol withdrawal [47]. Mice lacking the Il1r1 gene, encoding IL-1R1, exhibit the opposite phenotype of Il1rn KO mice – decreased ethanol-induced sedation and increased severity of ethanol withdrawal, indicating that IL-1R1 signaling plays a crucial role in these behaviors. However, the findings that ethanol intake and preference are not altered in Il1r1 KO mice and recovery from ethanol-induced motor incoordination is only altered in female Il1r1 KO mice suggests that these alcohol-related behaviors are not solely regulated by the IL-1β system [47]. Moreover, systemic administration of IL-1Ra reduces alcohol-induced sedation and motor impairment recovery time in mice [51] and also prevents alcohol-induced neuroinflammation [53].

Pharmacological manipulation of the IL-1 system selectively in the CNS provides further evidence for a critical role of the brain IL-1 system in several alcohol-related behaviors. Intracerebroventricular administration of IL-1β increases alcohol withdrawal-induced anxiety [104], while bilateral infusion of IL-1Ra into the basolateral amygdala (BLA), but not the central nucleus of the amygdala (CeA), reduces ethanol consumption with no impact on either sucrose drinking or open-field locomotor activity, a behavioral measure of anxiety [105]. Overall, these evidences indicate that IL-1β plays a critical role in activation of the ethanol-induced immune response in the brain and is involved in the regulation of critical neurocircuitries mediating the alcohol-related behaviors.

IL-1β mechanisms of action

Evidence for the involvement of IL-1β and its signaling pathways in alcohol-related behaviors are compelling. Indeed, ethanol increases IL-1β levels in neuronal and glial cell cultures [74, 106108], and in specific brain regions in animal models of AUDs as well as in humans (see Table 1). In this regard, the hippocampus, PFC, and cerebellum seems to be the most sensitive to ethanol-induced dysregulation of IL-1 - IL-1R1 signaling [53, 109111]. However, the mechanisms through which IL-1β modulates alcohol-related behaviors are still not fully understood. Therefore, the focus of current research has extended to the other brain regions such as the amygdala, which plays a critical role in alcohol dependence and withdrawal [112]. Thus, here, we will summarize our current understanding of the mechanisms of action of the IL-1β and the IL-1β-ethanol interactions at the cellular and behavioral levels in the CeA, BLA and hippocampus.

IL-1β in the CeA

The CeA, a major component of the extended amygdala, is a primarily GABAergic nucleus involved in stress-, fear- and anxiety-like behavior [113] and excessive drinking [112, 114]. The GABAergic system tightly controls neuronal excitability [115, 116], and it is critical in the development of alcohol dependence [117, 118].

Modulation of GABA-A receptors alters many ethanol behaviors [119121]. Specifically, muscimol, a GABA-A receptor agonist, injection into the CeA greatly reduces ethanol self-administration, but only in dependent rats [122], and a GABA-A antagonist reduces ethanol self-administration [123] in non-dependent rats. The CeA has abundant corticotrophin releasing factor (CRF)-containing fibers and CRF receptors [124], and is thought to be a target of the peripheral neuroimmune system [125]. CRF1 receptors play an essential role in ethanol’s effects on GABA release in the CeA and in ethanol dependence [114, 126128]. Interestingly, facilitation of ethanol withdrawal-induced anxiety by TNFα or MCP-1/CCL2 microinjection into the CeA is dependent on CRF [129], and CRF-amplified neuronal TLR4/MCP-1 signaling in the CeA regulates alcohol self-administration [44]. Moreover, IL-1β and IL-1Ra regulate GABAergic transmission in the CeA [45, 81]. Under basal conditions IL-1R is detected in the amygdala [130], but expression of IL-1β and IL-1Ra are not detectable but rather appears to be inducible in the CeA [125, 131], suggesting that modulation of basal GABAergic transmission with acute application IL-1Ra is likely through IL-1α. Systemic IL-1β and LPS administration activates the CeA, as indicated by an increase in gene expression of the immediate early gene product cFos [125, 130, 132]. At the cellular level, IL-1β significantly decreases amplitudes of evoked inhibitory GABA-A mediated postsynaptic potentials (eIPSP), without affecting paired-pulse facilitation (PPF), a paradigm to assess pre- and postsynaptic mechanisms for evoked responses. Interestingly, IL-1β has dual effects on action potential independent miniature inhibitory postsynaptic currents (mIPSCs) in CeA neurons: in the majority of the cells, IL-1β increases mIPSC frequency suggesting an increase in presynaptic vesicular GABA release. However, in some CeA neurons, IL-1β decreases vesicular GABA release as well as postsynaptic GABA-A receptor function represented by a decrease in mIPSC amplitude. Consistent with the IL-1β effects, IL-1Ra alone had dual effects on mIPSCs, and it also blocks the effects of IL-1β on CeA GABA transmission [45, 81].

Further, acute ethanol facilitates GABA transmission in the CeA [128]. IL-1β interacts with the effects of acute ethanol on GABA transmission in the CeA. Although IL-1β pretreatment does not block the ethanol-induced facilitation of evoked responses, IL-1β occludes ethanol’s effects on presynaptic vesicular GABA release in CeA neurons responding to IL-1β. Overall, these findings indicate that the IL-1 system is involved in tonic regulation of GABA transmission and that IL-1β interacts with the ethanol-induced enhancement of the GABAergic transmission in the CeA [81].

The endogenous IL-1Ra is an anti-inflammatory element that may play a critical role in the development of alcohol dependence [43, 49, 51, 53]. Transgenic mice lacking endogenous IL-1Ra (Il1rn KO) exhibit reduced alcohol intake [49], prolonged loss of the righting reflex (LORR) induced by ethanol or by the GABA-A receptor positive allosteric modulator flurazepam [47]. Also, GABAergic neurotransmission in the CeA of Il1rn KO mice is disrupted. Notably, both baseline evoked GABA responses and baseline frequency of action potential dependent spontaneous inhibitory postsynaptic currents (sIPSCs), but not mIPSCs, are significantly increased in these KO mice compared to wildtype (WT) mice, indicating increased GABA release in the CeA of KO mice. Acute application of ethanol increases the frequency of sIPSCs and mIPSCs in a vast majority of the WT CeA neurons, but these effects are observed only in about half of the KO CeA neurons. Pretreatment with exogenous IL-1Ra (Kineret) reverses this increase in KO mice without altering the frequency in WT mice. Kineret is also capable of restoring the ethanol-induced increase in GABA release in KO mice, indicating that some of the cellular phenotypes in Il1rn KO mice are rescued by application of exogenous IL-1Ra [45]. This suggests that IL-1R1 antagonism regulates basal GABA release and plays a key role in the effects of ethanol at inhibitory synapses in the CeA

IL-1β in the BLA

Acute application of IL-1β hyperpolarizes the membrane and decreases input resistance in most BLA neurons. The hyperpolarization induced by IL-1β is dose-dependent, reversible, action potential independent, and blocked with a GABA-A antagonist. IL-1β inhibits excitatory and inhibitory responses evoked by stimulating either the bed nucleus of stria terminalis or the lateral amygdala via presynaptic mechanisms. Thus, IL-1β hyperpolarizes the membrane through indirect mechanisms, possibly by enhancing the action of endogenous GABA in the BLA and inhibits excitatory and inhibitory transmission at presynaptic sites [133].

Binge-like ethanol drinking induces a significant increase in IL-1β mRNA and protein expression within the amygdala, but not CeA. Interestingly, bilateral infusions of IL-1Ra into the BLA, but not the CeA, reduces ethanol drinking without affecting sucrose drinking or open-field locomotor activity [105]. These results highlight a specific role for IL-1 receptor signaling in the BLA in modulating binge-like ethanol consumption and indicate that pro-inflammatory cytokines can be induced prior to progression into alcohol dependence.

IL-1β in the hippocampus

The hippocampus expresses a high density of IL-1β receptors presumably on dendrites of granule cells [134]. Exogenously applied IL-1β enhances neuronal excitability and increases NMDA receptor function. Indeed, data from primary rat hippocampal neuron cultures suggests that IL-1β increases NMDA receptor function through activation of tyrosine kinases and subsequent NR2A/B subunit phosphorylation [89]. IL-1β reduces seizure thresholds and inhibition of IL-1R1 by its antagonist limits seizures [135]. Moreover, convulsant and/or excitotoxic stimuli increases the production of IL-1β in microglia-like cells in the hippocampus [135]. A later study, clarified the mechanism of IL-1β-associated seizures and the interaction between IL-1β and Ca2+ mobilization on glutamate and GABA releases using mice hippocampal mini-slice [136]. Both basal and K+-evoked GABA releases are regulated by Ca2+ influx and Ca2+-induced Ca2+ releasing system (CICR). Similarly, the K+-evoked glutamate release is also regulated by Ca2+ influx and CICR, but basal glutamate release is not. IL-1β increased basal releases of glutamate and GABA depending on the activation of Ca2+ influx and ryanodine receptor (RyR)-sensitive CICR. During neuronal hyperexcitability, the effect of IL-1β on GABA release is more predominantly modulated by Ca2+ influx and RyR-sensitive CICR than that on glutamate [136].

IL-1β can also impact neuronal plasticity. Low, physiological levels of IL-1β play a role in long-term potentiation (LTP), an important cellular correlate of learning and memory, while high, pathological levels can disrupt this process. Blockade of IL-1β signaling by its antagonist, IL-1Ra, impairs memory. However, addition of excessive IL-1β also impairs memory [137]. Therefore, immune signaling impacts plasticity through finely tuned changes in cytokine levels that alter neuronal activity, neural circuitry and consequently behavioral phenotypes [4].

IL-1β can also affect neurogenesis, the process of generating functional neurons from neural precursors, in the hippocampus. Inflammation [138] and chronic stress [139] reduce neurogenesis and cause depression-like behavior. In particular, stress induces IL-1β expression in the hippocampus, which decreases neurogenesis and contributes to depression. Blockade of IL-1β signaling inhibits stress-induced decreases in neurogenesis and depression-like behavior [140]. The increased gene expression and protein levels of IL-1β in the hippocampus following prolonged/binge and chronic ethanol exposure, found in animal models as well as in human alcoholics (Table 1), indicate that IL-1β may play a critical role in the ethanol-reduced hippocampal neurogenesis [141]. Indeed, blocking IL-1β or inflammasome signaling reverses the ethanol effects on the neurogenesis [141]. These findings indicate that inflammasome and IL-1β mediate the ethanol-induced inhibition of the hippocampal neurogenesis [74].

Thus, it is clear that the IL-1β system plays a neuromodulatory role and interacts with ethanol in CeA/BLA/hippocampus neurons. At the same time, there are still many unanswered questions regarding the mechanisms mediating brain region differences in the IL-1β effects and neuroadaptive changes of the IL-1β system induced by chronic ethanol exposure and withdrawal.

Interleukin-6

Interleukin-6 (IL-6) is a prototypical pro-inflammatory cytokine involved in the transition from innate to adaptive immunity. IL-6 plays a major role in the neuroimmune response to brain injury and is associated with multiple neurobiological (e.g., multiple sclerosis, Parkinson’s disease, Alzheimer’s disease) and psychiatric (major depression, post-traumatic stress disorder, substance use disorders) disorders [142]. In addition to mediating the neuroimmune response, IL-6 is critical in neurogenesis and the regulation of various physiological processes (e.g., food intake, body weight, body temperature, stress, sleep-awake behavior, etc.) [143148]. Neurons, astrocytes, microglia and endothelial cells are the essential sources of IL-6, but astrocytes are the primary source of the IL-6 under physiological conditions and during alcohol exposure in the CNS [149152]. Production of IL-6 in brain cells is regulated by other cytokines and inflammatory factors (e.g., IL-1β and TNF-α) as well as by neurotransmitters and neuropeptides (e.g., norepinephrine, serotonin, substance P) [142, 153156]. IL-6 signaling is initiated by binding of IL-6 to the IL-6 receptor (IL-6R) and recruitment of additional accessory proteins including gp130, which leads to the activation of major signaling pathways including JAK2/ STAT3, p44/42 MAPK, and PI3-K [157]. IL-6 modulates gene expression of many inflammatory mediators and proteins involved in apoptosis and other processes [142]. At the cellular level, IL-6 has an inhibitory effect on sodium (Na+) and Ca2+ voltage-gated ion channels that may serve as a neuroprotective mechanism in the CNS [11, 158]. Moreover, IL-6 modulates glutamate receptor (mGluR2/3) expression and glutamate-mediated excitotoxicity [159161]. IL-6 also reversibly decreases GABA-A mediated currents, likely via modulation of GABA-A receptor compartmentalization and PI3-K-Akt pathway [162]. The direct effects of IL-6 on cellular physiology and synaptic transmission indicate that a dysregulation of the IL-6 signaling may lead to a significant disturbance in network activity in a brain region-specific manner.

IL-6 and Alcohol.

A polymorphism in the Il6, the gene encoding IL-6, is associated with alcoholism in humans [41, 42], and genomic studies show modifications in Il6 gene expression in alcohol-preferring rodents [43]. Transgenic mice with a null mutant Il6 have lower ethanol intake and ethanol preference compared to WT mice [49]. On the other hand, transgenic mice with elevated astrocyte production of IL-6 in the CNS (IL-6tg mice) show increased susceptibility to acute alcohol withdrawal hyperexcitability [163].

While ethanol has mixed effects on the IL-6 levels in neuronal and glial cell cultures [107, 108, 164166], both acute and chronic ethanol exposures primarily increase IL-6 levels in a brain-region and ethanol-exposure (time and dose) specific manner (see Table 1.). In addition to the direct effects of ethanol on IL-6 levels in the brain, ethanol’s effects on IL-6 levels might also be under conditioned control. Repeated pairings between distinctive odor cues (conditional stimulus) and ethanol can result in elevation of IL-6 levels in the hippocampus and amygdala upon presentation of the odor cues alone [167]. At the synaptic level, IL-6tg mice exhibit an altered response in hippocampal LTP to acute ethanol. While acute ethanol depresses fEPSPs (field excitatory postsynaptic potentials), PTP (post-tetanic potentiation) and LTP, and does not affect sPS (secondary population spikes) in WT (non-tg) mice, acute ethanol increases fEPSPs and sPS and does not affect the PTP and LTP in Il-6tg mice [163, 168]. These studies on IL-6tg mice suggest possible mechanisms mediating IL-6 and ethanol interactions, particularly following the ethanol-induced increase in IL-6 levels in the brain.

Interleukin 10

Interleukin-10 (IL-10) is an anti-inflammatory cytokine that negatively regulates inflammation. IL-10 is expressed in the brain; specifically, it is produced by microglia, astrocytes, and neurons [169]. IL-10 binds to its cognate cell surface receptor, a heterotetrameric complex consisting of two ligand binding IL-10 receptor 1 (IL-10R1) chains and two accessory IL-10 receptor 2 (IL-10R2) chains [169, 170] also expressed in glia and neurons. This interaction leads to the activation of downstream signaling cascades including the JAK/STAT3 and PI3-K/AKT pathways [169172] and ultimately results in diverse biological effects in the brain such as: limiting the synthesis of pro-inflammatory mediators and reducing cytokine receptor expression and activation [173], neuroprotection [171, 174], and modulation of synaptic structure and activity [175, 176]. At the cellular level, IL-10 regulates GABAergic transmission in the hippocampal (dentate gyrus) neurons via both pre- and postsynaptic mechanisms; IL-10 decreases mIPSCs and tonic GABA currents, and its postsynaptic mechanisms of actions are mediated by PI3K pathways [176].

IL-10 and Alcohol.

IL-10 is implicated in alcoholism in humans. Human genetic studies show that a −592C>A polymorphism in the IL-10 gene is associated with alcoholism [41]. Further, IL-10R2 levels are decreased in the CeA and cortex of alcoholics [177]. Notably, IL-10 regulates SOCS (suppressor of cytokine signaling), and SOCS mRNA levels are also decreased in the CeA and cortex of alcoholics [177].

IL-10 expression and signaling are altered in several CNS pathologies [169]. Expression studies show that a single intoxicating dose of ethanol increases IL-10 content in rat hippocampus and primary cultured cortical neurons [176], 24-hour ethanol exposure increases IL-10 production by human monocytes [172], 4-day binge ethanol exposure results in protracted increases in IL-10 levels in the rat hippocampus [178], and 12-day withdrawal after chronic ethanol exposure increases IL-10 content in the rat hippocampus, prefrontal cortex, and brainstem [179]. In contrast, 4-day binge drinking in the dark paradigm decreases IL-10 levels in the mouse BLA, but not in the CeA, and IL-10 infusion into the BLA, but not the CeA, decreases binge-like drinking [52]. A 10-day binge ethanol exposure decreases mouse brain IL-10 levels [110]. The differential effects on IL-10 expression are likely due to differences between species, animal models, and examination of region-specific versus whole brain changes. Despite the growing body of evidence on an important role of IL-10 in regulation of the alcohol-related behavior, particularly binge drinking, the mechanistic and functional aspects of IL-10 and ethanol interactions are very limited.

Tumor Necrosis Factor-alpha

Tumor necrosis factor-alpha (TNF-α) is a member of the TNF superfamily and is central to the innate immune response and maintenance of homeostasis at the cellular, tissue and organism levels. In the CNS, TNF-α displays pro-inflammatory effects and is considered a major mediator of the secondary CNS damage following acute injury and during chronic inflammation. However, TNF-α also exerts essential beneficial functions in the CNS. Its potent pro-inflammatory effects require very tight temporal and spatial control, as dysregulation of TNF-α production and activity can trigger cell death and tissue degeneration [8]. TNF-α is produced in two bioactive forms transmembrane (tmTNF) and soluble (solTNF), that differ in their biological activity and intracellular signaling [180]. In general, solTNF has systemic inflammatory effects and is necessary for optimization of the immune response, whereas tmTNF mediates a subset of beneficial TNF-α activities and only basic immune responses [8]. The TNF-α system has two receptors – TNFR1 and TNFR2. While both TNF-α forms bind to TNFR1, tmTNF is the sole ligand for TNFR2 [181]. TNFR1 is ubiquitously and constitutively expressed, and its activation induces pro-inflammatory signaling through the NF-kB and AP1 transcription factors [182]. TNFR2 expression is restricted to immune cells, endothelial cells, and CNS cells – including neurons, astrocytes, and oligodendrocytes. Activation of TNFR2 leads mainly to induction of pro-survival signals mediated by AKT and NF-kB signaling pathways [183, 184]. Under physiological conditions, TNF-α plays an important role in the regulation of homeostatic processes such as synaptic scaling and plasticity [185187] and regulation of inhibitory and excitatory neurotransmission [11]. Under pathological conditions in the CNS, TNF-α has both protective and pro-inflammatory effects [8]. For example, a mechanism by which TNF-α mediates neurotoxicity is by inhibiting glutamate uptake causing glutamate excitotoxicity [188].

Similar to IL-1β and IL-6, TNF-α has neuromodulatory effects in the CNS. TNF-α enhances Na+ channels and has mixed effects on voltage-gated Ca2+ channels (solTNF-α decreasing and mTNF-α increasing Ca2+ currents) [11]. Presynaptically, TNF-α increases action potential dependent spontaneous excitatory postsynaptic currents (sEPSCs) in corticostriatal projections, through AMPA receptors [189]. Also, TNF-α can modify extracellular glutamate levels indirectly by inducing glutamate release from microglia [190] and astrocytes [191] and by inhibiting glutamate uptake by astrocytes [188]. TNF-α-TNF-R1 signaling preferentially affects AMPARs in a brain region-specific manner. TNF-α facilitates AMPAR-mediated glutamatergic transmission and enhances neuronal excitability in the hippocampus, cortex, amygdala and spinal cord [192194]. Notably, activation of the CB1 cannabinoid receptor reverses TNF-α effects on AMPAR [195]. In the striatum, however, TNF-α induces the internalization of GluR1-GluR2 AMPAR subunits leading to a decrease in the excitatory drive on inhibitory GABA neurons. Also, TNF-α promotes the endocytosis of GABA-A receptors (subunits b2/3) [192]. Thus, TNF-α effects on glutamate and GABA receptors lead to enhanced neuronal excitability and in some instances to excitotoxicity [185, 192, 196, 197]. In the CeA, TNF-α increases the amplitude of mEPSCs via the PI3-K signaling pathway but does not affect mEPSC frequencies, suggesting a predominantly postsynaptic mechanism of action. Further, TNF-α increases the firing of CeA neurons through its action on glutamate receptors [129, 198]. Finally, TNF-α increases the frequency of mIPSC, indicating an increase in presynaptic GABA release, and this effect is blocked by a CRF1 antagonist and minocycline, which is an inhibitor of glial activation [198]. These findings indicate that TNF-α interacts and modulates key neurotransmitters (GABA and glutamate) and neuropeptide (CRF) systems involved in alcohol-related behaviors [199].

TNF-α and Alcohol.

Elevated plasma levels of TNF-α in alcoholics is associated with increased craving and relapse to drinking [200]. In contrast to IL-1β and IL-6 cytokines, genomic studies in rodents did not find alterations in TNF-α gene expression in alcohol-preferring mice [43]. In general, TNF-α levels are predominantly decreased following acute ethanol treatment and increased after chronic ethanol exposure. Intracerebroventricular (i.c.v.) and intra-CeA administration of TNF-α before a single chronic ethanol exposure and ethanol withdrawal sensitizes ethanol withdrawal-induced anxiety-like behavior [104, 129], and this effect is mediated by CRF1, as a CRF1 antagonist reduces the TNF-α induced elevation of withdrawal-induced anxiety [129]. The interactions of TNF-α and the CRF system particularly in the CeA, where CRF1 is known to mediate ethanol effects on GABAergic transmission, may represent one of the mechanisms involved in TNF-α -induced modulation of the synaptic transmission. However, the mechanisms and functional consequences of TNF-α and its interaction with ethanol on neuromodulation are not known.

Chemokine ligand 2

The chemokine ligand 2 (CCL2), also known as monocyte chemotactic protein 1 (MCP-1), is a member of the monocyte chemo-attractant protein (MCP) family. CCL2 is a vital chemokine that controls the migration and infiltration of monocytes/macrophages [201]. In the brain, CCL2 is produced mainly by astrocytes and microglia and to a lesser extent by endothelial cells [202]. The neuronal expression of CCL2 is present in several brain regions including the cortex, hippocampus, hypothalamus, substantia nigra, and cerebellum [203]. Importantly, the expression of CCL2 colocalizes with classical neurotransmitters, particularly acetylcholine and dopamine [203], and cell depolarization can induce Ca2+-dependent CCL2 release [204, 205]. Compared to CCL2, its receptor CCR2 is expressed by resident immune cells, such as microglia [206, 207]. Moreover, CCR2 production is also found in cultured rat microglia [208], human fetal astrocytes [209], and in neurons of the adult rat brain [210], mainly from the cortex, hippocampus, hypothalamus, amygdala, substantia nigra, ventral tegmental area (VTA) and cerebellum [211]. There are two splice variants of CCR2 - CCR2A and CCR2B. The splice variants are expressed in different immune cells, and they activate different signaling pathways and exert distinct actions. CCR2, as a Gαi class G-protein coupled receptor [212], signals through inhibition of adenylate cyclase, and PI3-K, MAPK and protein kinase C pathways [213215]. CCL2 enhances neuronal excitability and excitatory synaptic transmission in CA1 hippocampal neurons via presynaptic mechanisms [216]. Importantly, CCR2 has both pro-inflammatory and anti-inflammatory actions [201], and CCL2-CCR2 are involved in some physiological processes and the pathogenesis of neurodegenerative disorders and AUD.

CCL2 and Alcohol

CCL2 levels are elevated in several brain regions (e.g. hippocampus and cortex) in postmortem tissue of human alcoholics [217, 218] and the cerebrospinal fluid (CF) of alcohol-dependent human subjects [219]. Similarly, ethanol exposure and withdrawal increases levels of CCL2 in several brain regions [36, 110, 129, 220226]. Indeed, alterations in the CCL2 system affect ethanol intake and motivation as mice deficient in Ccl2 or Cclr2, the genes encoding CCL2 and its receptor CCLR2, drink less ethanol and show reduced ethanol-induced aversion [50]. There is no significant difference in ethanol intake between ethanol non-dependent CCL2-tg (mice overexpressing CCL2 in astrocytes) and their control WT (non-tg) mice, whereas ethanol-dependent CCL2-tg mice drink less than the dependent non-tg mice [227]. Notably, chronic infusions of CCL2 results in long-lasting heightened ethanol intake in rats suggesting that persistent exposure to CCL2 may be required for CCL2/alcohol interactions [228]. Interestingly, CCL2-tg mice did not show acute alcohol-induced impairments in contextual learning that are observed in non-tg mice[168]. However, ethanol induced a spatial learning impairment in non-dependent CCL2-tg mice but not in non-dependent non-tg mice. Overexpression of CCL2 has a protective effect against alcohol-induced impairments in associative learning [227]. Like IL-1β and TNF-a, intracerebral injection of CCL2 before ethanol exposure and withdrawal elevates ethanol withdrawal-induced anxiety-like behavior [104]. At the cellular level, CCL2-tg mice are resistant to the depressing effects of acute alcohol (20–60 mM) on hippocampal LTP in non-tg mice. CCL2 can enhance neuronal excitability and excitatory synaptic transmission in CA1 hippocampal neurons via presynaptic mechanisms [168]. These studies on transgenic animals targeting CCL2 have significantly advanced our understanding of the potential role of CCL2 in the neuropathology of AUD.

Mechanistically, CCL2/CCR2 system involvement in the neurobiology of AUD includes interactions with other neurotransmitter and neuropeptide systems, particularly CRF and the orexigenic peptide MCH (melanin-concentrating hormone). CCL2 is expressed in cholinergic and dopaminergic neurons [203], and it modulates neuronal activity and synaptic transmission [229, 230]. CCL2 levels in the CeA and VTA are increased in alcohol-preferring P rats compared to non-preferring NP rats, and CCL2 in these brain regions, but not in ventral pallidum, mediates binge drinking in P rats. Importantly, CRF mediates feedback regulation of TLR4 (toll-like receptor 4) and CCL2 signaling in the CeA and VTA during ethanol consumption [44] suggesting that CRF, TLR4 and CCL2 in these regions regulate the initiation of excessive drinking [44]. Moreover, prenatal exposure to ethanol increases later adolescent ethanol drinking which is associated with increased CCR2 levels and increased density of neurons co-expressing CCR2 and MCH in the lateral hypothalamus [224]. As both CCR2 and MCH are believed to promote ethanol intake, these findings suggest that these systems may work together to promote ethanol drinking. Although our understanding of mechanisms mediating CCL2’s contribution to the AUD has advanced, there is still a lot of unknowns regarding CCL2 regulation of the synaptic transmission in other alcohol-related brain regions.

Conclusion

The role of the neuroimmune system and cytokines in the neurobiology of AUDs is supported by several lines of evidence. Ethanol-induced cytokine responses in the CNS are dynamic and depend on multiple factors including the duration and amount of ethanol exposure, sex, brain region, cellular specificity, and history of previous immune challenges (e.g. infection, trauma, stress, etc). Cytokines contribute to the neuroadaptive changes in the CNS induced by ethanol exposure through their direct and indirect effects on all CNS cell types, which lead to the modulation of neuronal activity, glia cells, neurogenesis, and potentially neurodegeneration.

Although our understanding of the role of key cytokines in the ethanol-induced immune response has advanced, there are still many unanswered questions especially regarding the therapeutic implications of targeting cytokines and their downstream signaling pathways. The critical role of the neuroimmune system in the neuropathology of AUD suggests its potential to be targeted for the development of new treatments for AUDs. Currently, the focus of preclinical research is on inhibiting the alcohol-induced neuroimmune response and associated alcohol-related behaviors, particularly alcohol drinking. The strategies involve targeting individual components of the neuroimmune system (e.g. TLR4 [231], IL-1R1 [51], or IL-10 [52]) or to use drugs that simultaneously target several inflammatory pathways as well as other brain signaling systems (e.g. peroxisome proliferator-activated receptor agonists (fenofibrate, pioglitazone, tesaglitazar, bezafibrate) [232234], phosphodiesterase inhibitors (e.g. ibudilast, rolipram) [235, 236], and naloxone/naltrexone [237, 238]. Regarding cytokines, preclinical studies suggest that activation or increased expression of anti-inflammatory cytokines such as IL-1Ra and IL-10 might have therapeutic value. There are, however, several challenges in targeting the neuroimmune system for the development of therapeutic strategies for alcoholism: 1) different inflammatory pathways seem to be critical for different stages of alcohol addiction and alcohol-related behaviors [239], 2) the peripheral immune system is compromised in human alcoholics [240], and 3) the neuroimmune response has both neurotoxic as well as neuroprotective roles and thus, strategies based solely on blocking the neuroimmune system may be counterproductive [241]. Understanding the role and mechanisms of action of individual components of the neuroimmune systems in the development and maintenance of alcohol addiction and relapse will be crucial for identification of new, more target specific and efficacious therapies for AUD.

Acknowledgements

This publication was supported by National Institutes of Health grants U01 AA013498, AA015566, AA006420, AA017447, T32 AA007456 and R01 AA021491 from the National Institute on Alcohol Abuse and Alcoholism, and the Pearson Center for Alcoholism and Addiction Research. The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.

References

  • 1.Abbas AK, et al. , Cellular and molecular immunology. 2018.
  • 2.Lydyard PM, et al. , Immunology. 2011, New York: Garland Science. [Google Scholar]
  • 3.Montesinos J, Alfonso-Loeches S, and Guerri C, Impact of the Innate Immune Response in the Actions of Ethanol on the Central Nervous System. Alcohol Clin Exp Res, 2016. 40(11): p. 2260–2270. [DOI] [PubMed] [Google Scholar]
  • 4.Crews FT, et al. , The role of neuroimmune signaling in alcoholism. Neuropharmacology, 2017. 122: p. 56–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Vosshenrich CA and Di Santo JP, Interleukin signaling. Curr Biol, 2002. 12(22): p. R760–3. [DOI] [PubMed] [Google Scholar]
  • 6.Bachelerie F, et al. , International Union of Basic and Clinical Pharmacology. [corrected]. LXXXIX. Update on the extended family of chemokine receptors and introducing a new nomenclature for atypical chemokine receptors. Pharmacol Rev, 2014. 66(1): p. 1–79. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Sedger LM and McDermott MF, TNF and TNF-receptors: From mediators of cell death and inflammation to therapeutic giants - past, present and future. Cytokine Growth Factor Rev, 2014. 25(4): p. 453–72. [DOI] [PubMed] [Google Scholar]
  • 8.Probert L, TNF and its receptors in the CNS: The essential, the desirable and the deleterious effects. Neuroscience, 2015. 302: p. 2–22. [DOI] [PubMed] [Google Scholar]
  • 9.Nallar SC and Kalvakolanu DV, Interferons, signal transduction pathways, and the central nervous system. J Interferon Cytokine Res, 2014. 34(8): p. 559–76. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Becher B, Spath S, and Goverman J, Cytokine networks in neuroinflammation. Nat Rev Immunol, 2017. 17(1): p. 49–59. [DOI] [PubMed] [Google Scholar]
  • 11.Vezzani A and Viviani B, Neuromodulatory properties of inflammatory cytokines and their impact on neuronal excitability. Neuropharmacology, 2015. 96(Pt A): p. 70–82. [DOI] [PubMed] [Google Scholar]
  • 12.Allan SM, Tyrrell PJ, and Rothwell NJ, Interleukin-1 and neuronal injury. Nat Rev Immunol, 2005. 5(8): p. 629–40. [DOI] [PubMed] [Google Scholar]
  • 13.Vezzani A, et al. , The role of inflammation in epilepsy. Nat Rev Neurol, 2011. 7(1): p. 31–40. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Erickson MA, Dohi K, and Banks WA, Neuroinflammation: a common pathway in CNS diseases as mediated at the blood-brain barrier. Neuroimmunomodulation, 2012. 19(2): p. 121–30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Erickson MA and Banks WA, Cytokine and chemokine responses in serum and brain after single and repeated injections of lipopolysaccharide: multiplex quantification with path analysis. Brain Behav Immun, 2011. 25(8): p. 1637–48. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Banks WA, The blood-brain barrier in neuroimmunology: Tales of separation and assimilation. Brain Behav Immun, 2015. 44: p. 1–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Callahan MK and Ransohoff RM, Analysis of leukocyte extravasation across the blood-brain barrier: conceptual and technical aspects. Curr Allergy Asthma Rep, 2004. 4(1): p. 65–73. [DOI] [PubMed] [Google Scholar]
  • 18.Annunziata P, et al. , Substance P antagonist blocks leakage and reduces activation of cytokine-stimulated rat brain endothelium. J Neuroimmunol, 2002. 131(1–2): p. 41–9. [DOI] [PubMed] [Google Scholar]
  • 19.Tsao N, et al. , Tumour necrosis factor-alpha causes an increase in blood-brain barrier permeability during sepsis. J Med Microbiol, 2001. 50(9): p. 812–21. [DOI] [PubMed] [Google Scholar]
  • 20.Pan W and Kastin AJ, Upregulation of the transport system for TNFalpha at the blood-brain barrier. Arch Physiol Biochem, 2001. 109(4): p. 350–3. [DOI] [PubMed] [Google Scholar]
  • 21.Librizzi L, et al. , Seizure-induced brain-borne inflammation sustains seizure recurrence and blood-brain barrier damage. Ann Neurol, 2012. 72(1): p. 82–90. [DOI] [PubMed] [Google Scholar]
  • 22.Vezzani A and Friedman A, Brain inflammation as a biomarker in epilepsy. Biomark Med, 2011. 5(5): p. 607–14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Rochfort KD and Cummins PM, The blood-brain barrier endothelium: a target for pro-inflammatory cytokines. Biochem Soc Trans, 2015. 43(4): p. 702–6. [DOI] [PubMed] [Google Scholar]
  • 24.Hosoi T, Okuma Y, and Nomura Y, The mechanisms of immune-to-brain communication in inflammation as a drug target. Curr Drug Targets Inflamm Allergy, 2002. 1(3): p. 257–62. [DOI] [PubMed] [Google Scholar]
  • 25.Maier SF, et al. , The role of the vagus nerve in cytokine-to-brain communication. Ann N Y Acad Sci, 1998. 840: p. 289–300. [DOI] [PubMed] [Google Scholar]
  • 26.Boulanger LM, Immune proteins in brain development and synaptic plasticity. Neuron, 2009. 64(1): p. 93–109. [DOI] [PubMed] [Google Scholar]
  • 27.Kohman RA and Rhodes JS, Neurogenesis, inflammation and behavior. Brain Behav Immun, 2013. 27(1): p. 22–32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Levin SG and Godukhin OV, Modulating Effect of Cytokines on Mechanisms of Synaptic Plasticity in the Brain. Biochemistry (Mosc), 2017. 82(3): p. 264–274. [DOI] [PubMed] [Google Scholar]
  • 29.Pribiag H and Stellwagen D, Neuroimmune regulation of homeostatic synaptic plasticity. Neuropharmacology, 2014. 78: p. 13–22. [DOI] [PubMed] [Google Scholar]
  • 30.Williamson LL and Bilbo SD, Chemokines and the hippocampus: a new perspective on hippocampal plasticity and vulnerability. Brain Behav Immun, 2013. 30: p. 186–94. [DOI] [PubMed] [Google Scholar]
  • 31.Marin I and Kipnis J, Learning and memory … and the immune system. Learn Mem, 2013. 20(10): p. 601–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Barker BR, Taxman DJ, and Ting JP, Cross-regulation between the IL-1beta/IL-18 processing inflammasome and other inflammatory cytokines. Curr Opin Immunol, 2011. 23(5): p. 591–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Lobo-Silva D, et al. , Balancing the immune response in the brain: IL-10 and its regulation. Journal of Neuroinflammation, 2016. 13(1): p. 297. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Bardou I, et al. , Age and duration of inflammatory environment differentially affect the neuroimmune response and catecholaminergic neurons in the midbrain and brainstem. Neurobiol Aging, 2014. 35(5): p. 1065–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Pascual M, et al. , Gender differences in the inflammatory cytokine and chemokine profiles induced by binge ethanol drinking in adolescence. Addict Biol, 2016. [DOI] [PubMed] [Google Scholar]
  • 36.Knapp DJ, et al. , Stress and Withdrawal from Chronic Ethanol Induce Selective Changes in Neuroimmune mRNAs in Differing Brain Sites. Brain Sci, 2016. 6(3). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Biswas SK and Lopez-Collazo E, Endotoxin tolerance: new mechanisms, molecules and clinical significance. Trends Immunol, 2009. 30(10): p. 475–87. [DOI] [PubMed] [Google Scholar]
  • 38.Marshall SA, Geil CR, and Nixon K, Prior Binge Ethanol Exposure Potentiates the Microglial Response in a Model of Alcohol-Induced Neurodegeneration. Brain Sci, 2016. 6(2). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Topper LA, Baculis BC, and Valenzuela CF, Exposure of neonatal rats to alcohol has differential effects on neuroinflammation and neuronal survival in the cerebellum and hippocampus. J Neuroinflammation, 2015. 12: p. 160. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Pastor IJ, et al. , Interleukin-1 gene cluster polymorphisms and alcoholism in Spanish men. Alcohol Alcohol, 2005. 40(3): p. 181–6. [DOI] [PubMed] [Google Scholar]
  • 41.Marcos M, et al. , Interleukin-10 gene polymorphism is associated with alcoholism but not with alcoholic liver disease. Alcohol Alcohol, 2008. 43(5): p. 523–8. [DOI] [PubMed] [Google Scholar]
  • 42.Sery O, et al. , Association between −174 G/C polymorphism of interleukin-6 gene and alcoholism. Acta Neuropsychiatr, 2003. 15(5): p. 257–61. [DOI] [PubMed] [Google Scholar]
  • 43.Mulligan MK, et al. , Toward understanding the genetics of alcohol drinking through transcriptome meta-analysis. Proc Natl Acad Sci U S A, 2006. 103(16): p. 6368–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.June HL, et al. , CRF-amplified neuronal TLR4/MCP-1 signaling regulates alcohol self-administration. Neuropsychopharmacology, 2015. 40(6): p. 1549–59. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Bajo M, et al. , Role of the IL-1 receptor antagonist in ethanol-induced regulation of GABAergic transmission in the central amygdala. Brain Behav Immun, 2015. 45: p. 189–97. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Bajo M, et al. , Innate immune factors modulate ethanol interaction with GABAergic transmission in mouse central amygdala. Brain Behav Immun, 2014. 40: p. 191–202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Blednov YA, et al. , Role of interleukin-1 receptor signaling in the behavioral effects of ethanol and benzodiazepines. Neuropharmacology, 2015. 95: p. 309–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Blednov YA, et al. , Loss of ethanol conditioned taste aversion and motor stimulation in knockin mice with ethanol-insensitive alpha2-containing GABA(A) receptors. J Pharmacol Exp Ther, 2011. 336(1): p. 145–54. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Blednov YA, et al. , Neuroimmune regulation of alcohol consumption: behavioral validation of genes obtained from genomic studies. Addict Biol, 2012. 17(1): p. 108–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Blednov YA, et al. , Perturbation of chemokine networks by gene deletion alters the reinforcing actions of ethanol. Behav Brain Res, 2005. 165(1): p. 110–25. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Wu Y, et al. , Attenuation of microglial and IL-1 signaling protects mice from acute alcohol-induced sedation and/or motor impairment. Brain Behav Immun, 2011. 25 Suppl 1: p. S155–64. [DOI] [PubMed] [Google Scholar]
  • 52.Marshall SA, et al. , Modulation of Binge-like Ethanol Consumption by IL-10 Signaling in the Basolateral Amygdala. J Neuroimmune Pharmacol, 2017. 12(2): p. 249–259. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Lippai D, et al. , Alcohol-induced IL-1beta in the brain is mediated by NLRP3/ASC inflammasome activation that amplifies neuroinflammation. J Leukoc Biol, 2013. 94(1): p. 171–82. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Crews FT and Vetreno RP, Neuroimmune basis of alcoholic brain damage. Int Rev Neurobiol, 2014. 118: p. 315–57. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Dinarello CA, Interleukin-1 in the pathogenesis and treatment of inflammatory diseases. Blood, 2011. 117(14): p. 3720–32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Parker LC, et al. , IL-1beta induced changes in hypothalamic IL-1R1 and IL-1R2 mRNA expression in the rat. Brain Res Mol Brain Res, 2000. 79(1–2): p. 156–8. [DOI] [PubMed] [Google Scholar]
  • 57.French RA, et al. , Expression and localization of p80 and p68 interleukin-1 receptor proteins in the brain of adult mice. J Neuroimmunol, 1999. 93(1–2): p. 194–202. [DOI] [PubMed] [Google Scholar]
  • 58.Ericsson A, et al. , Type 1 interleukin-1 receptor in the rat brain: distribution, regulation, and relationship to sites of IL-1-induced cellular activation. J Comp Neurol, 1995. 361(4): p. 681–98. [DOI] [PubMed] [Google Scholar]
  • 59.Heida JG and Pittman QJ, Causal links between brain cytokines and experimental febrile convulsions in the rat. Epilepsia, 2005. 46(12): p. 1906–13. [DOI] [PubMed] [Google Scholar]
  • 60.Johnson JD, et al. , The role of IL-1beta in stress-induced sensitization of proinflammatory cytokine and corticosterone responses. Neuroscience, 2004. 127(3): p. 569–77. [DOI] [PubMed] [Google Scholar]
  • 61.Hosoi T, Okuma Y, and Nomura Y, Leptin induces IL-1 receptor antagonist expression in the brain. Biochem Biophys Res Commun, 2002. 294(2): p. 215–9. [DOI] [PubMed] [Google Scholar]
  • 62.Gayle D, et al. , Basal and IL-1beta-stimulated cytokine and neuropeptide mRNA expression in brain regions of young and old Long-Evans rats. Brain Res Mol Brain Res, 1999. 70(1): p. 92–100. [DOI] [PubMed] [Google Scholar]
  • 63.Cartmell T, Luheshi GN, and Rothwell NJ, Brain sites of action of endogenous interleukin-1 in the febrile response to localized inflammation in the rat. J Physiol, 1999. 518 (Pt 2): p. 585–94. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Taishi P, et al. , Diurnal variations of interleukin-1 beta mRNA and beta-actin mRNA in rat brain. J Neuroimmunol, 1997. 75(1–2): p. 69–74. [DOI] [PubMed] [Google Scholar]
  • 65.Quan N, Whiteside M, and Herkenham M, Time course and localization patterns of interleukin-1beta messenger RNA expression in brain and pituitary after peripheral administration of lipopolysaccharide. Neuroscience, 1998. 83(1): p. 281–93. [DOI] [PubMed] [Google Scholar]
  • 66.Quan N, et al. , Detection of interleukin-1 bioactivity in various brain regions of normal healthy rats. Neuroimmunomodulation, 1996. 3(1): p. 47–55. [DOI] [PubMed] [Google Scholar]
  • 67.Hagan P, Poole S, and Bristow AF, Endotoxin-stimulated production of rat hypothalamic interleukin-1 beta in vivo and in vitro, measured by specific immunoradiometric assay. J Mol Endocrinol, 1993. 11(1): p. 31–6. [DOI] [PubMed] [Google Scholar]
  • 68.Blanco AM, et al. , Involvement of TLR4/type I IL-1 receptor signaling in the induction of inflammatory mediators and cell death induced by ethanol in cultured astrocytes. J Immunol, 2005. 175. [DOI] [PubMed] [Google Scholar]
  • 69.Blanco AM and Guerri C, Ethanol intake enhances inflammatory mediators in brain: role of glial cells and TLR4/IL-1RI receptors. Front Biosci, 2007. 12: p. 2616–30. [DOI] [PubMed] [Google Scholar]
  • 70.Gardoni F, et al. , Distribution of interleukin-1 receptor complex at the synaptic membrane driven by interleukin-1beta and NMDA stimulation. J Neuroinflammation, 2011. 8(1): p. 14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Viviani B, et al. , Early maternal deprivation immunologically primes hippocampal synapses by redistributing interleukin-1 receptor type I in a sex dependent manner. Brain Behav Immun, 2014. 35: p. 135–43. [DOI] [PubMed] [Google Scholar]
  • 72.Keyel PA, How is inflammation initiated? Individual influences of IL-1, IL-18 and HMGB1. Cytokine, 2014. 69(1): p. 136–45. [DOI] [PubMed] [Google Scholar]
  • 73.Lamkanfi M and Dixit VM, Mechanisms and functions of inflammasomes. Cell, 2014. 157(5): p. 1013–22. [DOI] [PubMed] [Google Scholar]
  • 74.Zou J and Crews FT, Inflammasome-IL-1beta Signaling Mediates Ethanol Inhibition of Hippocampal Neurogenesis. Front Neurosci, 2012. 6: p. 77. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Wang X, et al. , Ethanol directly induced HMGB1 release through NOX2/NLRP1 inflammasome in neuronal cells. Toxicology, 2015. 334: p. 104–10. [DOI] [PubMed] [Google Scholar]
  • 76.Alfonso-Loeches S, et al. , Role of mitochondria ROS generation in ethanol-induced NLRP3 inflammasome activation and cell death in astroglial cells. Front Cell Neurosci, 2014. 8: p. 216. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Alfonso-Loeches S, et al. , Ethanol-Induced TLR4/NLRP3 Neuroinflammatory Response in Microglial Cells Promotes Leukocyte Infiltration Across the BBB. Neurochem Res, 2015. [DOI] [PubMed] [Google Scholar]
  • 78.Cohen P, The TLR and IL-1 signalling network at a glance. J Cell Sci, 2014. 127(Pt 11): p. 2383–90. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Garlanda C, Dinarello CA, and Mantovani A, The interleukin-1 family: back to the future. Immunity, 2013. 39(6): p. 1003–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Krumm B, Xiang Y, and Deng J, Structural biology of the IL-1 superfamily: key cytokines in the regulation of immune and inflammatory responses. Protein Sci, 2014. 23(5): p. 526–38. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Bajo M, et al. , IL-1 interacts with ethanol effects on GABAergic transmission in the mouse central amygdala. Front Pharmacol, 2015. 6: p. 49. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Szabo G and Lippai D, Converging actions of alcohol on liver and brain immune signaling. Int Rev Neurobiol, 2014. 118: p. 359–80. [DOI] [PubMed] [Google Scholar]
  • 83.Motoki K, et al. , The direct excitatory effect of IL-1beta on cerebellar Purkinje cell. Biochem Biophys Res Commun, 2009. 379(3): p. 665–8. [DOI] [PubMed] [Google Scholar]
  • 84.Brambilla D, et al. , Interleukin-1 inhibits firing of serotonergic neurons in the dorsal raphe nucleus and enhances GABAergic inhibitory post-synaptic potentials. Eur J Neurosci, 2007. 26(7): p. 1862–9. [DOI] [PubMed] [Google Scholar]
  • 85.Manfridi A, et al. , Interleukin-1beta enhances non-rapid eye movement sleep when microinjected into the dorsal raphe nucleus and inhibits serotonergic neurons in vitro. Eur J Neurosci, 2003. 18(5): p. 1041–9. [DOI] [PubMed] [Google Scholar]
  • 86.Lukats B, Egyed R, and Karadi Z, Single neuron activity changes to interleukin-1beta in the orbitofrontal cortex of the rat. Brain Res, 2005. 1038(2): p. 243–6. [DOI] [PubMed] [Google Scholar]
  • 87.Serantes R, et al. , Interleukin-1beta enhances GABAA receptor cell-surface expression by a phosphatidylinositol 3-kinase/Akt pathway: relevance to sepsis-associated encephalopathy. J Biol Chem, 2006. 281(21): p. 14632–43. [DOI] [PubMed] [Google Scholar]
  • 88.Wang DS, et al. , Memory deficits induced by inflammation are regulated by alpha5-subunit-containing GABAA receptors. Cell Rep, 2012. 2(3): p. 488–96. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Viviani B, et al. , Interleukin-1beta enhances NMDA receptor-mediated intracellular calcium increase through activation of the Src family of kinases. J Neurosci, 2003. 23(25): p. 8692–700. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Feleder C, et al. , Interleukin-1 stimulates hypothalamic inhibitory amino acid neurotransmitter release. Neuroimmunomodulation, 1998. 5(1–2): p. 1–4. [DOI] [PubMed] [Google Scholar]
  • 91.Miller LG, et al. , Interleukin-1 augments gamma-aminobutyric acidA receptor function in brain. Mol Pharmacol, 1991. 39(2): p. 105–8. [PubMed] [Google Scholar]
  • 92.Mishra A, et al. , Synapse loss induced by interleukin-1beta requires pre- and post-synaptic mechanisms. J Neuroimmune Pharmacol, 2012. 7(3): p. 571–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Murray CA, et al. , Interleukin-1 beta inhibits glutamate release in hippocampus of young, but not aged, rats. Neurobiol Aging, 1997. 18(3): p. 343–8. [DOI] [PubMed] [Google Scholar]
  • 94.Sama MA, et al. , Interleukin-1beta-dependent signaling between astrocytes and neurons depends critically on astrocytic calcineurin/NFAT activity. J Biol Chem, 2008. 283(32): p. 21953–64. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Tabarean IV, Korn H, and Bartfai T, Interleukin-1beta induces hyperpolarization and modulates synaptic inhibition in preoptic and anterior hypothalamic neurons. Neuroscience, 2006. 141(4): p. 1685–95. [DOI] [PubMed] [Google Scholar]
  • 96.Zeise ML, et al. , Interleukin-1beta does not increase synaptic inhibition in hippocampal CA3 pyramidal and dentate gyrus granule cells of the rat in vitro. Brain Res, 1997. 768(1–2): p. 341–4. [DOI] [PubMed] [Google Scholar]
  • 97.Zeise ML, Madamba S, and Siggins GR, Interleukin-1 beta increases synaptic inhibition in rat hippocampal pyramidal neurons in vitro. Regul Pept, 1992. 39(1): p. 1–7. [DOI] [PubMed] [Google Scholar]
  • 98.Dunn AJ, Wang J, and Ando T, Effects of cytokines on cerebral neurotransmission. Comparison with the effects of stress. Adv Exp Med Biol, 1999. 461: p. 117–27. [DOI] [PubMed] [Google Scholar]
  • 99.Lin HW, et al. , Astrogliosis is delayed in type 1 interleukin-1 receptor-null mice following a penetrating brain injury. J Neuroinflammation, 2006. 3: p. 15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.O’Connor JJ and Coogan AN, Actions of the pro-inflammatory cytokine IL-1 beta on central synaptic transmission. Exp Physiol, 1999. 84(4): p. 601–14. [PubMed] [Google Scholar]
  • 101.Rothwell NJ and Luheshi GN, Interleukin 1 in the brain: biology, pathology and therapeutic target. Trends Neurosci, 2000. 23(12): p. 618–25. [DOI] [PubMed] [Google Scholar]
  • 102.Towne JE, et al. , Interleukin (IL)-1F6, IL-1F8, and IL-1F9 signal through IL-1Rrp2 and IL-1RAcP to activate the pathway leading to NF-kappaB and MAPKs. J Biol Chem, 2004. 279(14): p. 13677–88. [DOI] [PubMed] [Google Scholar]
  • 103.O’Neill LA, The interleukin-1 receptor/Toll-like receptor superfamily: 10 years of progress. Immunol Rev, 2008. 226: p. 10–8. [DOI] [PubMed] [Google Scholar]
  • 104.Breese GR, et al. , Repeated lipopolysaccharide (LPS) or cytokine treatments sensitize ethanol withdrawal-induced anxiety-like behavior. Neuropsychopharmacology, 2008. 33(4): p. 867–76. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Marshall SA, et al. , IL-1 receptor signaling in the basolateral amygdala modulates binge-like ethanol consumption in male C57BL/6J mice. Brain Behav Immun, 2016. 51: p. 258–67. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Rajayer SR, et al. , Cold-inducible RNA-binding protein is an important mediator of alcohol-induced brain inflammation. PLoS One, 2013. 8(11): p. e79430. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Boyadjieva NI and Sarkar DK, Role of microglia in ethanol’s apoptotic action on hypothalamic neuronal cells in primary cultures. Alcohol Clin Exp Res, 2010. 34(11): p. 1835–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Lawrimore CJ and Crews FT, Ethanol, TLR3, and TLR4 Agonists Have Unique Innate Immune Responses in Neuron-Like SH-SY5Y and Microglia-Like BV2. Alcohol Clin Exp Res, 2017. 41(5): p. 939–954. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Lippai D, et al. , Chronic alcohol-induced microRNA-155 contributes to neuroinflammation in a TLR4-dependent manner in mice. PLoS One, 2013. 8(8): p. e70945. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Qin L, et al. , Increased systemic and brain cytokine production and neuroinflammation by endotoxin following ethanol treatment. J Neuroinflammation, 2008. 5: p. 10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Valles SL, et al. , Chronic ethanol treatment enhances inflammatory mediators and cell death in the brain and in astrocytes. Brain Pathol, 2004. 14(4): p. 365–71. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Koob GF and Volkow ND, Neurobiology of addiction: a neurocircuitry analysis. Lancet Psychiatry, 2016. 3(8): p. 760–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Davis M, Rainnie D, and Cassell M, Neurotransmission in the rat amygdala related to fear and anxiety. Trends Neurosci, 1994. 17(5): p. 208–14. [DOI] [PubMed] [Google Scholar]
  • 114.Roberto M, et al. , Corticotropin releasing factor-induced amygdala gamma-aminobutyric Acid release plays a key role in alcohol dependence. Biol Psychiatry, 2010. 67(9): p. 831–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Klausberger T and Somogyi P, Neuronal diversity and temporal dynamics: the unity of hippocampal circuit operations. Science, 2008. 321(5885): p. 53–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Nuss P, Anxiety disorders and GABA neurotransmission: a disturbance of modulation. Neuropsychiatr Dis Treat, 2015. 11: p. 165–75. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Holmes A and Wellman CL, Stress-induced prefrontal reorganization and executive dysfunction in rodents. Neurosci Biobehav Rev, 2009. 33(6): p. 773–83. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Silveri MM, GABAergic contributions to alcohol responsivity during adolescence: insights from preclinical and clinical studies. Pharmacol Ther, 2014. 143(2): p. 197–216. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Blednov YA, et al. , Linking GABA(A) receptor subunits to alcohol-induced conditioned taste aversion and recovery from acute alcohol intoxication. Neuropharmacology, 2013. 67: p. 46–56. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Blednov YA, et al. , Deletion of the alpha1 or beta2 subunit of GABAA receptors reduces actions of alcohol and other drugs. J Pharmacol Exp Ther, 2003. 304(1): p. 30–6. [DOI] [PubMed] [Google Scholar]
  • 121.Boehm SL 2nd, et al. , gamma-Aminobutyric acid A receptor subunit mutant mice: new perspectives on alcohol actions. Biochem Pharmacol, 2004. 68(8): p. 1581–602. [DOI] [PubMed] [Google Scholar]
  • 122.Roberts AJ, Cole M, and Koob GF, Intra-amygdala muscimol decreases operant ethanol self-administration in dependent rats. Alcohol Clin Exp Res, 1996. 20(7): p. 1289–98. [DOI] [PubMed] [Google Scholar]
  • 123.Hyytia P and Koob GF, GABAA receptor antagonism in the extended amygdala decreases ethanol self-administration in rats. Eur J Pharmacol, 1995. 283(1–3): p. 151–9. [DOI] [PubMed] [Google Scholar]
  • 124.Uryu K, et al. , Fine structure and possible origins of nerve fibers with corticotropin-releasing factor-like immunoreactivity in the rat central amygdaloid nucleus. Brain Res, 1992. 577(1): p. 175–9. [DOI] [PubMed] [Google Scholar]
  • 125.Konsman JP, et al. , Central nervous action of interleukin-1 mediates activation of limbic structures and behavioural depression in response to peripheral administration of bacterial lipopolysaccharide. Eur J Neurosci, 2008. 28(12): p. 2499–510. [DOI] [PubMed] [Google Scholar]
  • 126.Nie Z, et al. , Ethanol augments GABAergic transmission in the central amygdala via CRF1 receptors. Science, 2004. 303(5663): p. 1512–4. [DOI] [PubMed] [Google Scholar]
  • 127.Nie Z, et al. , Presynaptic CRF1 receptors mediate the ethanol enhancement of GABAergic transmission in the mouse central amygdala. ScientificWorldJournal, 2009. 9: p. 68–85. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Roberto M, et al. , Ethanol increases GABAergic transmission at both pre- and postsynaptic sites in rat central amygdala neurons. Proc Natl Acad Sci U S A, 2003. 100(4): p. 2053–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Knapp DJ, et al. , Cytokine involvement in stress may depend on corticotrophin releasing factor to sensitize ethanol withdrawal anxiety. Brain Behav Immun, 2011. 25 Suppl 1: p. S146–54. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Frost P, et al. , IL-1 receptor type I gene expression in the amygdala of inflammatory susceptible Lewis and inflammatory resistant Fischer rats. J Neuroimmunol, 2001. 121(1–2): p. 32–9. [DOI] [PubMed] [Google Scholar]
  • 131.Eriksson C, et al. , Increased expression of mRNA encoding interleukin-1beta and caspase-1, and the secreted isoform of interleukin-1 receptor antagonist in the rat brain following systemic kainic acid administration. J Neurosci Res, 2000. 60(2): p. 266–79. [DOI] [PubMed] [Google Scholar]
  • 132.Dayas CV, et al. , Stressor categorization: acute physical and psychological stressors elicit distinctive recruitment patterns in the amygdala and in medullary noradrenergic cell groups. Eur J Neurosci, 2001. 14(7): p. 1143–52. [DOI] [PubMed] [Google Scholar]
  • 133.Yu B and Shinnick-Gallagher P, Interleukin-1 beta inhibits synaptic transmission and induces membrane hyperpolarization in amygdala neurons. J Pharmacol Exp Ther, 1994. 271(2): p. 590–600. [PubMed] [Google Scholar]
  • 134.Takao T, et al. , Interleukin-1 receptors in mouse brain: characterization and neuronal localization. Endocrinology, 1990. 127(6): p. 3070–8. [DOI] [PubMed] [Google Scholar]
  • 135.Vezzani A, et al. , Interleukin-1beta immunoreactivity and microglia are enhanced in the rat hippocampus by focal kainate application: functional evidence for enhancement of electrographic seizures. J Neurosci, 1999. 19(12): p. 5054–65. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Zhu YN, et al. , Periplocoside E inhibits experimental allergic encephalomyelitis by suppressing interleukin 12-dependent CCR5 expression and interferon-gamma-dependent CXCR3 expression in T lymphocytes. J Pharmacol Exp Ther, 2006. 318(3): p. 1153–62. [DOI] [PubMed] [Google Scholar]
  • 137.Goshen I, et al. , A dual role for interleukin-1 in hippocampal-dependent memory processes. Psychoneuroendocrinology, 2007. 32(8–10): p. 1106–15. [DOI] [PubMed] [Google Scholar]
  • 138.Ryan SM and Nolan YM, Neuroinflammation negatively affects adult hippocampal neurogenesis and cognition: can exercise compensate? Neurosci Biobehav Rev, 2016. 61: p. 121–31. [DOI] [PubMed] [Google Scholar]
  • 139.Kreisel T, et al. , Dynamic microglial alterations underlie stress-induced depressive-like behavior and suppressed neurogenesis. Mol Psychiatry, 2014. 19. [DOI] [PubMed] [Google Scholar]
  • 140.Koo JW and Duman RS, IL-1beta is an essential mediator of the antineurogenic and anhedonic effects of stress. Proc Natl Acad Sci U S A, 2008. 105(2): p. 751–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Geil CR, et al. , Alcohol and adult hippocampal neurogenesis: promiscuous drug, wanton effects. Prog Neuropsychopharmacol Biol Psychiatry, 2014. 54: p. 103–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Erta M, Quintana A, and Hidalgo J, Interleukin-6, a major cytokine in the central nervous system. Int J Biol Sci, 2012. 8(9): p. 1254–66. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Wallenius V, et al. , Interleukin-6-deficient mice develop mature-onset obesity. Nat Med, 2002. 8(1): p. 75–9. [DOI] [PubMed] [Google Scholar]
  • 144.Herrmann O, et al. , Regulation of body temperature and neuroprotection by endogenous interleukin-6 in cerebral ischemia. J Cereb Blood Flow Metab, 2003. 23(4): p. 406–15. [DOI] [PubMed] [Google Scholar]
  • 145.Chai Z, et al. , Interleukin (IL)-6 gene expression in the central nervous system is necessary for fever response to lipopolysaccharide or IL-1 beta: a study on IL-6-deficient mice. J Exp Med, 1996. 183(1): p. 311–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Mastorakos G, Chrousos GP, and Weber JS, Recombinant interleukin-6 activates the hypothalamic-pituitary-adrenal axis in humans. J Clin Endocrinol Metab, 1993. 77(6): p. 1690–4. [DOI] [PubMed] [Google Scholar]
  • 147.Morrow JD and Opp MR, Sleep-wake behavior and responses of interleukin-6-deficient mice to sleep deprivation. Brain Behav Immun, 2005. 19(1): p. 28–39. [DOI] [PubMed] [Google Scholar]
  • 148.Dimitrov S, et al. , Sleep enhances IL-6 trans-signaling in humans. FASEB J, 2006. 20(12): p. 2174–6. [DOI] [PubMed] [Google Scholar]
  • 149.Ye SM and Johnson RW, Increased interleukin-6 expression by microglia from brain of aged mice. J Neuroimmunol, 1999. 93(1–2): p. 139–48. [DOI] [PubMed] [Google Scholar]
  • 150.Fattori E, et al. , IL-6 expression in neurons of transgenic mice causes reactive astrocytosis and increase in ramified microglial cells but no neuronal damage. Eur J Neurosci, 1995. 7(12): p. 2441–9. [DOI] [PubMed] [Google Scholar]
  • 151.Choi SS, et al. , Human astrocytes: secretome profiles of cytokines and chemokines. PLoS One, 2014. 9(4): p. e92325. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Farina C, Aloisi F, and Meinl E, Astrocytes are active players in cerebral innate immunity. Trends Immunol, 2007. 28(3): p. 138–45. [DOI] [PubMed] [Google Scholar]
  • 153.Norris JG and Benveniste EN, Interleukin-6 production by astrocytes: induction by the neurotransmitter norepinephrine. J Neuroimmunol, 1993. 45(1–2): p. 137–45. [DOI] [PubMed] [Google Scholar]
  • 154.Maimone D, et al. , Norepinephrine and vasoactive intestinal peptide induce IL-6 secretion by astrocytes: synergism with IL-1 beta and TNF alpha. J Neuroimmunol, 1993. 47(1): p. 73–81. [DOI] [PubMed] [Google Scholar]
  • 155.Lieb K, et al. , Serotonin via 5-HT7 receptors activates p38 mitogen-activated protein kinase and protein kinase C epsilon resulting in interleukin-6 synthesis in human U373 MG astrocytoma cells. J Neurochem, 2005. 93(3): p. 549–59. [DOI] [PubMed] [Google Scholar]
  • 156.Gitter BD, et al. , Interleukin-6 secretion from human astrocytoma cells induced by substance P. J Neuroimmunol, 1994. 51(1): p. 101–8. [DOI] [PubMed] [Google Scholar]
  • 157.Schaper F and Rose-John S, Interleukin-6: Biology, signaling and strategies of blockade. Cytokine Growth Factor Rev, 2015. 26(5): p. 475–87. [DOI] [PubMed] [Google Scholar]
  • 158.Li X, et al. , Interleukin-6 inhibits voltage-gated sodium channel activity of cultured rat spinal cord neurons. Acta Neuropsychiatr, 2014. 26(3): p. 170–7. [DOI] [PubMed] [Google Scholar]
  • 159.Conroy SM, et al. , Interleukin-6 produces neuronal loss in developing cerebellar granule neuron cultures. J Neuroimmunol, 2004. 155(1–2): p. 43–54. [DOI] [PubMed] [Google Scholar]
  • 160.Orellana DI, et al. , Role of the JAKs/STATs pathway in the intracellular calcium changes induced by interleukin-6 in hippocampal neurons. Neurotox Res, 2005. 8(3–4): p. 295–304. [DOI] [PubMed] [Google Scholar]
  • 161.Qiu Z, Parsons KL, and Gruol DL, Interleukin-6 selectively enhances the intracellular calcium response to NMDA in developing CNS neurons. J Neurosci, 1995. 15(10): p. 6688–99. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Garcia-Oscos F, et al. , The stress-induced cytokine interleukin-6 decreases the inhibition/excitation ratio in the rat temporal cortex via trans-signaling. Biol Psychiatry, 2012. 71(7): p. 574–82. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Hernandez RV, et al. , Transgenic mice with increased astrocyte expression of IL-6 show altered effects of acute ethanol on synaptic function. Neuropharmacology, 2016. 103: p. 27–43. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Sarc L, Wraber B, and Lipnik-Stangelj M, Ethanol and acetaldehyde disturb TNF-alpha and IL-6 production in cultured astrocytes. Hum Exp Toxicol, 2011. 30(9): p. 1256–65. [DOI] [PubMed] [Google Scholar]
  • 165.Chaturvedi LS, Zhang P, and Basson MD, Effects of extracellular pressure and alcohol on the microglial response to inflammatory stimulation. Am J Surg, 2012. 204(5): p. 602–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Wilhelm CJ, et al. , Astrocyte Dysfunction Induced by Alcohol in Females but Not Males. Brain Pathol, 2016. 26(4): p. 433–51. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Gano A, et al. , Conditioned effects of ethanol on the immune system. Exp Biol Med (Maywood), 2017. 242(7): p. 718–730. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Bray JG, et al. , Synaptic plasticity in the hippocampus shows resistance to acute ethanol exposure in transgenic mice with astrocyte-targeted enhanced CCL2 expression. Neuropharmacology, 2013. 67: p. 115–25. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Kwilasz AJ, et al. , The therapeutic potential of interleukin-10 in neuroimmune diseases. Neuropharmacology, 2015. 96. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Fickenscher H, et al. , The interleukin-10 family of cytokines. Trends Immunol, 2002. 23(2): p. 89–96. [DOI] [PubMed] [Google Scholar]
  • 171.Sharma S, et al. , IL-10 directly protects cortical neurons by activating PI-3 kinase and STAT-3 pathways. Brain Res, 2011. 1373: p. 189–94. [DOI] [PubMed] [Google Scholar]
  • 172.Norkina O, et al. , Acute alcohol activates STAT3, AP-1, and Sp-1 transcription factors via the family of Src kinases to promote IL-10 production in human monocytes. J Leukoc Biol, 2007. 82(3): p. 752–62. [DOI] [PubMed] [Google Scholar]
  • 173.Curtale G, et al. , Negative regulation of Toll-like receptor 4 signaling by IL-10-dependent microRNA-146b. Proc Natl Acad Sci U S A, 2013. 110(28): p. 11499–504. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Segev-Amzaleg N, Trudler D, and Frenkel D, Preconditioning to mild oxidative stress mediates astroglial neuroprotection in an IL-10-dependent manner. Brain Behav Immun, 2013. 30: p. 176–85. [DOI] [PubMed] [Google Scholar]
  • 175.Lim SH, et al. , Neuronal synapse formation induced by microglia and interleukin 10. PLoS One, 2013. 8(11): p. e81218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Suryanarayanan A, et al. , Role of interleukin-10 (IL-10) in regulation of GABAergic transmission and acute response to ethanol. Neuropharmacology, 2016. 107: p. 181–188. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Ponomarev I, et al. , Gene coexpression networks in human brain identify epigenetic modifications in alcohol dependence. J Neurosci, 2012. 32(5): p. 1884–97. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Marshall SA, et al. , Microglial activation is not equivalent to neuroinflammation in alcohol-induced neurodegeneration: The importance of microglia phenotype. Neurobiol Dis, 2013. 54: p. 239–51. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Schunck RV, et al. , Protracted alcohol abstinence induces analgesia in rats: Possible relationships with BDNF and interleukin-10. Pharmacol Biochem Behav, 2015. 135: p. 64–9. [DOI] [PubMed] [Google Scholar]
  • 180.Kriegler M, et al. , A novel form of TNF/cachectin is a cell surface cytotoxic transmembrane protein: ramifications for the complex physiology of TNF. Cell, 1988. 53(1): p. 45–53. [DOI] [PubMed] [Google Scholar]
  • 181.Grell M, et al. , The transmembrane form of tumor necrosis factor is the prime activating ligand of the 80 kDa tumor necrosis factor receptor. Cell, 1995. 83(5): p. 793–802. [DOI] [PubMed] [Google Scholar]
  • 182.Walczak H, TNF and ubiquitin at the crossroads of gene activation, cell death, inflammation, and cancer. Immunol Rev, 2011. 244(1): p. 9–28. [DOI] [PubMed] [Google Scholar]
  • 183.Medvedev AE, Sundan A, and Espevik T, Involvement of the tumor necrosis factor receptor p75 in mediating cytotoxicity and gene regulating activities. Eur J Immunol, 1994. 24(11): p. 2842–9. [DOI] [PubMed] [Google Scholar]
  • 184.Rao P, Hsu KC, and Chao MV, Upregulation of NF-kappa B-dependent gene expression mediated by the p75 tumor necrosis factor receptor. J Interferon Cytokine Res, 1995. 15(2): p. 171–7. [DOI] [PubMed] [Google Scholar]
  • 185.Stellwagen D and Malenka RC, Synaptic scaling mediated by glial TNF-alpha. Nature, 2006. 440(7087): p. 1054–9. [DOI] [PubMed] [Google Scholar]
  • 186.Turrigiano GG, The self-tuning neuron: synaptic scaling of excitatory synapses. Cell, 2008. 135(3): p. 422–35. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Kaneko M, et al. , Tumor necrosis factor-alpha mediates one component of competitive, experience-dependent plasticity in developing visual cortex. Neuron, 2008. 58(5): p. 673–80. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Zou JY and Crews FT, TNF alpha potentiates glutamate neurotoxicity by inhibiting glutamate uptake in organotypic brain slice cultures: neuroprotection by NF kappa B inhibition. Brain Res, 2005. 1034(1–2): p. 11–24. [DOI] [PubMed] [Google Scholar]
  • 189.Musumeci G, et al. , Transient receptor potential vanilloid 1 channels modulate the synaptic effects of TNF-alpha and of IL-1beta in experimental autoimmune encephalomyelitis. Neurobiol Dis, 2011. 43(3): p. 669–77. [DOI] [PubMed] [Google Scholar]
  • 190.Takeuchi H, et al. , Tumor necrosis factor-alpha induces neurotoxicity via glutamate release from hemichannels of activated microglia in an autocrine manner. J Biol Chem, 2006. 281(30): p. 21362–8. [DOI] [PubMed] [Google Scholar]
  • 191.Bezzi P, et al. , Neuron-astrocyte cross-talk during synaptic transmission: physiological and neuropathological implications. Prog Brain Res, 2001. 132: p. 255–65. [DOI] [PubMed] [Google Scholar]
  • 192.Stellwagen D, et al. , Differential regulation of AMPA receptor and GABA receptor trafficking by tumor necrosis factor-alpha. J Neurosci, 2005. 25(12): p. 3219–28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.He P, et al. , Genetic deletion of TNF receptor suppresses excitatory synaptic transmission via reducing AMPA receptor synaptic localization in cortical neurons. FASEB J, 2012. 26(1): p. 334–45. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Ferguson AR, et al. , Cell death after spinal cord injury is exacerbated by rapid TNF alpha-induced trafficking of GluR2-lacking AMPARs to the plasma membrane. J Neurosci, 2008. 28(44): p. 11391–400. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Zhao P, et al. , Cannabinoid receptor activation reduces TNFalpha-induced surface localization of AMPAR-type glutamate receptors and excitotoxicity. Neuropharmacology, 2010. 58(2): p. 551–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Beattie MS, et al. , Cell death in models of spinal cord injury. Prog Brain Res, 2002. 137: p. 37–47. [DOI] [PubMed] [Google Scholar]
  • 197.Leonoudakis D, et al. , TNFalpha-induced AMPA-receptor trafficking in CNS neurons; relevance to excitotoxicity? Neuron Glia Biol, 2004. 1(3): p. 263–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Ming Z, Criswell HE, and Breese GR, Evidence for TNFalpha action on excitatory and inhibitory neurotransmission in the central amygdala: a brain site influenced by stress. Brain Behav Immun, 2013. 33: p. 102–11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Roberto M, Gilpin NW, and Siggins GR, The Central Amygdala and Alcohol: Role of gamma-Aminobutyric Acid, Glutamate, and Neuropeptides. Cold Spring Harb Perspect Med, 2012. 2(12). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Kiefer F, et al. , Alcohol intake, tumour necrosis factor-alpha, leptin and craving: factors of a possibly vicious circle? Alcohol Alcohol, 2002. 37(4): p. 401–4. [DOI] [PubMed] [Google Scholar]
  • 201.Reaux-Le Goazigo A, et al. , Current status of chemokines in the adult CNS. Prog Neurobiol, 2013. 104: p. 67–92. [DOI] [PubMed] [Google Scholar]
  • 202.Semple BD, Frugier T, and Morganti-Kossmann MC, CCL2 modulates cytokine production in cultured mouse astrocytes. J Neuroinflammation, 2010. 7: p. 67. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Banisadr G, et al. , Highly regionalized neuronal expression of monocyte chemoattractant protein-1 (MCP-1/CCL2) in rat brain: evidence for its colocalization with neurotransmitters and neuropeptides. J Comp Neurol, 2005. 489(3): p. 275–92. [DOI] [PubMed] [Google Scholar]
  • 204.Jung H, et al. , Monocyte chemoattractant protein-1 functions as a neuromodulator in dorsal root ganglia neurons. J Neurochem, 2008. 104(1): p. 254–63. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205.Dansereau MA, et al. , Spinal CCL2 pronociceptive action is no longer effective in CCR2 receptor antagonist-treated rats. J Neurochem, 2008. 106(2): p. 757–69. [DOI] [PubMed] [Google Scholar]
  • 206.Conductier G, et al. , The role of monocyte chemoattractant protein MCP1/CCL2 in neuroinflammatory diseases. J Neuroimmunol, 2010. 224(1–2): p. 93–100. [DOI] [PubMed] [Google Scholar]
  • 207.Yamasaki T, et al. , Tumor microvasculature with endothelial fenestrations in VHL null clear cell renal cell carcinomas as a potent target of anti-angiogenic therapy. Cancer Sci, 2012. 103(11): p. 2027–37. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Boddeke EW, et al. , Cultured rat microglia express functional beta-chemokine receptors. J Neuroimmunol, 1999. 98(2): p. 176–84. [DOI] [PubMed] [Google Scholar]
  • 209.Andjelkovic AV, et al. , Functional expression of CCR2 by human fetal astrocytes. J Neurosci Res, 2002. 70(2): p. 219–31. [DOI] [PubMed] [Google Scholar]
  • 210.Rostene W, Kitabgi P, and Parsadaniantz SM, Chemokines: a new class of neuromodulator? Nat Rev Neurosci, 2007. 8(11): p. 895–903. [DOI] [PubMed] [Google Scholar]
  • 211.Banisadr G, et al. , Constitutive neuronal expression of CCR2 chemokine receptor and its colocalization with neurotransmitters in normal rat brain: functional effect of MCP-1/CCL2 on calcium mobilization in primary cultured neurons. J Comp Neurol, 2005. 492(2): p. 178–92. [DOI] [PubMed] [Google Scholar]
  • 212.Kuang Y, et al. , Selective G protein coupling by C-C chemokine receptors. J Biol Chem, 1996. 271(8): p. 3975–8. [DOI] [PubMed] [Google Scholar]
  • 213.Wain JH, Kirby JA, and Ali S, Leucocyte chemotaxis: Examination of mitogen-activated protein kinase and phosphoinositide 3-kinase activation by Monocyte Chemoattractant Proteins-1, −2, −3 and −4. Clin Exp Immunol, 2002. 127(3): p. 436–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 214.Old EA and Malcangio M, Chemokine mediated neuron-glia communication and aberrant signalling in neuropathic pain states. Curr Opin Pharmacol, 2012. 12(1): p. 67–73. [DOI] [PubMed] [Google Scholar]
  • 215.Dawson J, et al. , Targeting monocyte chemoattractant protein-1 signalling in disease. Expert Opin Ther Targets, 2003. 7(1): p. 35–48. [DOI] [PubMed] [Google Scholar]
  • 216.Zhou Y, et al. , Chemokine CCL2 modulation of neuronal excitability and synaptic transmission in rat hippocampal slices. J Neurochem, 2011. 116(3): p. 406–14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Lewohl JM, et al. , Gene expression in human alcoholism: microarray analysis of frontal cortex. Alcohol Clin Exp Res, 2000. 24(12): p. 1873–82. [PubMed] [Google Scholar]
  • 218.He J and Crews FT, Increased MCP-1 and microglia in various regions of the human alcoholic brain. Exp Neurol, 2008. 210(2): p. 349–58. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219.Umhau JC, et al. , Cerebrospinal fluid monocyte chemoattractant protein-1 in alcoholics: support for a neuroinflammatory model of chronic alcoholism. Alcohol Clin Exp Res, 2014. 38(5): p. 1301–6. [DOI] [PubMed] [Google Scholar]
  • 220.Freeman K, et al. , Temporal changes in innate immune signals in a rat model of alcohol withdrawal in emotional and cardiorespiratory homeostatic nuclei. J Neuroinflammation, 2012. 9: p. 97. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 221.Kane CJ, et al. , Effects of ethanol on immune response in the brain: region-specific changes in adolescent versus adult mice. Alcohol Clin Exp Res, 2014. 38(2): p. 384–91. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Kane CJ, et al. , Effects of ethanol on immune response in the brain: region-specific changes in aged mice. J Neuroinflammation, 2013. 10: p. 66. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Vetreno RP, Qin L, and Crews FT, Increased receptor for advanced glycation end product expression in the human alcoholic prefrontal cortex is linked to adolescent drinking. Neurobiol Dis, 2013. 59: p. 52–62. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 224.Chang GQ, Karatayev O, and Leibowitz SF, Prenatal exposure to ethanol stimulates hypothalamic CCR2 chemokine receptor system: Possible relation to increased density of orexigenic peptide neurons and ethanol drinking in adolescent offspring. Neuroscience, 2015. 310: p. 163–75. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Drew PD, et al. , Pioglitazone blocks ethanol induction of microglial activation and immune responses in the hippocampus, cerebellum, and cerebral cortex in a mouse model of fetal alcohol spectrum disorders. Alcohol Clin Exp Res, 2015. 39(3): p. 445–54. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 226.Pascual M, et al. , Cytokines and chemokines as biomarkers of ethanol-induced neuroinflammation and anxiety-related behavior: role of TLR4 and TLR2. Neuropharmacology, 2015. 89: p. 352–9. [DOI] [PubMed] [Google Scholar]
  • 227.Gruol DL, Impact of Increased Astrocyte Expression of IL-6, CCL2 or CXCL10 in Transgenic Mice on Hippocampal Synaptic Function. Brain Sci, 2016. 6(2). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228.Valenta JP and Gonzales RA, Chronic Intracerebroventricular Infusion of Monocyte Chemoattractant Protein-1 Leads to a Persistent Increase in Sweetened Ethanol Consumption During Operant Self-Administration But Does Not Influence Sucrose Consumption in Long-Evans Rats. Alcohol Clin Exp Res, 2016. 40(1): p. 187–95. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229.Guyon A, et al. , Long term exposure to the chemokine CCL2 activates the nigrostriatal dopamine system: a novel mechanism for the control of dopamine release. Neuroscience, 2009. 162(4): p. 1072–80. [DOI] [PubMed] [Google Scholar]
  • 230.Apartis E, et al. , [Chemokines as new actors in the dopaminergic system]. Biol Aujourdhui, 2010. 204(4): p. 295–300. [DOI] [PubMed] [Google Scholar]
  • 231.Wu Y, et al. , Inhibiting the TLR4-MyD88 signalling cascade by genetic or pharmacological strategies reduces acute alcohol-induced sedation and motor impairment in mice. Br J Pharmacol, 2012. 165(5): p. 1319–29. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 232.Stopponi S, et al. , Activation of PPARgamma by pioglitazone potentiates the effects of naltrexone on alcohol drinking and relapse in msP rats. Alcohol Clin Exp Res, 2013. 37(8): p. 1351–60. [DOI] [PubMed] [Google Scholar]
  • 233.Stopponi S, et al. , Activation of nuclear PPARgamma receptors by the antidiabetic agent pioglitazone suppresses alcohol drinking and relapse to alcohol seeking. Biol Psychiatry, 2011. 69(7): p. 642–9. [DOI] [PubMed] [Google Scholar]
  • 234.Blednov YA, et al. , Peroxisome proliferator-activated receptors alpha and gamma are linked with alcohol consumption in mice and withdrawal and dependence in humans. Alcohol Clin Exp Res, 2015. 39(1): p. 136–45. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 235.Bell RL, et al. , Ibudilast reduces alcohol drinking in multiple animal models of alcohol dependence. Addict Biol, 2015. 20(1): p. 38–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 236.Blednov YA, et al. , Inhibition of phosphodiesterase 4 reduces ethanol intake and preference in C57BL/6J mice. Front Neurosci, 2014. 8: p. 129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 237.Kamdar NK, et al. , Acute effects of naltrexone and GBR 12909 on ethanol drinking-in-the-dark in C57BL/6J mice. Psychopharmacology (Berl), 2007. 192(2): p. 207–17. [DOI] [PubMed] [Google Scholar]
  • 238.Tomie A, Azogu I, and Yu L, Effects of naltrexone on post-abstinence alcohol drinking in C57BL/6NCRL and DBA/2J mice. Prog Neuropsychopharmacol Biol Psychiatry, 2013. 44: p. 240–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 239.Robinson G, et al. , Neuroimmune pathways in alcohol consumption: evidence from behavioral and genetic studies in rodents and humans. Int Rev Neurobiol, 2014. 118: p. 13–39. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 240.Szabo G and Saha B, Alcohol’s Effect on Host Defense. Alcohol Res, 2015. 37(2): p. 159–70. [PMC free article] [PubMed] [Google Scholar]
  • 241.Du L, et al. , Role of Microglia in Neurological Disorders and Their Potentials as a Therapeutic Target. Mol Neurobiol, 2016. [DOI] [PubMed] [Google Scholar]
  • 242.Doremus-Fitzwater TL, et al. , Intoxication- and withdrawal-dependent expression of central and peripheral cytokines following initial ethanol exposure. Alcohol Clin Exp Res, 2014. 38. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 243.Ahlers KE, et al. , Transient activation of microglia following acute alcohol exposure in developing mouse neocortex is primarily driven by BAX-dependent neurodegeneration. Glia, 2015. 63(10): p. 1694–713. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244.Whitman BA, et al. , The cytokine mRNA increase induced by withdrawal from chronic ethanol in the sterile environment of brain is mediated by CRF and HMGB1 release. Alcohol Clin Exp Res, 2013. 37. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 245.Teng SX and Molina PE, Acute alcohol intoxication prolongs neuroinflammation without exacerbating neurobehavioral dysfunction following mild traumatic brain injury. J Neurotrauma, 2014. 31(4): p. 378–86. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 246.Gottesfeld Z, Moore AN, and Dash PK, Acute ethanol intake attenuates inflammatory cytokines after brain injury in rats: a possible role for corticosterone. J Neurotrauma, 2002. 19(3): p. 317–26. [DOI] [PubMed] [Google Scholar]
  • 247.Doremus-Fitzwater TL, et al. , Male adolescent rats display blunted cytokine responses in the CNS after acute ethanol or lipopolysaccharide exposure. Physiol Behav, 2015. 148: p. 131–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 248.Qin L and Crews FT, Focal thalamic degeneration from ethanol and thiamine deficiency is associated with neuroimmune gene induction, microglial activation, and lack of monocarboxylic acid transporters. Alcohol Clin Exp Res, 2014. 38(3): p. 657–71. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 249.Montesinos J, Gil A, and Guerri C, Nalmefene Prevents Alcohol-Induced Neuroinflammation and Alcohol Drinking Preference in Adolescent Female Mice: Role of TLR4. Alcohol Clin Exp Res, 2017. 41(7): p. 1257–1270. [DOI] [PubMed] [Google Scholar]
  • 250.Tiwari V and Chopra K, Attenuation of oxidative stress, neuroinflammation, and apoptosis by curcumin prevents cognitive deficits in rats postnatally exposed to ethanol. Psychopharmacology (Berl), 2012. 224(4): p. 519–35. [DOI] [PubMed] [Google Scholar]
  • 251.Tiwari V and Chopra K, Resveratrol abrogates alcohol-induced cognitive deficits by attenuating oxidative-nitrosative stress and inflammatory cascade in the adult rat brain. Neurochem Int, 2013. 62(6): p. 861–9. [DOI] [PubMed] [Google Scholar]
  • 252.Alfonso-Loeches S, Pascual M, and Guerri C, Gender differences in alcohol-induced neurotoxicity and brain damage. Toxicology, 2013. 311(1–2): p. 27–34. [DOI] [PubMed] [Google Scholar]
  • 253.Schneider R Jr., et al. , N-acetylcysteine Prevents Alcohol Related Neuroinflammation in Rats. Neurochem Res, 2017. 42(8): p. 2135–2141. [DOI] [PubMed] [Google Scholar]
  • 254.Diaz A, et al. , Energy Drink Administration in Combination with Alcohol Causes an Inflammatory Response and Oxidative Stress in the Hippocampus and Temporal Cortex of Rats. Oxid Med Cell Longev, 2016. 2016: p. 8725354. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 255.Zhu Q, et al. , Vitamin E prevents ethanol-induced inflammatory, hormonal, and cytotoxic changes in reproductive tissues. Endocrine, 2007. 32(1): p. 59–68. [DOI] [PubMed] [Google Scholar]
  • 256.Emanuele N, et al. , Effects of chronic ethanol (EtOH) administration on pro-inflammatory cytokines of the hypothalamic-pituitary-gonadal (HPG) axis in female rats. Endocr Res, 2005. 31(1): p. 9–16. [DOI] [PubMed] [Google Scholar]
  • 257.Amin FU, Shah SA, and Kim MO, Glycine inhibits ethanol-induced oxidative stress, neuroinflammation and apoptotic neurodegeneration in postnatal rat brain. Neurochem Int, 2016. 96: p. 1–12. [DOI] [PubMed] [Google Scholar]
  • 258.Zahr NM, et al. , Measurement of serum, liver, and brain cytokine induction, thiamine levels, and hepatopathology in rats exposed to a 4-day alcohol binge protocol. Alcohol Clin Exp Res, 2010. 34(11): p. 1858–70. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 259.McClain JA, et al. , Adolescent binge alcohol exposure induces long-lasting partial activation of microglia. Brain Behav Immun, 2011. 25 Suppl 1: p. S120–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 260.Roberson R, et al. , Neuroprotective fractalkine in fetal alcohol syndrome. Am J Obstet Gynecol, 2011. 204(5): p. 400 e1–3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 261.Dennis CV, et al. , Microglial proliferation in the brain of chronic alcoholics with hepatic encephalopathy. Metab Brain Dis, 2014. 29(4): p. 1027–39. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES