Skip to main content

REVIEW article

Front. Neurosci., 17 August 2018
Sec. Neuropharmacology
This article is part of the Research Topic Involvement of Inflammation in the Pathophysiology and Treatment of Mood Disorders View all 5 articles

Involvement of Innate and Adaptive Immune Systems Alterations in the Pathophysiology and Treatment of Depression

\r\nEva M. Medina-RodriguezEva M. Medina-Rodriguez1Jeffrey A. LowellJeffrey A. Lowell1Ryan J. WorthenRyan J. Worthen1Shariful A. SyedShariful A. Syed1Elonore Beurel,*Eléonore Beurel1,2*
  • 1Department of Psychiatry and Behavioral Sciences, Miller School of Medicine, University of Miami, Miami, FL, United States
  • 2Department of Biochemistry and Molecular Biology, Miller School of Medicine, University of Miami, Miami, FL, United States

Major depressive disorder (MDD) is a prevalent and debilitating disorder, often fatal. Treatment options are few and often do not provide immediate relief to the patients. The increasing involvement of inflammation in the pathology of MDD has provided new potential therapeutic avenues. Cytokine levels are elevated in the blood and cerebrospinal fluid of MDD patients whereas immune cells often exhibit an immunosuppressed phenotype in MDD patients. Blocking cytokine actions in patients exhibiting MDD show some antidepressant efficacy. However, the role of cytokines, and the immune response in MDD patients remain to be determined. We reviewed here the roles of the innate and adaptive immune systems in MDD, as well as potential mechanisms whereby the immune response might be regulated in MDD.

Introduction

Once considered primarily as a component of the body’s responses to injury and infection, inflammation is now also known to be an important component of neuronal function and dysfunction. Here we consider recent evidence of these roles of inflammation and its potential as a therapeutic target in major depressive disorder (MDD). MDD is a heterogeneous, recurrent, and debilitating disease that has a high prevalence worldwide, has unknown etiology, and is inadequately treated in approximately two-thirds of patients (Rush et al., 2006). During the last two to three decades inflammation has emerged as an important factor contributing to MDD (Miller and Raison, 2016) and evidence is emerging that it may provide a new target for therapeutic intervention.

Immune System and Depression

The Innate and Adaptive Immune Systems

The immune system is a complex network of many types of immune cells and secreted factors that contribute to regulating inflammation in all tissues of the body. The immune system constantly surveys its environment to ensure health and is activated following infection for pathogen clearance, following tissue damage to support repair and by psychological stress.

The immune system is commonly characterized as consisting of two arms: the innate immune system and the adaptive immune system, even though both the innate and adaptive immune systems communicate and cooperate with each other to direct pro- and anti-inflammatory responses with the main goals of eradicating pathogens and restoring homeostasis. Cytokines and chemokines, which are produced during innate and adaptive immune responses, are thought to mediate some of these effects, particularly by mobilizing and modulating the function of immune cells. The innate immune system composed of antimicrobial peptides and myeloid lineage cells (e.g., monocytes/macrophages, dendritic cells, neutrophils, and others) is the first line of response to pathogens or insults. Innate immune cells are sentinels that constantly survey the environment for non-self signals, using pattern-recognition receptors [e.g., Toll-like receptors (TLRs)] that recognize common microbial features (virus, bacteria, fungi, parasites) or other damage signals. If the innate immune system fails to resolve the insult, a second line of defense by the adaptive immune system is activated. While the innate immune response occurs quickly, the adaptive immune response, composed of lymphocytes and secreted antibodies, relies on the proliferation and differentiation of antigen-specific T cells and B cells that requires time to be mounted (den Haan et al., 2014). The advantage of the adaptive immune response is the selectivity for antigen, since the T cell-receptor and B cell-receptor systems provide a pathogen-tailored immune response. In addition, the adaptive immune system is thought to be responsible for immunological memory which is critical in the case of re-exposure to the same antigen.

It is important to note that within each organ, there are also often specialized immune cells, such as microglia in the central nervous system (CNS). Microglial cells derive from a myeloid lineage in early development when they colonize the neural tube. It was thought for a long time that microglial cells were the only immune competent cells to mediate immune responses in the CNS because the CNS was thought to be isolated from the immune system by the presence of the blood–brain barrier (BBB) that prevents the free entry of immune cells into the CNS. However, recent evidence shows that immune cells can infiltrate the CNS (Korn and Kallies, 2017). In addition, stress, the most common factor that precipitates depressive episodes, has been shown to induce inflammatory responses both in the brain and in the periphery by activating similar pathways (e.g., TLRs) that are activated by pathogens on innate immune cells. Here, we review the role of both the innate and adaptive systems in MDD and their potential contributions to disease.

Roles of the Innate Immune System in Mouse Models of Depression and in MDD Patients

A diverse array of evidence has been reported implicating the innate immune system involvement in the pathology of MDD (Hodes et al., 2015). It has been known for decades that MDD patients often have elevated levels of circulating granulocytes and monocytes (Maes et al., 1992; Maes, 1995; Seidel et al., 1996b), pro-inflammatory cytokines and their receptors, acute phase proteins, chemokines, and prostaglandins (Miller et al., 2009), indicative of activation of the peripheral immune system in MDD. Similarly, an increased number of circulating monocytes has been observed in mice exhibiting depressive-like behaviors (Engler et al., 2004) and neutralization of monocytes is sufficient to reduce depressive-like behaviors (Zheng et al., 2016). While examining the mechanisms whereby monocytes promote depressive-like behaviors, it has been proposed that this is due to secretion by monocytes of cytokines that are capable of promoting depressive-like behaviors (Miller et al., 2009). Furthermore, activated peripheral monocytes traffic to the brain vasculature and parenchyma during the induction of depressive-like behavior, triggering inflammation in the CNS and therefore contributing to the pathophysiology of depression (Wohleb et al., 2013). Monocytes infiltrate and colonize brain regions associated with depression and anxiety in mice exposed to social defeat stress (Sawicki et al., 2015). Conversely, depressive-like behavior induced following peripheral organ inflammation is reversed by blocking the brain infiltration of monocytes (D’Mello et al., 2009). This points out the importance of the influx of peripheral innate immune cells such as monocytes into the brain to maintain depressive-like behaviors. Among the mechanisms proposed for the increase of brain monocytes after stress is increased release of bone marrow derived monocytes to the bloodstream after stimulation by catecholamines produced by the sympathetic nervous system after stress (Wohleb et al., 2011, 2013; Heidt et al., 2014). Once released, monocytes are recruited to the brain upon receiving signals secreted by activated microglia, including two key chemokine receptors, chemokine receptor-2 (CCR2) and fractalkine receptor (CX3CR1) (D’Mello et al., 2009; Wohleb et al., 2013). Once in the brain, monocytes recruited by microglial cells are thought to produce interleukin (IL)-1β, which promoted anxiogenesis through the activation of neurovascular IL-1R1 (McKim et al., 2017). Consistent with the presence of monocytes in the brain of mice exhibiting depressive-like behaviors, post-mortem analyses showed increased numbers of perivascular tissue resident macrophages associated with elevated levels of the C–C chemokine ligand 2 (CCL2) in the brain of MDD patients (Torres-Platas et al., 2014). These and other studies indicate that peripheral innate immune cells are activated and reach the brain in depression and animal models of depression.

Role of Astrocytes and Microglia in Mouse Models of Depression and in MDD Patients

Both astrocytes and microglia have been implicated as contributing to the pathophysiology of depression. MDD is associated with loss and hypotrophy of astrocytes (Davis et al., 2002; Miguel-Hidalgo et al., 2010). This mainly occurs in the frontolimbic systems that are relevant for major depression (Ongur et al., 1998; Rajkowska et al., 1999; Miguel-Hidalgo et al., 2000; Cotter et al., 2002; Altshuler et al., 2010; Gittins and Harrison, 2011; Rajkowska and Stockmeier, 2013; Torres-Platas et al., 2014; Nagy et al., 2015; Rial et al., 2015; Cobb et al., 2016; Medina et al., 2016; Rubinow et al., 2016). Interestingly, there is no astrogliosis in MDD patients, which contrasts with the prominent astrogliosis that commonly occurs in patients with neurodegenerative diseases. Loss and shrinkage of astrocytes also have often been reported in mouse models of depression (Czeh et al., 2006; Rajkowska and Stockmeier, 2013; Sanacora and Banasr, 2013). That reduced astrocyte functions may contribute to depression is supported by the finding that elimination of astrocytes from the rat prefrontal cortex (PFC) is sufficient to induce depressive-like behaviors (Banasr and Duman, 2008). Antidepressant treatments reverse these stress-induced astrocytic morphological changes to restore astrocytic function, as well as reverse associated depressive-like behaviors (Rajkowska and Stockmeier, 2013; Rial et al., 2015). These findings support the current theory that astrocytic dysfunction contributes to the pathophysiology of MDD and that the cellular actions of antidepressants may correct or compensate for impaired function of astrocytes (Manji et al., 2003; Czeh et al., 2006).

Promotion of depression by impaired astrocyte function may be due to consequential neuronal damage, including induction of neuronal excitotoxicity, because of the roles of astrocytes in supporting neuronal function by the uptake of synaptic glutamate or the production of growth factors (Sofroniew and Vinters, 2010). Consistent with this, astrocyte density (Miguel-Hidalgo et al., 2000) and synapse-related genes (Kang et al., 2012) are both decreased in the dorsolateral PFC of MDD patients. In addition, knockout of the astrocytic glutamate transporter GLT-1, which is important for the reuptake of extracellular glutamate, in the lateral habenula of mice, is sufficient to induce depressive-like behaviors (Cui et al., 2014). Altogether, these findings suggest that loss of astrocytes might promote depression by impairing neuronal functions.

Many other functions of astrocytes are altered in depression, including formation of gap junctions, neurotrophic support, energy metabolism, water homeostasis, γ-aminobutyric acid (GABA) and monoamine recycling, and gliogenesis, and antidepressant treatments tend to reverse most of these alterations (Tavares et al., 2002; Simard and Nedergaard, 2004; Czeh et al., 2006; Haydon and Carmignoto, 2006; Hundal, 2007; Rouach et al., 2008; Eulenburg and Gomeza, 2010; Sofroniew and Vinters, 2010; Cao et al., 2013; Jiang and Salton, 2013; Martin et al., 2013; Rajkowska and Stockmeier, 2013; Rajkowska et al., 2013; Nagy et al., 2015; Rial et al., 2015; Bjorkholm and Monteggia, 2016). Therefore, it remains to be determined by which mechanism(s) astrocytic atrophy promotes depression. In addition, it is unclear what causes astrocytic atrophy in MDD patients and whether this atrophy results from an inflammatory process in the brain mediated by immune cells.

Microglial cells have been shown to be activated in various brain regions of MDD patients, including in the dorsolateral PFC, anterior cingulate cortex, mediodorsal thalamus, insula and hippocampus, both in post-mortem tissue and by positron emission tomography (PET) imaging (Steiner et al., 2008; Schnieder et al., 2014; Setiawan et al., 2015). Activation of microglia often is the consequence of increased inflammation in the brain (Dantzer et al., 2008; Rivest, 2009; Benarroch, 2013; Slavich and Irwin, 2014; Delpech et al., 2015; Jo et al., 2015; Ransohoff et al., 2015; Menard et al., 2016; Miller and Raison, 2016; Wachholz et al., 2016; Haroon et al., 2017). Characterization of microglia in mouse models of depression showed that microglial cells adopt an activated phenotype expressing surface markers such as major histocompatibility complex (MHC) class II, CD86 or CD54 (Wachholz et al., 2016) and producing pro-inflammatory cytokines [e.g., IL-1β, IL-6, tumor necrosis factor α (TNFα)] (Chabry et al., 2015), toxic molecules (e.g., nitric oxide) or extracellular vesicles, such as exosomes, which might be responsible for propagating inflammatory signals throughout the brain (Fruhbeis et al., 2013). Microglial cells, once activated, are often eliminated (Kreisel et al., 2014). In a model of unpredictable chronic stress, it was shown that after a first phase of microglial activation, microglial cells decrease in number in the hippocampus of stressed mice, whereas the blockade of microglia activation was sufficient to prevent the reduction in microglia number, and the induction of unpredictable chronic stress-associated depressive-like behaviors (Kreisel et al., 2014). In contrast, neuroprotective microglia diminish in the hippocampus (Branchi et al., 2014), PFC (Hinwood et al., 2012), and amygdala (Hamidi et al., 2004) after chronic stress.

Microglial cells express several inflammatory pathways that are responsible for cytokine production, including TLR pathways. Microglial cells are the main producers of cytokines in the brain after exposure to stress. In response to stress, microglial cells are activated by molecules called alarmins, danger-associated molecular pattern molecules, or damage-associated molecular pattern molecules. For example, the protein high mobility group box-1 (HMGB1) is up-regulated and it activates the TLR4 and receptor for advanced glycation end-products (RAGE) pathways resulting in increased production of proinflammatory cytokines (Weber et al., 2015; Franklin et al., 2018). Central administration of HMGB-1 also induces depressive-like behaviors (Wu et al., 2015; Cheng et al., 2016; Franklin et al., 2018), altogether suggesting that up-regulation of alarmins such as HMGB1 contribute to the development of depressive-like symptoms by activating microglial cytokine production. Additionally, cytokines, such as TNFα, also can disrupt BBB permeability, raising the possibility that increased BBB permeability caused by microglia-derived cytokines contributes to depression (Menard et al., 2017; Cheng et al., 2018) by promoting the infiltration of peripheral inflammatory molecules and immune cells (Wohleb et al., 2011, 2013).

Microglial cells contribute to synaptic pruning by phagocytosis of synapses during development but also possibly during adulthood (Kettenmann et al., 2013; Miyamoto et al., 2013; Wake et al., 2013; Wohleb et al., 2018). Therefore microglia can modulate neuronal circuits (Wake et al., 2013), and these processes can increase the production of neurotrophic factors in response to neuronal stimulus (Nakajima et al., 2007). Therefore, microglia activation has direct effects on neuronal function. Microglia activation impedes microglia-neuron communication via the CX3CR1-fractaline pathway after chronic stress (Corona et al., 2010; Milior et al., 2016). Furthermore, stress-induced anxiety- and depressive-like behaviors are associated with increases in neuronal colony stimulating factor 1 (CSF1) which activates microglial phagocytosis of synaptic elements in the medial PFC. This demonstrates the importance of reshaping neuronal connectivity by the neuronal-activated microglia communication in promoting depressive-like behaviors (Wohleb et al., 2018). Moreover, microglial activation after stress-induced depressive-like behavior induces the activation of the kynurenic acid pathway (Jo et al., 2015), leading to the depletion of monoamines and the production of quinolinic acid (Steiner et al., 2011), and ultimately to the dysregulation of monoaminergic and glutamatergic circuits (Jo et al., 2015). Consistent with the fact that microglial cells are particularly enriched in the hippocampus, a brain region relevant for stress response and neuroplasticity, the role of microglia appears to be critical for maintaining neuronal homeostasis and neurotransmission, whereas activation of microglia by stress alters this equilibrium, which can promote susceptibility to depressive-like behaviors.

Role of the Adaptive Immune System (T Cells and B Cells) in Mouse Models of Depression and in MDD Patients

In addition to increased peripheral monocytes, atrophy of astrocytes, and increased activation of microglia, MDD patients also exhibit decreased numbers of circulating T cells and a reduction in regulatory B cells (Ahmetspahic et al., 2018), suggestive of a dysregulation of the adaptive immune system in MDD patients. In addition, T cells from MDD patients have a reduced capacity to respond to stimulation, indicative of an immunosuppressed phenotype of T cells in MDD (Herbert and Cohen, 1993; Zorrilla et al., 2001; Irwin and Miller, 2007). Besides a global immunosuppression of T cells in MDD, there are also alterations of the T helper (Th) (CD4+) and T cytotoxic (CD8+) cells in MDD patients. Both Th (CD4+) and T cytotoxic (CD8+) cells have altered activities in MDD patients exemplified by an increase of the ratio of CD4+ T cells relative to CD8+ T cells, suggestive of an increase of CD4+ cells and/or a decrease of CD8+ cells favoring the Th cells. Differentiation of the different subsets of CD4+ Th effector cells, Th1, Th2, Th17, and regulatory T cells (Tregs) require specific cytokines, antigen presentation, and co-stimulatory signals for the production of each subtype. Evaluation of the differentiation status revealed a decrease in circulating Tregs (reviewed in Li et al., 2010; Martino et al., 2012; Toben and Baune, 2015) and an increase in the Th1/Th2 ratio (Myint et al., 2005), whereas circulating Th17 cells are increased in MDD patients (Chen et al., 2011). These findings indicate that there is a massive remodeling of the Th compartment in MDD patients that might be dictated by the cytokine milieu present in MDD patients. The exacerbation of the Th2 responses in MDD or asthma patients might contribute to the higher rate of both MDD in asthma patients (Van Lieshout et al., 2009), and asthma in MDD patients (Shen et al., 2017). However, further studies are required to identify the underlying mechanisms. Mice exhibiting depressive-like behaviors also show increased levels of brain Th1 and Th17 cells (Beurel et al., 2013). Although the contribution of these various Th cells to MDD pathology remains to be determined, it is worth noting that administration of Th17 cells is sufficient to promote depressive-like behaviors in mice (Beurel et al., 2013). Recent findings revealed that the Th17 cells that accumulate in the brains of mice exhibiting depressive-like behavior express characteristics of both pathogenic Th17 cells and of follicular T cells (Beurel et al., 2018). To further corroborate the notion that T cell alterations contribute to MDD, antidepressant treatments have been shown to restore the Th1/Th2 imbalance in MDD patients (Martino et al., 2012), which may be mediated by increasing Treg levels (Himmerich et al., 2010). Together these findings suggest that T cell alterations are present in MDD patients and might contribute to the development and maintenance of MDD.

The role of T cells in modulating mood and behaviors has mainly been studied in the context of examining the effects of the absence of T cells using lymphopenic mice devoid of both T cells and B cells. T cells have been shown to be beneficial to brain function through neuroprotective roles (Miller, 2010; Filiano et al., 2017), providing continual brain immune surveillance (Lewitus et al., 2008), and promoting neurogenesis (Ziv and Schwartz, 2008), cognition (Schwartz and Kipnis, 2011), and mood (Herkenham and Kigar, 2017). Thus, CNS-targeted T cells promote brain health and function, supporting the notion of “protective autoimmunity” (Rook and Lowry, 2008; Lewitus and Schwartz, 2009; Lewitus et al., 2009). Similarly, immunization of mice with a modified CNS antigen to generate weak CNS-autoreactive T cells protected mice from the development of depressive-like behavior following stress (Cohen et al., 2006; Lewitus et al., 2009). However, T cells may not require CNS-specificity to be beneficial to the CNS, as adoptive transfer of lymphocytes to lymphopenic mice rescues sociability deficits displayed by lymphopenic mice (Filiano et al., 2016) and confers antidepressant-like effects (Clark et al., 2016). These results suggest that T cells contribute to normal affective behavior in mice by supporting an adaptive response to stress (Herkenham and Kigar, 2017). Furthermore, administration of splenocytes from socially defeated mice, but not from unstressed mice, protects lymphopenic recipient mice from the effects of social defeat and confers both anxiolytic and antidepressant-like effects (Brachman et al., 2015; Scheinert et al., 2016), suggesting that stress may produce changes in T cells that promote resilience to future stress. These beneficial effects of T cells contrast with the deleterious effects of T cell subsets in autoimmune diseases, where subsets of T cells drive pathogenicity. It is possible that in depression, similar deleterious roles of T cell subsets exist, whereas most of the studies have been focused on the overall importance of T cells in mediating depressive-like behaviors using lymphopenic mice.

Cytokines In Depression and Pathways Controlling Inflammation In Depression

Overview of Meta-Analyses of Cytokine Levels in Depressed Patients Compared to Healthy Controls

Major depressive disorder (MDD) has been associated with alterations of both the innate and the adaptive immune responses (Dantzer, 2009) and MDD patients often experience elevated levels of cytokines (Felger and Lotrich, 2013). Although it is difficult to find a consensus in the many different cytokines that have been reported to be altered in MDD patients in various studies, the levels in both blood and cerebrospinal fluid (CSF) of TNFα, IL-6 and IL-1β have very often been found to be significantly elevated in individuals with MDD compared to healthy controls (Maes et al., 1997; Levine et al., 1999; Tuglu et al., 2003; O’Brien et al., 2004; Howren et al., 2009; Lindqvist et al., 2009; Dowlati et al., 2010; Hiles et al., 2012; Haapakoski et al., 2015; Kohler et al., 2017). In addition to these three pro-inflammatory cytokines, elevated levels of several anti-inflammatory cytokines [IL-2, IL-4, IL-10, transforming growth factor β (TGFβ)] have also been associated with MDD (Kubera et al., 2001b; Hernández et al., 2008; Sutcigil et al., 2008; Fazzino et al., 2009), whereas other cytokines, such as sIL-2R, CCL-2, IL-13, IL-18, and IL-12, have been found to be elevated in the blood of MDD patients in a few studies (Maes, 1995; Kim et al., 2002; Merendino et al., 2002; Pavón et al., 2006; Köhler et al., 2017). In addition, the acute phase protein, C-reactive protein (CRP), which is an innate immune marker reflecting peripheral inflammation, correlates with central inflammation, and is associated with depressive symptom severity (Felger et al., 2018). Overall, these findings suggest a general induction of cytokine production in MDD patients. However, not all cytokines have been shown to be elevated in MDD. For example, interferon (IFN)γ was often found to be lower in MDD patients compared to healthy controls (Hernández et al., 2008; Fornaro et al., 2013; Ho et al., 2015). In addition to being elevated in the blood and CSF of MDD patients, there is also evidence that cytokines elevated in the brains of MDD patients. Post-mortem measurements of brain tissue samples from Broadman area 10 (frontal cortex) showed significant increases of IL-1α, IL-2, IL-10, and IFNγ mRNA levels in suicide completers (Shelton et al., 2011).

The presence of elevated levels of cytokines in MDD patients raises the question of whether inflammation causes depression or is a consequence of the disease. The literature suggests a bidirectional relationship between depression and inflammation: inflammation is capable of inducing depressive-like symptomatology and depressive-like behavior is associated with elevated cytokine levels (Dantzer, 2009). Thus, for example, elevation of certain cytokines has been associated with the severity of MDD (Lindqvist et al., 2009) and in a prospective study of geriatric patients with no psychiatric history, the elevation of cytokines preceded the onset of depressive symptoms (Baune et al., 2012). In the Avon Longitudinal Study of Parents and Children birth cohort, higher serum concentration of IL-6 at 9 years old predicted the likelihood of developing MDD symptoms at 18 years old (Khandaker et al., 2014) and another longitudinal study found a similar association (Gimeno et al., 2009). Consistent with the notion that a low grade inflammation might increase susceptibility to MDD, healthy volunteers receiving a low dose of lipopolysaccharide (LPS), which triggers cytokine production, exhibit depressive symptoms whose severity correlated with cytokine production (Reichenberg et al., 2001). Furthermore, IFNα infusion therapy is associated with 30–50% of patients developing depression, in a dose-dependent manner (Capuron et al., 2005). Moreover, many inflammatory diseases, such as multiple sclerosis, are co-morbid with depression with an estimation of 60–80% of multiple sclerosis patients exhibiting depression (Caine and Schwid, 2002; Pucak et al., 2007). Other diseases with an inflammatory component such as diabetes mellitus (Anderson et al., 2001) and coronary artery disease (Rudisch and Nemeroff, 2003) have been shown to have a high prevalence of comorbidities with depression. Similar findings were reported in mice where higher inflammation is associated with depressive-like behaviors (Henry et al., 2008; O’Connor et al., 2009). Additionally, psychological stress (Maes et al., 1998) that often triggers depression is sufficient to induce pro-inflammatory responses. Taken together, depression is accompanied by significant changes of cytokine production which indicate activation of the inflammatory response in MDD patients, and inflammation appears to promote susceptibility to depression as well as to be induced by depression.

Mechanisms Regulated by NF-κB and GSK3 Contributing to Inflammation in Depression

Although changes in cytokine production in MDD are well-established, the mechanisms whereby these cytokines are produced in MDD patients remain largely unknown. Multiple mechanisms likely contribute to inflammation in depression. Mechanistic studies in rodents have shown the importance of a few key pathways in controlling cytokines production. Among these, here we will focus on nuclear factor-κB (NF-κB) and glycogen synthase kinase 3 (GSK3) pathways, which are predominant molecular pathways regulated by stress.

TLR4/GSK3/NF-κB

One of the major pathways regulating the production of most cytokines after infection, injury or stress is the TLR4 pathway. TLR4 is a pattern-recognition receptor that leads to activation of the transcription factor NF-κB to increase the production of many cytokines. TLR4 is activated following recognition of microbial peptides (such as LPS) after infection or endogenous molecules (HMGB1, heat shock proteins, ATP) that are released after cellular damage or psychological stress. Although TLR4s are most well-recognized as being expressed by innate immune system cells, they are also expressed by neurons and glia (Slavich and Irwin, 2014). Altogether 10 functional human TLRs have been identified (12 in mice) and among these, TLR4 has been most repeatedly linked with the induction of inflammation following psychological stressors (Jope et al., 2017). TLR4 activation up-regulates its own expression, and elevations of TLR4 mRNA and protein have been detected in both the periphery and CNS of MDD patients (Hung et al., 2014). Furthermore, successful treatment of MDD restores levels of TLR4, pointing toward a role for TLR4 in MDD patients (Raison and Miller, 2017). In addition, TLR4 ligands (e.g., HMGB1) are induced by stress in animal models of depression (Fleshner, 2013). Stress has also been shown to promote NF-κB activation, which likely results at least in part from activation of TLR4 (Cheng et al., 2016). Another important link in the TLR4 pathway is GSK3. Indeed, TLR4-mediated cytokine production is highly dependent on active GSK3 (Martin et al., 2005). GSK3 promotes NF-κB activation and NF-κB-dependent production of proinflammatory cytokines, whereas GSK3 inhibits cAMP response element binding (CREB)-dependent production of anti-inflammatory cytokines (Martin et al., 2005), so GSK3 activation, which occurs after many types of stress, favors proinflammatory responses. Furthermore, GSK3 is activated in post-mortem brains from MDD patients and in mice exhibiting depressive-like behaviors (Beurel, 2011; Beurel et al., 2015), and most antidepressant drugs inhibit GSK3 (Beurel et al., 2015). In addition, GSK3 promote IFNγ-dependent indoleamine 2,3-dioxygenase (IDO) expression by modulating the Janus kinase (JAK)/signal transducer and activator of transcription (STAT) pathway in dendritic cells (Noh et al., 2015). This might also be mediated by NF-κB because non-canonical NF-κB signaling is required to induce IDO in dendritic cells by modulating the JAK/STAT pathway (Tas et al., 2007). Altogether this suggests another potential role of both GSK3 and NF-κB in promoting depressive-like behavior by modulating IDO. Therefore, GSK3 is an important hub linking stress, inflammation, and depression via the TLR4–GSK3–NF-κB signaling axis.

Inflammasome

The inflammasome pathway is selectively responsible for the production of IL-1β and IL-18 and is part of the innate immune response that contributes to cytokine production. It is a cytosolic complex, composed of Nod-like receptor (NLR), caspase-1, and apoptosis-associated speck-like protein containing C-terminal caspase recruitment domain (ASC)-1. Once activated, this complex leads to the caspase-1-mediated cleavage of pro-IL1β or pro-IL-18 to produce active IL-1β and IL-18 (Fleshner et al., 2017). Among the NLR subtypes, the NLR protein 3 (NLRP3) inflammasome has often been studied in the context of CNS inflammation related to stress and depression. The NLRP3 inflammasome requires a two-step process to produce cytokines, consisting first of the assembly of a primed inflammasome with procaspase-1 followed by activation of the inflammasome involving cleavage, and thereby activation, of procaspase-1. Thus, HMGB1, for example, induces the assembly of the NLRP3 inflammasome by activating TLR4 (Fleshner et al., 2017). The NLRP3 inflammasome is then activated by another or the same signal (e.g., HMGB1, ATP, heat-shock proteins) (Miller and Raison, 2016; Fleshner et al., 2017). NLRP3- and caspase-1-deficient mice are resilient to depressive-like behavior, indicating that inflammation caused by activation of the NLRP3 inflammasome regulates susceptibility to depression-like behaviors (Alcocer-Gomez et al., 2016; Iwata et al., 2016; Wong et al., 2016). In contrast, antidepressant drugs prevent inflammasome activation (Alcocer-Gomez et al., 2017). In addition, MDD patients have increased expression of NLRP3 and caspase-1 in circulating immune cells, suggesting that the NLRP3 inflammasome is activated in MDD patients, which correlates with increased levels of IL-1β and IL-18 in MDD patients (Alcocer-Gomez et al., 2014). Interestingly, the inflammasome assembly is controlled by both NF-κB and GSK3 (Guo et al., 2015; Jope et al., 2017). Thus, the inflammasome is primed by a NF-κB-dependent signal (Guo et al., 2015). Moreover, in lupus nephritis, GSK3 activates NLRP3 and therefore, IL-1β production, and psychological stress activating GSK3 induces the NLRP3/IL-1β pathway (Liu et al., 2015; Cheng et al., 2016). It remains to be determined if GSK3 actions on the inflammasome are NF-κB dependent.

HPA Axis

The hypothalamus–pituitary–adrenal (HPA) axis activation is the main response to stress. It is a complex interaction between the paraventricular nuclei of the hypothalamus, the pituitary gland, and the adrenal glands, a major neuroendocrine system that controls many functions, such as digestion, immune responses, and mood and emotions. In response to stress, corticotropin-releasing factor (CRF) is released, which initiates a cascade of events comprising the release of adrenocorticotropic hormone (ACTH) leading to the production of glucocorticoids (cortisol in humans, corticosterone in rodents) from the adrenal glands (Herkenham and Kigar, 2017). The HPA axis is finely tuned in healthy individuals via negative-feedback since its activation leads to profound metabolic, physiologic, and immunologic changes across nearly all body systems, including the brain (Sapolsky et al., 2000; Herbert et al., 2006; Silverman and Sternberg, 2012). Thus, glucocorticoids regulate the magnitude and duration of the glucocorticoids release. In addition, the HPA axis receives feedback from afferent projections from various brain regions, such as brain stem noradrenergic neurons. Glucocorticoids bind to the glucocorticoid receptor (GR), which is part of a multiprotein complex consisting of several heat shock proteins. Once activated upon binding of its ligand, the GR dissociates from the heat shock protein complex to migrate to the nucleus where it binds to DNA binding elements or interacts with other transcription factors, such as NF-κB. Two-thirds of MDD patients have elevated levels of circulating glucocorticoids (Stetler and Miller, 2011), and genetic studies have identified polymorphisms and epigenetic changes in the GR and associated proteins (Pariante and Lightman, 2008; Miller and Raison, 2016). These modifications lead to a desensitization to cortisol (Herkenham and Kigar, 2017). Yet, one of the functions of glucocorticoids is to inhibit the peripheral immune response by inducing amongst other actions, immune cell apoptosis and suppression of NF-κB signaling, which is the main pathway responsible for the production of cytokines (Slavich and Irwin, 2014). These findings produce a seemingly paradoxical situation in which depression is characterized by high levels of both cortisol and proinflammatory factors. In animal models of glucocorticoid insensitivity, immune cells become desensitized to the immunosuppressing effects of glucocorticoids (Silverman and Sternberg, 2012), leading to higher levels of proinflammatory cytokines which further inhibit GR function (Slavich and Irwin, 2014; Miller and Raison, 2016). In addition, in rodents, GSK3, which is activated during depressive-like behaviors and which promotes proinflammatory cytokine production, also phosphorylates the GR, inhibiting its transcriptional activity (Rogatsky et al., 1998). Furthermore, GSK3 knockin mice, which have constitutively active GSK3, also exhibit increased levels of glucocorticoids after exposure to stress (Polter et al., 2010). GSK3 is degraded upon glucocorticoid signaling activation, leading to the formation of adherens junctions and increased tight junction strength (Failor et al., 2007), which might favor the immune suppressive actions of the glucocorticoids. Altogether, glucocorticoids are important immunoregulators.

Vagus Nerve

The vagus nerve is the tenth cranial nerve and is a complex nerve that innervates multiple organs through afferent and efferent fibers throughout the body (Carreno and Frazer, 2017). One major hub of vagal afferent termination is the nucleus tractus solitarius of the medulla which directly and indirectly innervates several regions of the brain associated with depression (e.g., limbic forebrain) and inhibits the HPA axis, reducing cortisol secretion (Ondicova et al., 2010). The vagus nerve functions as a brake slowing the heart, inhibiting the sympathetic and adrenal activity and ultimately reducing stress reactivity. The vagus nerve has been implicated in immune function. Thus, subdiaphragmatic vagal afferents provide information to the brain regarding visceral state. During inflammation with elevation of proinflammatry cytokines (IL-6, TNFα, IL-1β), the vagus nerve mediates in part the induction of sickness behavior (Dantzer et al., 2008), through immune cells (e.g., macrophages and dendritic cells) present in the perineural sheath of the vagus nerve (Dantzer, 2009) that relay the inflammatory signal to the brain. Consistent with this, cytokines elevated by LPS administration activate brain regions downstream of vagal afferent input (e.g., brainstem, hypothalamus, and limbic structures) and sectioning of the vagus nerve abolishes LPS-induced sickness behavior despite an elevation of cytokines (Pavlov and Tracey, 2012). This suggests that afferent terminals of the vagus nerve sense peripheral inflammation and affects depression-associated brain regions that may contribute to behavioral and neuroendocrine changes observed in MDD patients. Reciprocally, stimulation of the vagus nerve is associated with reduced production of proinflammatory cytokines (IL-6, TNFα, and others) (Bernik et al., 2002), which has been named the “inflammatory reflex.” The latter is mediated by cholinergic signaling and regulated by the afferent fibers of the vagus nerve (Pavlov and Tracey, 2012). Cholinergic signaling through α7-nicotinic acetylcholine receptors expressed on immune cells inhibits NF-κB-dependent proinflammatory cytokine (e.g., IL-1β and TNFα) secretion by macrophages, T cells, and dendritic cells (Pavlov and Tracey, 2012). These effects may contribute to the efficacy of chronic vagal nerve stimulation as a treatment in patients with treatment-resistant depression (Carreno and Frazer, 2017). Overall, the vagus nerve serves the brain-immune axis as a centrally controlled inflammation rheostat. GSK3 has been shown to mediate parasympathetic dysfunction in a diabetic mice model (Zhang et al., 2014), and to promote neuropathic pain, whereas GSK3 inhibition ameliorates neuropathic pain by reducing proinflammatory cytokines, and increasing serotonergic and catecholaminergic pathways (Mazzardo-Martins et al., 2012).

Effects of Inflammation on Brain Functions

Neuroinflammation is a condition of elevated levels of pro-inflammatory cytokines in the CNS, and is induced in response to psychosocial stress as well as infection or injury (Rivest, 2009; Shatz, 2009; Benarroch, 2013; Le Thuc et al., 2015; Ransohoff et al., 2015). Acute neuroinflammation is often beneficial and neuroprotective following infection or injury of the CNS (Bitzer-Quintero and Gonzalez-Burgos, 2012; Le Thuc et al., 2015). For example, following injury peripheral immune cells, such as monocytes and lymphocytes (Stoll et al., 2002; Whitney et al., 2009), and microglia (Bitzer-Quintero and Gonzalez-Burgos, 2012), are recruited to the site of insult by chemokines (Ransohoff and Engelhardt, 2012). Once recruited, the cells act to resolve the insult in part by phagocytosing cellular debris caused by damage, and releasing anti-inflammatory cytokines and neurotrophic factors, promoting neurogenesis and synaptogenesis, overall fostering recovery after neuronal damage (Mathieu et al., 2010; Bitzer-Quintero and Gonzalez-Burgos, 2012; Le Thuc et al., 2015). However, excessive or prolonged neuroinflammation is detrimental to CNS functions, impairing learning and memory, inhibiting long-term potentiation, decreasing neurogenesis, and altering dendritic spine density (Bilbo and Schwarz, 2009), synaptic plasticity, and synaptic scaling (Jakubs et al., 2008; McAfoose and Baune, 2009; Wood et al., 2011; Chugh et al., 2013; Kohman and Rhodes, 2013; Green and Nolan, 2014; Vezzani and Viviani, 2015). Additionally, chronic neuroinflammation leads to decreased production of brain-derived neurotrophic factor (BDNF) by neurons and astrocytes which may impair neurogenesis in the dentate gyrus of the hippocampus (Sapolsky, 2001; Jo et al., 2015) and alter neuronal functional integration (Jakubs et al., 2008; Wood et al., 2011). Neuroinflammation increases the production of neurotoxic factors such as reactive oxygen species (Whitney et al., 2009) and glutamate leading to excitotoxicity, increases BBB permeability (Pan et al., 2011) allowing inappropriate peripheral immune cell infiltration into the CNS (Lossinsky and Shivers, 2004), and perpetuates further production of pro-inflammatory cytokines (Le Thuc et al., 2015). Altogether, these processes cause damage to neurons, impairing neurogenesis and healthy synaptic connectivity in both the adult (Das and Basu, 2008) and developing brain (Harry and Kraft, 2012). Overall, excessive or prolonged neuronflammation is associated with pathogenicity in multiple neurological diseases, including mood disorders (Borsini et al., 2015).

Interactions Between Antidepressant Actions and the Immune Response Of Depression

Effects of Antidepressants on Cytokine Levels in Depressed Patients and in Mouse Models of Depression

Major depressive disorder is associated with inflammation, and antidepressant treatments are generally thought to shift the balance toward anti-inflammatory responses (Sluzewska et al., 1997; Maes, 1999; Lanquillon et al., 2000; Kubera et al., 2001a) leading to an overall normalization of cytokine levels to healthy control levels (Maes et al., 1997; Tuglu et al., 2003; O’Brien et al., 2004; Howren et al., 2009; Hannestad et al., 2011). However, the role of antidepressant medications in regulating cytokine production differs according to the class of antidepressant used. Thus, selective serotonin reuptake inhibitors (SSRIs) seem to reduce IL-1β, IL-6, and TNFα, whereas serotonin and norepinephrine reuptake inhibitors (SNRIs) can have proinflammatory effects, increasing in particular TNFα and IL-6 levels (Piletz et al., 2009) consistent with the proinflammatory effects of norepinephrine on innate immune cells (Thayer and Sternberg, 2010). Moreover, sertraline (SSRI) therapy decreases IL-12 levels but increases IL-4 and TGFβ levels in responders (Sutcigil et al., 2008). Amitriptyline [tricyclic antidepressant (TCA)] therapy in MDD patients is associated with decreased TNFα (Hinze-Selch et al., 2000), whereas desipramine, increases IL-10 in MDD patients but does not affect pro-inflammatory cytokines (Roque et al., 2009). In addition to decrease proinflammatory cytokine levels, acute antidepressant treatment can increase anti-inflammatory cytokine levels (e.g., IL-1ra, IL-10, TGFβ) (Seidel et al., 1995, 1996a; Maes et al., 1999; Roumestan et al., 2007). In mice, antidepressants also inhibit LPS- (Ohgi et al., 2013) or cytokine-induced depressive-like behaviors (Yirmiya et al., 2001), whereas antidepressant medications have also been shown to increase proinflammatory cytokines in mice (Warner-Schmidt et al., 2011).

Conversely, blocking proinflammatory cytokines provides antidepressant effects. Thus, FDA-approved monoclonal antibodies and other cytokine inhibitors, approved for other inflammatory conditions, to block individual cytokines in MDD patients, provide significant antidepressant effects in some cohorts of MDD patients (Kappelmann et al., 2016). Thus, TNFα inhibitors such as adalimumab (Loftus et al., 2008; Menter et al., 2010), etanercept (Tyring et al., 2006, 2013), or infliximab (Raison et al., 2013); IL-12/IL-23 antagonists (Langley et al., 2010); IL-6 antagonists (Kappelmann et al., 2016); or IL-4Ra antagonists (Simpson et al., 2015) are more efficacious than placebo in the treatment of MDD. Taken together, modulation of pro-inflammatory and/or anti-inflammatory cytokines might provide antidepressant actions and antidepressant medications might shift the balance toward anti-inflammatory responses.

Modulatory Effects of Cytokines on Therapeutic Responses to Antidepressants

Antidepressant resistance has often been associated with an activation of the inflammatory system. Thus, MDD patients displaying inflammation prior to treatment are less responsive to antidepressants (Mikova et al., 2001), or lithium (Sluzewska et al., 1997). Furthermore, IL-6 levels, but not TNFα, prior to treatment with amitriptyline predict the patient response (Lanquillon et al., 2000; Kubera et al., 2001a; Benedetti et al., 2002). Altogether, response to antidepressant treatments is impaired by certain proinflammatory cytokines, and can be overcome by co-administering anti-inflammatory drugs (Köhler et al., 2014). The results of a meta-analysis supported the notion that adjuvant anti-inflammatory therapy in addition to antidepressant treatment of depression can provide significant clinical benefit (Köhler et al., 2014). However, non-steroidal anti-inflammatory drugs (NSAIDs) might not be the optimal anti-inflammatory approach to use, as in the SADHART study, all patients receiving NSAIDs continued to exhibit cytokine elevations (Glassman et al., 2002). Recent findings point toward potential cytokine biomarker for antidepressant response. Thus, higher level of IL-17A predicts better outcomes in the response to bupropion and SSRI combination therapy (Jha et al., 2017). Furthermore, levels of CRP predict treatment response differentially according to the antidepressant. Thus, low level of CRP predicts better response to escitalopram, whereas high level CRP is associated with better outcome with nortriptyline (Uher et al., 2014). Overall these findings suggest that cytokine levels may influence a patient’s response to antidepressant treatment.

Conclusion

Inflammation, mainly cytokine production, has become an accepted part of the etiology of MDD, however, the role of the cytokines in promoting depression remains unclear. Targeting cytokines seems to provide some clinical benefits, but unless the immune response is better characterized in MDD patients, attenuating cytokine production in MDD patients might remain challenging.

Author Contributions

EM-R, JL, RW, SS, and EB wrote the review.

Funding

The research in EB’s laboratory is supported by grants from the NIMH (MH104656 and MH110415).

Conflict of Interest Statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

We thank Dr. Richard S. Jope for his critical comments on the manuscript.

References

Ahmetspahic, D., Schwarte, K., Ambree, O., Burger, C., Falcone, V., Seiler, K., et al. (2018). Altered B cell homeostasis in patients with major depressive disorder and normalization of CD5 surface expression on regulatory B cells in treatment responders. J. Neuroimmune Pharmacol. 13, 90–99. doi: 10.1007/s11481-017-9763-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Alcocer-Gomez, E., Casas-Barquero, N., Williams, M. R., Romero-Guillena, S. L., Canadas-Lozano, D., Bullon, P., et al. (2017). Antidepressants induce autophagy dependent-NLRP3-inflammasome inhibition in Major depressive disorder. Pharmacol. Res. 121, 114–121. doi: 10.1016/j.phrs.2017.04.028

PubMed Abstract | CrossRef Full Text | Google Scholar

Alcocer-Gomez, E., de Miguel, M., Casas-Barquero, N., Nunez-Vasco, J., Sanchez-Alcazar, J. A., Fernandez-Rodriguez, A., et al. (2014). NLRP3 inflammasome is activated in mononuclear blood cells from patients with major depressive disorder. Brain Behav. Immun. 36, 111–117.

PubMed Abstract | Google Scholar

Alcocer-Gomez, E., Ulecia-Moron, C., Marin-Aguilar, F., Rybkina, T., Casas-Barquero, N., Ruiz-Cabello, J., et al. (2016). Stress-induced depressive behaviors require a functional NLRP3 inflammasome. Mol. Neurobiol. 53, 4874–4882. doi: 10.1007/s12035-015-9408-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Altshuler, L. L., Abulseoud, O. A., Foland-Ross, L., Bartzokis, G., Chang, S., Mintz, J., et al. (2010). Amygdala astrocyte reduction in subjects with major depressive disorder but not bipolar disorder. Bipolar Disord. 12, 541–549. doi: 10.1111/j.1399-5618.2010.00838.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Anderson, R. J., Freedland, K. E., Clouse, R. E., and Lustman, P. J. (2001). The prevalence of comorbid depression in adults with diabetes: a meta-analysis. Diabetes Care 24, 1069–1078.

Google Scholar

Banasr, M., and Duman, R. S. (2008). Glial loss in the prefrontal cortex is sufficient to induce depressive-like behaviors. Biol. Psychiatry 64, 863–870. doi: 10.1016/j.biopsych.2008.06.008

PubMed Abstract | CrossRef Full Text | Google Scholar

Baune, B. T., Smith, E., Reppermund, S., Air, T., Samaras, K., Lux, O., et al. (2012). Inflammatory biomarkers predict depressive, but not anxiety symptoms during aging: the prospective Sydney Memory and Aging Study. Psychoneuroendocrinology 37, 1521–1530.

Google Scholar

Benarroch, E. E. (2013). Microglia: multiple roles in surveillance, circuit shaping, and response to injury. Neurology 81, 1079–1088. doi: 10.1212/WNL.0b013e3182a4a577

PubMed Abstract | CrossRef Full Text | Google Scholar

Benedetti, F., Lucca, A., Brambilla, F., Colombo, C., and Smeraldi, E. (2002). Interleukine-6 serum levels correlate with response to antidepressant sleep deprivation and sleep phase advance. Prog. Neuropsychopharmacol. Biol. Psychiatry 26, 1167–1170.

PubMed Abstract | Google Scholar

Bernik, T. R., Friedman, S. G., Ochani, M., DiRaimo, R., Ulloa, L., Yang, H., et al. (2002). Pharmacological stimulation of the cholinergic antiinflammatory pathway. J. Exp. Med. 195, 781–788.

Google Scholar

Beurel, E. (2011). Regulation by glycogen synthase kinase-3 of inflammation and T cells in CNS diseases. Front. Mol. Neurosci. 4:18. doi: 10.3389/fnmol.2011.00018

PubMed Abstract | CrossRef Full Text | Google Scholar

Beurel, E., Grieco, S. F., and Jope, R. S. (2015). Glycogen synthase kinase-3 (GSK3): regulation, actions, and diseases. Pharmacol. Ther. 148, 114–131. doi: 10.1016/j.pharmthera.2014.11.016

PubMed Abstract | CrossRef Full Text | Google Scholar

Beurel, E., Harrington, L. E., and Jope, R. S. (2013). Inflammatory T helper 17 cells promote depression-like behavior in mice. Biol. Psychiatry 73, 622–630. doi: 10.1016/j.biopsych.2012.09.021

PubMed Abstract | CrossRef Full Text | Google Scholar

Beurel, E., Lowell, J., and Jope, R. (2018). Distinct characteristics of hippocampal pathogenic TH17 cells in a mouse model of depression. Brain Behav. Immun. doi: 10.1016/j.bbi.2018.04.012 [Epub ahead of print].

PubMed Abstract | CrossRef Full Text | Google Scholar

Bilbo, S. D., and Schwarz, J. M. (2009). Early-life programming of later-life brain and behavior: a critical role for the immune system. Front. Behav. Neurosci. 3:14. doi: 10.3389/neuro.08.014.2009

PubMed Abstract | CrossRef Full Text | Google Scholar

Bitzer-Quintero, O. K., and Gonzalez-Burgos, I. (2012). Immune system in the brain: a modulatory role on dendritic spine morphophysiology? Neural Plast. 2012:348642. doi: 10.1155/2012/348642

PubMed Abstract | CrossRef Full Text | Google Scholar

Bjorkholm, C., and Monteggia, L. M. (2016). BDNF - a key transducer of antidepressant effects. Neuropharmacology 102, 72–79. doi: 10.1016/j.neuropharm.2015.10.034

PubMed Abstract | CrossRef Full Text | Google Scholar

Borsini, A., Zunszain, P. A., Thuret, S., and Pariante, C. M. (2015). The role of inflammatory cytokines as key modulators of neurogenesis. Trends Neurosci. 38, 145–157.

Google Scholar

Brachman, R. A., Lehmann, M. L., Maric, D., and Herkenham, M. (2015). Lymphocytes from chronically stressed mice confer antidepressant-like effects to naive mice. J. Neurosci. 35, 1530–1538. doi: 10.1523/JNEUROSCI.2278-14.2015

PubMed Abstract | CrossRef Full Text | Google Scholar

Branchi, I., Alboni, S., and Maggi, L. (2014). The role of microglia in mediating the effect of the environment in brain plasticity and behavior. Front. Cell. Neurosci. 8:390. doi: 10.3389/fncel.2014.00390

PubMed Abstract | CrossRef Full Text | Google Scholar

Caine, E. D., and Schwid, S. R. (2002). Multiple sclerosis, depression, and the risk of suicide. Neurology 59, 662–663.

Google Scholar

Cao, X., Li, L. P., Wang, Q., Wu, Q., Hu, H. H., Zhang, M., et al. (2013). Astrocyte-derived ATP modulates depressive-like behaviors. Nat. Med. 19, 773–777. doi: 10.1038/nm.3162

PubMed Abstract | CrossRef Full Text | Google Scholar

Capuron, L., Pagnoni, G., Demetrashvili, M., Woolwine, B. J., Nemeroff, C. B., Berns, G. S., et al. (2005). Anterior cingulate activation and error processing during interferon-alpha treatment. Biol. Psychiatry 58, 190–196.

PubMed Abstract | Google Scholar

Carreno, F. R., and Frazer, A. (2017). Vagal nerve stimulation for treatment-resistant depression. Neurotherapeutics 14, 716–727. doi: 10.1007/s13311-017-0537-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Chabry, J., Nicolas, S., Cazareth, J., Murris, E., Guyon, A., Glaichenhaus, N., et al. (2015). Enriched environment decreases microglia and brain macrophages inflammatory phenotypes through adiponectin-dependent mechanisms: relevance to depressive-like behavior. Brain Behav. Immun. 50, 275–287. doi: 10.1016/j.bbi.2015.07.018

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, Y., Jiang, T., Chen, P., Ouyang, J., Xu, G., Zeng, Z., et al. (2011). Emerging tendency towards autoimmune process in major depressive patients: a novel insight from Th17 cells. Psychiatry Res. 188, 224–230. doi: 10.1016/j.psychres.2010.10.029

PubMed Abstract | CrossRef Full Text | Google Scholar

Cheng, Y., Desse, S., Martinez, A., Worthen, R. J., Jope, R. S., and Beurel, E. (2018). TNFalpha disrupts blood brain barrier integrity to maintain prolonged depressive-like behavior in mice. Brain Behav. Immun. 69, 556–567. doi: 10.1016/j.bbi.2018.02.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Cheng, Y., Pardo, M., Armini, R. S., Martinez, A., Mouhsine, H., Zagury, J. F., et al. (2016). Stress-induced neuroinflammation is mediated by GSK3-dependent TLR4 signaling that promotes susceptibility to depression-like behavior. Brain Behav. Immun. 53, 207–222. doi: 10.1016/j.bbi.2015.12.012

PubMed Abstract | CrossRef Full Text | Google Scholar

Chugh, D., Nilsson, P., Afjei, S. A., Bakochi, A., and Ekdahl, C. T. (2013). Brain inflammation induces post-synaptic changes during early synapse formation in adult-born hippocampal neurons. Exp. Neurol. 250, 176–188. doi: 10.1016/j.expneurol.2013.09.005

PubMed Abstract | CrossRef Full Text | Google Scholar

Clark, S. M., Soroka, J. A., Song, C., Li, X., and Tonelli, L. H. (2016). CD4+ T cells confer anxiolytic and antidepressant-like effects, but enhance fear memory processes in Rag2(-/-) mice. Stress 19, 303–311. doi: 10.1080/10253890.2016.1191466

PubMed Abstract | CrossRef Full Text | Google Scholar

Cobb, J. A., O’Neill, K., Milner, J., Mahajan, G. J., Lawrence, T. J., May, W. L., et al. (2016). Density of GFAP-immunoreactive astrocytes is decreased in left hippocampi in major depressive disorder. Neuroscience 316, 209–220. doi: 10.1016/j.neuroscience.2015.12.044

PubMed Abstract | CrossRef Full Text | Google Scholar

Cohen, H., Ziv, Y., Cardon, M., Kaplan, Z., Matar, M. A., Gidron, Y., et al. (2006). Maladaptation to mental stress mitigated by the adaptive immune system via depletion of naturally occurring regulatory CD4+CD25+ cells. J. Neurobiol. 66, 552–563.

Google Scholar

Corona, A. W., Huang, Y., O’Connor, J. C., Dantzer, R., Kelley, K. W., Popovich, P. G., et al. (2010). Fractalkine receptor (CX3CR1) deficiency sensitizes mice to the behavioral changes induced by lipopolysaccharide. J. Neuroinflammation 7:93. doi: 10.1186/1742-2094-7-93

PubMed Abstract | CrossRef Full Text | Google Scholar

Cotter, D., Mackay, D., Chana, G., Beasley, C., Landau, S., and Everall, I. P. (2002). Reduced neuronal size and glial cell density in area 9 of the dorsolateral prefrontal cortex in subjects with major depressive disorder. Cereb. Cortex 12, 386–394.

PubMed Abstract | Google Scholar

Cui, W., Mizukami, H., Yanagisawa, M., Aida, T., Nomura, M., Isomura, Y., et al. (2014). Glial dysfunction in the mouse habenula causes depressive-like behaviors and sleep disturbance. J. Neurosci. 34, 16273–16285. doi: 10.1523/JNEUROSCI.1465-14.2014

PubMed Abstract | CrossRef Full Text | Google Scholar

Czeh, B., Simon, M., Schmelting, B., Hiemke, C., and Fuchs, E. (2006). Astroglial plasticity in the hippocampus is affected by chronic psychosocial stress and concomitant fluoxetine treatment. Neuropsychopharmacology 31, 1616–1626.

PubMed Abstract | Google Scholar

Dantzer, R. (2009). Cytokine, sickness behavior, and depression. Immunol. Allergy Clin. North Am. 29, 247–264.

Google Scholar

Dantzer, R., O’Connor, J. C., Freund, G. G., Johnson, R. W., and Kelley, K. W. (2008). From inflammation to sickness and depression: when the immune system subjugates the brain. Nat. Rev. Neurosci. 9, 46–56.

PubMed Abstract | Google Scholar

Das, S., and Basu, A. (2008). Inflammation: a new candidate in modulating adult neurogenesis. J. Neurosci. Res. 86, 1199–1208.

PubMed Abstract | Google Scholar

Davis, S., Thomas, A., Perry, R., Oakley, A., Kalaria, R. N., and O’Brien, J. T. (2002). Glial fibrillary acidic protein in late life major depressive disorder: an immunocytochemical study. J. Neurol. Neurosurg. Psychiatry 73, 556–560.

PubMed Abstract | Google Scholar

Delpech, J. C., Madore, C., Nadjar, A., Joffre, C., Wohleb, E. S., and Laye, S. (2015). Microglia in neuronal plasticity: influence of stress. Neuropharmacology 96, 19–28. doi: 10.1016/j.neuropharm.2014.12.034

PubMed Abstract | CrossRef Full Text | Google Scholar

den Haan, J. M., Arens, R., and van Zelm, M. C. (2014). The activation of the adaptive immune system: cross-talk between antigen-presenting cells, T cells and B cells. Immunol. Lett. 162, 103–112. doi: 10.1016/j.imlet.2014.10.011

PubMed Abstract | CrossRef Full Text | Google Scholar

D’Mello, C., Le, T., and Swain, M. G. (2009). Cerebral microglia recruit monocytes into the brain in response to tumor necrosis factoralpha signaling during peripheral organ inflammation. J. Neurosci. 29, 2089–2102. doi: 10.1523/JNEUROSCI.3567-08.2009

PubMed Abstract | CrossRef Full Text | Google Scholar

Dowlati, Y., Herrmann, N., Swardfager, W., Liu, H., Sham, L., Reim, E. K., et al. (2010). A meta-analysis of cytokines in major depression. Biol. Psychiatry 67, 446–457. doi: 10.1016/j.biopsych.2009.09.033

PubMed Abstract | CrossRef Full Text | Google Scholar

Engler, H., Bailey, M. T., Engler, A., and Sheridan, J. F. (2004). Effects of repeated social stress on leukocyte distribution in bone marrow, peripheral blood and spleen. J. Neuroimmunol. 148, 106–115.

PubMed Abstract | Google Scholar

Eulenburg, V., and Gomeza, J. (2010). Neurotransmitter transporters expressed in glial cells as regulators of synapse function. Brain Res. Rev. 63, 103–112. doi: 10.1016/j.brainresrev.2010.01.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Failor, K. L., Desyatnikov, Y., Finger, L. A., and Firestone, G. L. (2007). Glucocorticoid-induced degradation of glycogen synthase kinase-3 protein is triggered by serum- and glucocorticoid-induced protein kinase and Akt signaling and controls beta-catenin dynamics and tight junction formation in mammary epithelial tumor cells. Mol. Endocrinol. 21, 2403–2415.

PubMed Abstract | Google Scholar

Fazzino, F., Urbina, M., Cedeño, N., and Lima, L. (2009). Fluoxetine treatment to rats modifies serotonin transporter and cAMP in lymphocytes, CD4+ and CD8+ subpopulations and interleukins 2 and 4. Int. Immunopharmacol. 9, 463–467. doi: 10.1016/j.intimp.2009.01.011

PubMed Abstract | CrossRef Full Text | Google Scholar

Felger, J. C., Haroon, E., Patel, T. A., Goldsmith, D. R., Wommack, E. C., Woolwine, B. J., et al. (2018). What does plasma CRP tell us about peripheral and central inflammation in depression? Mol. Psychiatry (in press). doi: 10.1038/s41380-018-0096-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Felger, J. C., and Lotrich, F. E. (2013). Inflammatory cytokines in depression: neurobiological mechanisms and therapeutic implications. Neuroscience 246, 199–229. doi: 10.1016/j.neuroscience.2013.04.060

PubMed Abstract | CrossRef Full Text | Google Scholar

Filiano, A. J., Gadani, S. P., and Kipnis, J. (2017). How and why do T cells and their derived cytokines affect the injured and healthy brain? Nat. Rev. Neurosci. 18, 375–384. doi: 10.1038/nrn.2017.39

PubMed Abstract | CrossRef Full Text | Google Scholar

Filiano, A. J., Xu, Y., Tustison, N. J., Marsh, R. L., Baker, W., Smirnov, I., et al. (2016). Unexpected role of interferon-gamma in regulating neuronal connectivity and social behaviour. Nature 535, 425–429.

PubMed Abstract | Google Scholar

Fleshner, M. (2013). Stress-evoked sterile inflammation, danger associated molecular patterns (DAMPs), microbial associated molecular patterns (MAMPs) and the inflammasome. Brain Behav. Immun. 27, 1–7. doi: 10.1016/j.bbi.2012.08.012

PubMed Abstract | CrossRef Full Text | Google Scholar

Fleshner, M., Frank, M., and Maier, S. F. (2017). Danger signals and inflammasomes: stress-evoked sterile inflammation in mood disorders. Neuropsychopharmacology 42, 36–45. doi: 10.1038/npp.2016.125

PubMed Abstract | CrossRef Full Text | Google Scholar

Fornaro, M., Rocchi, G., Escelsior, A., Contini, P., and Martino, M. (2013). Might different cytokine trends in depressed patients receiving duloxetine indicate differential biological backgrounds. J. Affect. Disord. 145, 300–307. doi: 10.1016/j.jad.2012.08.007

PubMed Abstract | CrossRef Full Text | Google Scholar

Franklin, T. C., Wohleb, E. S., Zhang, Y., Fogaca, M., Hare, B., and Duman, R. S. (2018). Persistent increase in microglial RAGE contributes to chronic stress-induced priming of depressive-like behavior. Biol. Psychiatry 83, 50–60. doi: 10.1016/j.biopsych.2017.06.034

PubMed Abstract | CrossRef Full Text | Google Scholar

Fruhbeis, C., Frohlich, D., Kuo, W. P., and Kramer-Albers, E. M. (2013). Extracellular vesicles as mediators of neuron-glia communication. Front. Cell. Neurosci. 7:182. doi: 10.3389/fncel.2013.00182

PubMed Abstract | CrossRef Full Text | Google Scholar

Gimeno, D., Kivimaki, M., Brunner, E. J., Elovainio, M., De Vogli, R., Steptoe, A., et al. (2009). Associations of C-reactive protein and interleukin-6 with cognitive symptoms of depression: 12-year follow-up of the Whitehall II study. Psychol. Med. 39, 413–423. doi: 10.1017/S0033291708003723

PubMed Abstract | CrossRef Full Text | Google Scholar

Gittins, R. A., and Harrison, P. J. (2011). A morphometric study of glia and neurons in the anterior cingulate cortex in mood disorder. J. Affect. Disord. 133, 328–332. doi: 10.1016/j.jad.2011.03.042

PubMed Abstract | CrossRef Full Text | Google Scholar

Glassman, A. H., O’Connor, C. M., Califf, R. M., Swedberg, K., Schwartz, P., Bigger, J. T. Jr., et al. (2002). Sertraline treatment of major depression in patients with acute MI or unstable angina. JAMA 288, 701–709.

Google Scholar

Green, H. F., and Nolan, Y. M. (2014). Inflammation and the developing brain: consequences for hippocampal neurogenesis and behavior. Neurosci. Biobehav. Rev. 40, 20–34. doi: 10.1016/j.neubiorev.2014.01.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Guo, H., Callaway, J. B., and Ting, J. P. (2015). Inflammasomes: mechanism of action, role in disease, and therapeutics. Nat. Med. 21, 677–687. doi: 10.1038/nm.3893

PubMed Abstract | CrossRef Full Text | Google Scholar

Haapakoski, R., Mathieu, J., Ebmeier, K. P., Alenius, H., and Kivimäki, M. (2015). Cumulative meta-analysis of interleukins 6 and 1β, tumour necrosis factor α and C-reactive protein in patients with major depressive disorder. Brain Behav. Immun. 49, 206–215.

Google Scholar

Hamidi, M., Drevets, W. C., and Price, J. L. (2004). Glial reduction in amygdala in major depressive disorder is due to oligodendrocytes. Biol. Psychiatry 55, 563–569.

Google Scholar

Hannestad, J., DellaGioia, N., and Bloch, M. (2011). The effect of antidepressant medication treatment on serum levels of inflammatory cytokines: a meta-analysis. Neuropsychopharmacology 36, 2452–2459. doi: 10.1038/npp.2011.132

PubMed Abstract | CrossRef Full Text | Google Scholar

Haroon, E., Miller, A. H., and Sanacora, G. (2017). Inflammation, glutamate, and glia: a trio of trouble in mood disorders. Neuropsychopharmacology 42, 193–215. doi: 10.1038/npp.2016.199

PubMed Abstract | CrossRef Full Text | Google Scholar

Harry, G. J., and Kraft, A. D. (2012). Microglia in the developing brain: a potential target with lifetime effects. Neurotoxicology 33, 191–206. doi: 10.1016/j.neuro.2012.01.012

PubMed Abstract | CrossRef Full Text | Google Scholar

Haydon, P. G., and Carmignoto, G. (2006). Astrocyte control of synaptic transmission and neurovascular coupling. Physiol. Rev. 86, 1009–1031.

Google Scholar

Heidt, T., Sager, H. B., Courties, G., Dutta, P., Iwamoto, Y., Zaltsman, A., et al. (2014). Chronic variable stress activates hematopoietic stem cells. Nat. Med. 20, 754–758. doi: 10.1038/nm.3589

PubMed Abstract | CrossRef Full Text | Google Scholar

Henry, C. J., Huang, Y., Wynne, A., Hanke, M., Himler, J., Bailey, M. T., et al. (2008). Minocycline attenuates lipopolysaccharide (LPS)-induced neuroinflammation, sickness behavior, and anhedonia. J. Neuroinflammation 5:15. doi: 10.1186/1742-2094-5-15

PubMed Abstract | CrossRef Full Text | Google Scholar

Herbert, J., Goodyer, I. M., Grossman, A. B., Hastings, M. H., de Kloet, E. R., Lightman, S. L., et al. (2006). Do corticosteroids damage the brain? J. Neuroendocrinol. 18, 393–411.

Google Scholar

Herbert, T. B., and Cohen, S. (1993). Depression and immunity: a meta-analytic review. Psychol. Bull. 113, 472–486.

Google Scholar

Herkenham, M., and Kigar, S. L. (2017). Contributions of the adaptive immune system to mood regulation: mechanisms and pathways of neuroimmune interactions. Prog. Neuropsychopharmacol. Biol. Psychiatry 79, 49–57. doi: 10.1016/j.pnpbp.2016.09.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Hernández, M. E., Mendieta, D., Martínez-Fong, D., Loría, F., Moreno, J., Estrada, I., et al. (2008). Variations in circulating cytokine levels during 52 week course of treatment with SSRI for major depressive disorder. Eur. Neuropsychopharmacol. 18, 917–924. doi: 10.1016/j.euroneuro.2008.08.001

PubMed Abstract | CrossRef Full Text | Google Scholar

Hiles, S. A., Baker, A. L., de Malmanche, T., and Attia, J. (2012). A meta-analysis of differences in IL-6 and IL-10 between people with and without depression: exploring the causes of heterogeneity. Brain Behav. Immun. 26, 1180–1188. doi: 10.1016/j.bbi.2012.06.001

PubMed Abstract | CrossRef Full Text | Google Scholar

Himmerich, H., Milenovic, S., Fulda, S., Plumakers, B., Sheldrick, A. J., Michel, T. M., et al. (2010). Regulatory T cells increased while IL-1beta decreased during antidepressant therapy. J. Psychiatr. Res. 44, 1052–1057. doi: 10.1016/j.jpsychires.2010.03.005

PubMed Abstract | CrossRef Full Text | Google Scholar

Hinwood, M., Morandini, J., Day, T. A., and Walker, F. R. (2012). Evidence that microglia mediate the neurobiological effects of chronic psychological stress on the medial prefrontal cortex. Cereb. Cortex 22, 1442–1454.

Google Scholar

Hinze-Selch, D., Schuld, A., Kraus, T., Kühn, M., Uhr, M., Haack, M., et al. (2000). Effects of antidepressants on weight and on the plasma levels of leptin, TNF-α and soluble TNF receptors: a longitudinal study in patients treated with amitriptyline or paroxetine. Neuropsychopharmacology 23, 13–19.

Google Scholar

Ho, P.-S., Yeh, Y.-W., Huang, S.-Y., and Liang, C.-S. (2015). A shift toward T helper 2 responses and an increase in modulators of innate immunity in depressed patients treated with escitalopram. Psychoneuroendocrinology 53, 246–255. doi: 10.1016/j.psyneuen.2015.01.008

PubMed Abstract | CrossRef Full Text | Google Scholar

Hodes, G. E., Kana, V., Menard, C., Merad, M., and Russo, S. J. (2015). Neuroimmune mechanisms of depression. Nat. Neurosci. 18, 1386–1393.

Google Scholar

Howren, M. B., Lamkin, D. M., and Suls, J. (2009). Associations of depression with C-reactive protein, IL-1, and IL-6: a meta-analysis. Psychosom. Med. 71, 171–186. doi: 10.1097/PSY.0b013e3181907c1b

PubMed Abstract | CrossRef Full Text | Google Scholar

Hundal, O. (2007). Major depressive disorder viewed as a dysfunction in astroglial bioenergetics. Med. Hypotheses 68, 370–377.

PubMed Abstract | Google Scholar

Hung, Y. Y., Kang, H. Y., Huang, K. W., and Huang, T. L. (2014). Association between toll-like receptors expression and major depressive disorder. Psychiatry Res. 220, 283–286. doi: 10.1016/j.psychres.2014.07.074

PubMed Abstract | CrossRef Full Text | Google Scholar

Irwin, M. R., and Miller, A. H. (2007). Depressive disorders and immunity: 20 years of progress and discovery. Brain Behav. Immun. 21, 374–383.

PubMed Abstract | Google Scholar

Iwata, M., Ota, K. T., Li, X. Y., Sakaue, F., Li, N., Dutheil, S., et al. (2016). Psychological stress activates the inflammasome via release of adenosine triphosphate and stimulation of the purinergic type 2X7 receptor. Biol. Psychiatry 80, 12–22. doi: 10.1016/j.biopsych.2015.11.026

PubMed Abstract | CrossRef Full Text | Google Scholar

Jakubs, K., Bonde, S., Iosif, R. E., Ekdahl, C. T., Kokaia, Z., Kokaia, M., et al. (2008). Inflammation regulates functional integration of neurons born in adult brain. J. Neurosci. 28, 12477–12488. doi: 10.1523/JNEUROSCI.3240-08.2008

PubMed Abstract | CrossRef Full Text | Google Scholar

Jha, M. K., Minhajuddin, A., Gadad, B. S., Greer, T. L., Mayes, T. L., and Trivedi, M. H. (2017). Interleukin 17 selectively predicts better outcomes with bupropion-SSRI combination: novel T cell biomarker for antidepressant medication selection. Brain Behav. Immun. 66, 103–110. doi: 10.1016/j.bbi.2017.07.005

PubMed Abstract | CrossRef Full Text | Google Scholar

Jiang, C., and Salton, S. R. (2013). The role of neurotrophins in major depressive disorder. Transl. Neurosci. 4, 46–58.

Google Scholar

Jo, W. K., Zhang, Y., Emrich, H. M., and Dietrich, D. E. (2015). Glia in the cytokine-mediated onset of depression: fine tuning the immune response. Front. Cell. Neurosci. 9:268. doi: 10.3389/fncel.2015.00268

PubMed Abstract | CrossRef Full Text | Google Scholar

Jope, R. S., Cheng, Y., Lowell, J. A., Worthen, R. J., Sitbon, Y. H., and Beurel, E. (2017). Stressed and inflamed, can GSK3 be blamed? Trends Biochem. Sci. 42, 180–192. doi: 10.1016/j.tibs.2016.10.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Kang, H. J., Voleti, B., Hajszan, T., Rajkowska, G., Stockmeier, C. A., Licznerski, P., et al. (2012). Decreased expression of synapse-related genes and loss of synapses in major depressive disorder. Nat. Med. 18, 1413–1417.

PubMed Abstract | Google Scholar

Kappelmann, N., Lewis, G., Dantzer, R., Jones, P. B., and Khandaker, G. M. (2016). Antidepressant activity of anti-cytokine treatment: a systematic review and meta-analysis of clinical trials of chronic inflammatory conditions. Mol. Psychiatry 23, 335–343. doi: 10.1038/mp.2016.167

PubMed Abstract | CrossRef Full Text | Google Scholar

Kettenmann, H., Kirchhoff, F., and Verkhratsky, A. (2013). Microglia: new roles for the synaptic stripper. Neuron 77, 10–18. doi: 10.1016/j.neuron.2012.12.023

PubMed Abstract | CrossRef Full Text | Google Scholar

Khandaker, G. M., Pearson, R. M., Zammit, S., Lewis, G., and Jones, P. B. (2014). Association of serum interleukin 6 and C-reactive protein in childhood with depression and psychosis in young adult life: a population-based longitudinal study. JAMA Psychiatry 71, 1121–1128. doi: 10.1001/jamapsychiatry.2014.1332

PubMed Abstract | CrossRef Full Text | Google Scholar

Kim, Y. K., Suh, I. B., Kim, H., Han, C. S., Lim, C. S., Choi, S. H., et al. (2002). The plasma levels of interleukin-12 in schizophrenia, major depression, and bipolar mania: effects of psychotropic drugs. Mol. Psychiatry 7, 1107–1114.

PubMed Abstract | Google Scholar

Köhler, C. A., Freitas, T. H., Maes, M., Andrade, N. Q., Liu, C. S., Fernandes, B. S., et al. (2017). Peripheral cytokine and chemokine alterations in depression: a meta-analysis of 82 studies. Acta Psychiatr. Scand. 135, 373–387. doi: 10.1111/acps.12698

PubMed Abstract | CrossRef Full Text | Google Scholar

Kohler, C. A., Freitas, T. H., Maes, M., de Andrade, N. Q., Liu, C. S., Fernandes, B. S., et al. (2017). Peripheral cytokine and chemokine alterations in depression: a meta-analysis of 82 studies. Acta Psychiatr. Scand. 135, 373–387. doi: 10.1111/acps.12698

PubMed Abstract | CrossRef Full Text | Google Scholar

Köhler, O., Benros, M. E., Nordentoft, M., Farkouh, M. E., Iyengar, R. L., Mors, O., et al. (2014). Effect of anti-inflammatory treatment on depression, depressive symptoms, and adverse effects: a systematic review and meta-analysis of randomized clinical trials. JAMA Psychiatry 71, 1381–1391. doi: 10.1001/jamapsychiatry.2014.1611

PubMed Abstract | CrossRef Full Text | Google Scholar

Kohman, R. A., and Rhodes, J. S. (2013). Neurogenesis, inflammation and behavior. Brain Behav. Immun. 27, 22–32. doi: 10.1016/j.bbi.2012.09.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Korn, T., and Kallies, A. (2017). T cell responses in the central nervous system. Nat. Rev. Immunol. 17, 179–194. doi: 10.1038/nri.2016.144

PubMed Abstract | CrossRef Full Text | Google Scholar

Kreisel, T., Frank, M. G., Licht, T., Reshef, R., Ben-Menachem-Zidon, O., Baratta, M. V., et al. (2014). Dynamic microglial alterations underlie stress-induced depressive-like behavior and suppressed neurogenesis. Mol. Psychiatry 19, 699–709. doi: 10.1038/mp.2013.155

PubMed Abstract | CrossRef Full Text | Google Scholar

Kubera, M., Lin, A. H., Kenis, G., Bosmans, E., van Bockstaele, D., and Maes, M. (2001a). Anti-inflammatory effects of antidepressants through suppression of the interferon-gamma/interleukin-10 production ratio. J. Clin. Psychopharmacol. 21, 199–206.

PubMed Abstract | Google Scholar

Kubera, M., Maes, M., Holan, V., Basta-Kaim, A., Roman, A., and Shani, J. (2001b). Prolonged desipramine treatment increases the production of interleukin-10, an anti-inflammatory cytokine, in C57BL/6 mice subjected to the chronic mild stress model of depression. J. Affect. Disord. 63, 171–178.

PubMed Abstract | Google Scholar

Langley, R. G., Feldman, S. R., Han, C., Schenkel, B., Szapary, P., Hsu, M.-C., et al. (2010). Ustekinumab significantly improves symptoms of anxiety, depression, and skin-related quality of life in patients with moderate-to-severe psoriasis: results from a randomized, double-blind, placebo-controlled phase III trial. J. Am. Acad. Dermatol. 63, 457–465. doi: 10.1016/j.jaad.2009.09.014

PubMed Abstract | CrossRef Full Text | Google Scholar

Lanquillon, S., Krieg, J. C., Bening-Abu-Shach, U., and Vedder, H. (2000). Cytokine production and treatment response in major depressive disorder. Neuropsychopharmacology 22, 370–379.

Google Scholar

Le Thuc, O., Blondeau, N., Nahon, J. L., and Rovere, C. (2015). The complex contribution of chemokines to neuroinflammation: switching from beneficial to detrimental effects. Ann. N. Y. Acad. Sci. 1351, 127–140. doi: 10.1111/nyas.12855

PubMed Abstract | CrossRef Full Text | Google Scholar

Levine, J., Barak, Y., Chengappa, K. N. R., Rapoport, A., Rebey, M., and Barak, V. (1999). Cerebrospinal cytokine levels in patients with acute depression. Neuropsychobiology 40, 171–176.

Google Scholar

Lewitus, G. M., Cohen, H., and Schwartz, M. (2008). Reducing post-traumatic anxiety by immunization. Brain Behav. Immun. 22, 1108–1114. doi: 10.1016/j.bbi.2008.05.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Lewitus, G. M., and Schwartz, M. (2009). Behavioral immunization: immunity to self-antigens contributes to psychological stress resilience. Mol. Psychiatry 14, 532–536. doi: 10.1038/mp.2008.103

PubMed Abstract | CrossRef Full Text | Google Scholar

Lewitus, G. M., Wilf-Yarkoni, A., Ziv, Y., Shabat-Simon, M., Gersner, R., Zangen, A., et al. (2009). Vaccination as a novel approach for treating depressive behavior. Biol. Psychiatry 65, 283–288. doi: 10.1016/j.biopsych.2008.07.014

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, Y., Xiao, B., Qiu, W., Yang, L., Hu, B., Tian, X., et al. (2010). Altered expression of CD4+CD25+ regulatory T cells and its 5-HT(1a) receptor in patients with major depression disorder. J. Affect. Disord. 124, 68–75. doi: 10.1016/j.jad.2009.10.018

PubMed Abstract | CrossRef Full Text | Google Scholar

Lindqvist, D., Janelidze, S., Hagell, P., Erhardt, S., Samuelsson, M., Minthon, L., et al. (2009). Interleukin-6 is elevated in the cerebrospinal fluid of suicide attempters and related to symptom severity. Biol. Psychiatry 66, 287–292. doi: 10.1016/j.biopsych.2009.01.030

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, W., Wang, H., Wang, Y., Li, H., and Ji, L. (2015). Metabolic factors-triggered inflammatory response drives antidepressant effects of exercise in CUMS rats. Psychiatry Res. 228, 257–264. doi: 10.1016/j.psychres.2015.05.102

PubMed Abstract | CrossRef Full Text | Google Scholar

Loftus, E. V., Feagan, B. G., Colombel, J.-F., Rubin, D. T., Wu, E. Q., Andrew, P. Y., et al. (2008). Effects of adalimumab maintenance therapy on health-related quality of life of patients with Crohn’s disease: patient-reported outcomes of the CHARM trial. Am. J. Gastroenterol. 103, 3132–3141. doi: 10.1111/j.1572-0241.2008.02175.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Lossinsky, A. S., and Shivers, R. R. (2004). Structural pathways for macromolecular and cellular transport across the blood-brain barrier during inflammatory conditions. Review. Histol. Histopathol. 19, 535–564. doi: 10.14670/HH-19.535

PubMed Abstract | CrossRef Full Text | Google Scholar

Maes, M. (1995). Evidence for an immune response in major depression: a review and hypothesis. Prog. Neuropsychopharmacol. Biol. Psychiatry 19, 11–38.

Google Scholar

Maes, M. (1999). Major depression and activation of the inflammatory response system. Adv. Exp. Med. Biol. 461, 25–46.

Google Scholar

Maes, M., Bosmans, E., De Jongh, R., Kenis, G., Vandoolaeghe, E., and Neels, H. (1997). Increased serum IL-6 and IL-1 receptor antagonist concentrations in major depression and treatment resistant depression. Cytokine 9, 853–858.

Google Scholar

Maes, M., Song, C., Lin, A., De Jongh, R., Van Gastel, A., Kenis, G., et al. (1998). The effects of psychological stress on humans: increased production of pro-inflammatory cytokines and Th1-like response in stress-induced anxiety. Cytokine 10, 313–318.

Google Scholar

Maes, M., Song, C., Lin, A.-H., Bonaccorso, S., Kenis, G., De Jongh, R., et al. (1999). Negative immunoregulatory effects of antidepressants: inhibition of interferon-γ and stimulation of interleukin-10 secretion. Neuropsychopharmacology 20, 370–379.

Google Scholar

Maes, M., Van der Planken, M., Stevens, W. J., Peeters, D., DeClerck, L. S., Bridts, C. H., et al. (1992). Leukocytosis, monocytosis and neutrophilia: hallmarks of severe depression. J. Psychiatr. Res. 26, 125–134.

PubMed Abstract | Google Scholar

Manji, H. K., Quiroz, J. A., Sporn, J., Payne, J. L., Denicoff, K., Gray, N., et al. (2003). Enhancing neuronal plasticity and cellular resilience to develop novel, improved therapeutics for difficult-to-treat depression. Biol. Psychiatry 53, 707–742.

PubMed Abstract | Google Scholar

Martin, J. L., Magistretti, P. J., and Allaman, I. (2013). Regulation of neurotrophic factors and energy metabolism by antidepressants in astrocytes. Curr. Drug Targets 14, 1308–1321.

PubMed Abstract | Google Scholar

Martin, M., Rehani, K., Jope, R. S., and Michalek, S. M. (2005). Toll-like receptor-mediated cytokine production is differentially regulated by glycogen synthase kinase 3. Nat. Immunol. 6, 777–784.

PubMed Abstract | Google Scholar

Martino, M., Rocchi, G., Escelsior, A., and Fornaro, M. (2012). Immunomodulation mechanism of antidepressants: interactions between serotonin/norepinephrine balance and Th1/Th2 balance. Curr. Neuropharmacol. 10, 97–123. doi: 10.2174/157015912800604542

PubMed Abstract | CrossRef Full Text | Google Scholar

Mathieu, P., Battista, D., Depino, A., Roca, V., Graciarena, M., and Pitossi, F. (2010). The more you have, the less you get: the functional role of inflammation on neuronal differentiation of endogenous and transplanted neural stem cells in the adult brain. J. Neurochem. 112, 1368–1385. doi: 10.1111/j.1471-4159.2009.06548.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Mazzardo-Martins, L., Martins, D. F., Stramosk, J., Cidral-Filho, F. J., and Santos, A. R. (2012). Glycogen synthase kinase 3-specific inhibitor AR-A014418 decreases neuropathic pain in mice: evidence for the mechanisms of action. Neuroscience 226, 411–420. doi: 10.1016/j.neuroscience.2012.09.020

PubMed Abstract | CrossRef Full Text | Google Scholar

McAfoose, J., and Baune, B. T. (2009). Evidence for a cytokine model of cognitive function. Neurosci. Biobehav. Rev. 33, 355–366. doi: 10.1016/j.neubiorev.2008.10.005

PubMed Abstract | CrossRef Full Text | Google Scholar

McKim, D. B., Weber, M. D., Niraula, A., Sawicki, C. M., Liu, X., Jarrett, B. L., et al. (2017). Microglial recruitment of IL-1beta-producing monocytes to brain endothelium causes stress-induced anxiety. Mol. Psychiatry 23, 1421–1431. doi: 10.1038/mp.2017.64

PubMed Abstract | CrossRef Full Text | Google Scholar

Medina, A., Watson, S. J., Bunney, W. Jr., Myers, R. M., Schatzberg, A., Barchas, J., et al. (2016). Evidence for alterations of the glial syncytial function in major depressive disorder. J. Psychiatr. Res. 72, 15–21. doi: 10.1016/j.jpsychires.2015.10.010

PubMed Abstract | CrossRef Full Text | Google Scholar

Menard, C., Hodes, G. E., and Russo, S. J. (2016). Pathogenesis of depression: insights from human and rodent studies. Neuroscience 321, 138–162. doi: 10.1016/j.neuroscience.2015.05.053

PubMed Abstract | CrossRef Full Text | Google Scholar

Menard, C., Pfau, M. L., Hodes, G. E., Kana, V., Wang, V. X., Bouchard, S., et al. (2017). Social stress induces neurovascular pathology promoting depression. Nat. Neurosci. 20, 1752–1760. doi: 10.1038/s41593-017-0010-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Menter, A., Augustin, M., Signorovitch, J., Andrew, P. Y., Wu, E. Q., Gupta, S. R., et al. (2010). The effect of adalimumab on reducing depression symptoms in patients with moderate to severe psoriasis: a randomized clinical trial. J. Am. Acad. Dermatol. 62, 812–818. doi: 10.1016/j.jaad.2009.07.022

PubMed Abstract | CrossRef Full Text | Google Scholar

Merendino, R. A., Di Rosa, A. E., Di Pasquale, G., Minciullo, P. L., Mangraviti, C., Costantino, A., et al. (2002). Interleukin-18 and CD30 serum levels in patients with moderate-severe depression. Mediators Inflamm. 11, 265–267.

PubMed Abstract | Google Scholar

Miguel-Hidalgo, J. J., Baucom, C., Dilley, G., Overholser, J. C., Meltzer, H. Y., Stockmeier, C. A., et al. (2000). Glial fibrillary acidic protein immunoreactivity in the prefrontal cortex distinguishes younger from older adults in major depressive disorder. Biol. Psychiatry 48, 861–873.

PubMed Abstract | Google Scholar

Miguel-Hidalgo, J. J., Waltzer, R., Whittom, A. A., Austin, M. C., Rajkowska, G., and Stockmeier, C. A. (2010). Glial and glutamatergic markers in depression, alcoholism, and their comorbidity. J. Affect. Disord. 127, 230–240. doi: 10.1016/j.jad.2010.06.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Mikova, O., Yakimova, R., Bosmans, E., Kenis, G., and Maes, M. (2001). Increased serum tumor necrosis factor alpha concentrations in major depression and multiple sclerosis. Eur. Neuropsychopharmacol. 11, 203–208.

PubMed Abstract | Google Scholar

Milior, G., Lecours, C., Samson, L., Bisht, K., Poggini, S., Pagani, F., et al. (2016). Fractalkine receptor deficiency impairs microglial and neuronal responsiveness to chronic stress. Brain Behav. Immun. 55, 114–125. doi: 10.1016/j.bbi.2015.07.024

PubMed Abstract | CrossRef Full Text | Google Scholar

Miller, A. H. (2010). Depression and immunity: a role for T cells? Brain Behav. Immun. 24, 1–8. doi: 10.1016/j.bbi.2009.09.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Miller, A. H., Maletic, V., and Raison, C. L. (2009). Inflammation and its discontents: the role of cytokines in the pathophysiology of major depression. Biol. Psychiatry 65, 732–741. doi: 10.1016/j.biopsych.2008.11.029

PubMed Abstract | CrossRef Full Text | Google Scholar

Miller, A. H., and Raison, C. L. (2016). The role of inflammation in depression: from evolutionary imperative to modern treatment target. Nat. Rev. Immunol. 16, 22–34. doi: 10.1038/nri.2015.5

PubMed Abstract | CrossRef Full Text | Google Scholar

Miyamoto, A., Wake, H., Moorhouse, A. J., and Nabekura, J. (2013). Microglia and synapse interactions: fine tuning neural circuits and candidate molecules. Front. Cell. Neurosci. 7:70. doi: 10.3389/fncel.2013.00070

PubMed Abstract | CrossRef Full Text | Google Scholar

Myint, A. M., Leonard, B. E., Steinbusch, H. W., and Kim, Y. K. (2005). Th1, Th2, and Th3 cytokine alterations in major depression. J. Affect. Disord. 88, 167–173.

Google Scholar

Nagy, C., Suderman, M., Yang, J., Szyf, M., Mechawar, N., Ernst, C., et al. (2015). Astrocytic abnormalities and global DNA methylation patterns in depression and suicide. Mol. Psychiatry 20, 320–328. doi: 10.1038/mp.2014.21

PubMed Abstract | CrossRef Full Text | Google Scholar

Nakajima, K., Tohyama, Y., Maeda, S., Kohsaka, S., and Kurihara, T. (2007). Neuronal regulation by which microglia enhance the production of neurotrophic factors for GABAergic, catecholaminergic, and cholinergic neurons. Neurochem. Int. 50, 807–820.

PubMed Abstract | Google Scholar

Noh, K. T., Son, K. H., Jung, I. D., Kang, T. H., Choi, C. H., and Park, Y. M. (2015). Glycogen synthase kinase-3beta (GSK-3beta) inhibition enhances dendritic cell-based cancer vaccine potency via suppression of interferon-gamma-induced indoleamine 2,3-dioxygenase expression. J. Biol. Chem. 290, 12394–12402. doi: 10.1074/jbc.M114.628578

PubMed Abstract | CrossRef Full Text | Google Scholar

O’Brien, S. M., Scott, L. V., and Dinan, T. G. (2004). Cytokines: abnormalities in major depression and implications for pharmacological treatment. Hum. Psychopharmacol. Clin. Exp. 19, 397–403.

PubMed Abstract | Google Scholar

O’Connor, J. C., Lawson, M. A., Andre, C., Moreau, M., Lestage, J., Castanon, N., et al. (2009). Lipopolysaccharide-induced depressive-like behavior is mediated by indoleamine 2, 3-dioxygenase activation in mice. Mol. Psychiatry 14, 511–522. doi: 10.1038/sj.mp.4002148

PubMed Abstract | CrossRef Full Text | Google Scholar

Ohgi, Y., Futamura, T., Kikuchi, T., and Hashimoto, K. (2013). Effects of antidepressants on alternations in serum cytokines and depressive-like behavior in mice after lipopolysaccharide administration. Pharmacol. Biochem. Behav. 103, 853–859. doi: 10.1016/j.pbb.2012.12.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Ondicova, K., Pecenak, J., and Mravec, B. (2010). The role of the vagus nerve in depression. Neuro Endocrinol. Lett. 31, 602–608.

Google Scholar

Ongur, D., Drevets, W. C., and Price, J. L. (1998). Glial reduction in the subgenual prefrontal cortex in mood disorders. Proc. Natl. Acad. Sci. U.S.A. 95, 13290–13295.

PubMed Abstract | Google Scholar

Pan, W., Stone, K. P., Hsuchou, H., Manda, V. K., Zhang, Y., and Kastin, A. J. (2011). Cytokine signaling modulates blood-brain barrier function. Curr. Pharm. Des. 17, 3729–3740.

Google Scholar

Pariante, C. M., and Lightman, S. L. (2008). The HPA axis in major depression: classical theories and new developments. Trends Neurosci. 31, 464–468. doi: 10.1016/j.tins.2008.06.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Pavlov, V. A., and Tracey, K. J. (2012). The vagus nerve and the inflammatory reflex–linking immunity and metabolism. Nat. Rev. Endocrinol. 8, 743–754. doi: 10.1038/nrendo.2012.189

PubMed Abstract | CrossRef Full Text | Google Scholar

Pavón, L., Sandoval-López, G., Hernández, M. E., Loría, F., Estrada, I., Pérez, M., et al. (2006). Th2 cytokine response in major depressive disorder patients before treatment. J. Neuroimmunol. 172, 156–165.

Google Scholar

Piletz, J. E., Halaris, A., Iqbal, O., Hoppensteadt, D., Fareed, J., Zhu, H., et al. (2009). Pro-inflammatory biomakers in depression: treatment with venlafaxine. World J. Biol. Psychiatry 10, 313–323. doi: 10.3109/15622970802573246

PubMed Abstract | CrossRef Full Text | Google Scholar

Polter, A., Beurel, E., Yang, S., Garner, R., Song, L., Miller, C. A., et al. (2010). Deficiency in the inhibitory serine-phosphorylation of glycogen synthase kinase-3 increases sensitivity to mood disturbances. Neuropsychopharmacology 35, 1761–1774. doi: 10.1038/npp.2010.43

PubMed Abstract | CrossRef Full Text | Google Scholar

Pucak, M. L., Carroll, K. A. L., Kerr, D. A., and Kaplin, A. L. (2007). Neuropsychiatric manifestations of depression in multiple sclerosis: neuroinflammatory, neuroendocrine, and neurotrophic mechanisms in the pathogenesis of immune-mediated depression. Dialogues Clin. Neurosci. 9, 125–139.

PubMed Abstract | Google Scholar

Raison, C. L., and Miller, A. H. (2017). Pathogen-host defense in the evolution of depression: insights into epidemiology, genetics, bioregional differences and female preponderance. Neuropsychopharmacology 42, 5–27. doi: 10.1038/npp.2016.194

PubMed Abstract | CrossRef Full Text | Google Scholar

Raison, C. L., Rutherford, R. E., Woolwine, B. J., Shuo, C., Schettler, P., Drake, D. F., et al. (2013). A randomized controlled trial of the tumor necrosis factor antagonist infliximab for treatment-resistant depression: the role of baseline inflammatory biomarkers. JAMA Psychiatry 70, 31–41. doi: 10.1001/2013.jamapsychiatry.4

PubMed Abstract | CrossRef Full Text | Google Scholar

Rajkowska, G., Hughes, J., Stockmeier, C. A., Javier Miguel-Hidalgo, J., and Maciag, D. (2013). Coverage of blood vessels by astrocytic endfeet is reduced in major depressive disorder. Biol. Psychiatry 73, 613–621. doi: 10.1016/j.biopsych.2012.09.024

PubMed Abstract | CrossRef Full Text | Google Scholar

Rajkowska, G., Miguel-Hidalgo, J. J., Wei, J., Dilley, G., Pittman, S. D., Meltzer, H. Y., et al. (1999). Morphometric evidence for neuronal and glial prefrontal cell pathology in major depression. Biol. Psychiatry 45, 1085–1098.

Google Scholar

Rajkowska, G., and Stockmeier, C. A. (2013). Astrocyte pathology in major depressive disorder: insights from human postmortem brain tissue. Curr. Drug Targets 14, 1225–1236.

PubMed Abstract | Google Scholar

Ransohoff, R. M., and Engelhardt, B. (2012). The anatomical and cellular basis of immune surveillance in the central nervous system. Nat. Rev. Immunol. 12, 623–635. doi: 10.1038/nri3265

PubMed Abstract | CrossRef Full Text | Google Scholar

Ransohoff, R. M., Schafer, D., Vincent, A., Blachere, N. E., and Bar-Or, A. (2015). Neuroinflammation: ways in which the immune system affects the brain. Neurotherapeutics 12, 896–909. doi: 10.1007/s13311-015-0385-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Reichenberg, A., Yirmiya, R., Schuld, A., Kraus, T., Haack, M., Morag, A., et al. (2001). Cytokine-associated emotional and cognitive disturbances in humans. Arch. Gen. Psychiatry 58, 445–452.

PubMed Abstract | Google Scholar

Rial, D., Lemos, C., Pinheiro, H., Duarte, J. M., Goncalves, F. Q., Real, J. I., et al. (2015). Depression as a glial-based synaptic dysfunction. Front. Cell. Neurosci. 9:521. doi: 10.3389/fncel.2015.00521

PubMed Abstract | CrossRef Full Text | Google Scholar

Rivest, S. (2009). Regulation of innate immune responses in the brain. Nat. Rev. Immunol. 9, 429–439. doi: 10.1038/nri2565

PubMed Abstract | CrossRef Full Text | Google Scholar

Rogatsky, I., Waase, C. L., and Garabedian, M. J. (1998). Phosphorylation and inhibition of rat glucocorticoid receptor transcriptional activation by glycogen synthase kinase-3 (GSK-3). Species-specific differences between human and rat glucocorticoid receptor signaling as revealed through GSK-3 phosphorylation. J. Biol. Chem. 273, 14315–14321.

PubMed Abstract | Google Scholar

Rook, G. A., and Lowry, C. A. (2008). The hygiene hypothesis and psychiatric disorders. Trends Immunol. 29, 150–158. doi: 10.1016/j.it.2008.01.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Roque, S., Correia-Neves, M., Mesquita, A. R., Palha, J. A., and Sousa, N. (2009). Interleukin-10: a key cytokine in depression? Cardiovasc. Psychiatry Neurol. 2009:187894. doi: 10.1155/2009/187894

PubMed Abstract | CrossRef Full Text | Google Scholar

Rouach, N., Koulakoff, A., Abudara, V., Willecke, K., and Giaume, C. (2008). Astroglial metabolic networks sustain hippocampal synaptic transmission. Science 322, 1551–1555. doi: 10.1126/science.1164022

PubMed Abstract | CrossRef Full Text | Google Scholar

Roumestan, C., Michel, A., Bichon, F., Portet, K., Detoc, M., Henriquet, C., et al. (2007). Anti-inflammatory properties of desipramine and fluoxetine. Respir. Res. 8:35.

Google Scholar

Rubinow, M. J., Mahajan, G., May, W., Overholser, J. C., Jurjus, G. J., Dieter, L., et al. (2016). Basolateral amygdala volume and cell numbers in major depressive disorder: a postmortem stereological study. Brain Struct. Funct. 221, 171–184. doi: 10.1007/s00429-014-0900-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Rudisch, B., and Nemeroff, C. B. (2003). Epidemiology of comorbid coronary artery disease and depression. Biol. Psychiatry 54, 227–240.

Google Scholar

Rush, A. J., Trivedi, M. H., Wisniewski, S. R., Nierenberg, A. A., Stewart, J. W., Warden, D., et al. (2006). Acute and longer-term outcomes in depressed outpatients requiring one or several treatment steps: a STARD report. Am. J. Psychiatry 163, 1905–1917.

PubMed Abstract | Google Scholar

Sanacora, G., and Banasr, M. (2013). From pathophysiology to novel antidepressant drugs: glial contributions to the pathology and treatment of mood disorders. Biol. Psychiatry 73, 1172–1179. doi: 10.1016/j.biopsych.2013.03.032

PubMed Abstract | CrossRef Full Text | Google Scholar

Sapolsky, R. M. (2001). Depression, antidepressants, and the shrinking hippocampus. Proc. Natl. Acad. Sci. U.S.A. 98, 12320–12322.

PubMed Abstract | Google Scholar

Sapolsky, R. M., Romero, L. M., and Munck, A. U. (2000). How do glucocorticoids influence stress responses? Integrating permissive, suppressive, stimulatory, and preparative actions. Endocr. Rev. 21, 55–89.

PubMed Abstract | Google Scholar

Sawicki, C. M., McKim, D. B., Wohleb, E. S., Jarrett, B. L., Reader, B. F., Norden, D. M., et al. (2015). Social defeat promotes a reactive endothelium in a brain region-dependent manner with increased expression of key adhesion molecules, selectins and chemokines associated with the recruitment of myeloid cells to the brain. Neuroscience 302, 151–164. doi: 10.1016/j.neuroscience.2014.10.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Scheinert, R. B., Haeri, M. H., Lehmann, M. L., and Herkenham, M. (2016). Therapeutic effects of stress-programmed lymphocytes transferred to chronically stressed mice. Prog. Neuropsychopharmacol. Biol. Psychiatry 70, 1–7. doi: 10.1016/j.pnpbp.2016.04.010

PubMed Abstract | CrossRef Full Text | Google Scholar

Schnieder, T. P., Trencevska, I., Rosoklija, G., Stankov, A., Mann, J. J., Smiley, J., et al. (2014). Microglia of prefrontal white matter in suicide. J. Neuropathol. Exp. Neurol. 73, 880–890. doi: 10.1097/NEN.0000000000000107

PubMed Abstract | CrossRef Full Text | Google Scholar

Schwartz, M., and Kipnis, J. (2011). A conceptual revolution in the relationships between the brain and immunity. Brain Behav. Immun. 25, 817–819.

PubMed Abstract | Google Scholar

Seidel, A., Arolt, V., Hunstiger, M., Rink, L., Behnisch, A., and Kirchner, H. (1995). Cytokine production and serum proteins in depression. Scand. J. Immunol. 41, 534–538.

Google Scholar

Seidel, A., Arolt, V., Hunstiger, M., Rink, L., Behnisch, A., and Kirchner, H. (1996a). Increased CD56+ natural killer cells and related cytokines in major depression. Clin. Immunol. Immunopathol. 78, 83–85.

Google Scholar

Seidel, A., Arolt, V., Hunstiger, M., Rink, L., Behnisch, A., and Kirchner, H. (1996b). Major depressive disorder is associated with elevated monocyte counts. Acta Psychiatr. Scand. 94, 198–204.

PubMed Abstract | Google Scholar

Setiawan, E., Wilson, A. A., Mizrahi, R., Rusjan, P. M., Miler, L., Rajkowska, G., et al. (2015). Role of translocator protein density, a marker of neuroinflammation, in the brain during major depressive episodes. JAMA Psychiatry 72, 268–275. doi: 10.1001/jamapsychiatry.2014.2427

PubMed Abstract | CrossRef Full Text | Google Scholar

Shatz, C. J. (2009). MHC class I: an unexpected role in neuronal plasticity. Neuron 64, 40–45. doi: 10.1016/j.neuron.2009.09.044

PubMed Abstract | CrossRef Full Text | Google Scholar

Shelton, R. C., Claiborne, J., Sidoryk-Wegrzynowicz, M., Reddy, R., Aschner, M., Lewis, D. A., et al. (2011). Altered expression of genes involved in inflammation and apoptosis in frontal cortex in major depression. Mol. Psychiatry 16, 751–762. doi: 10.1038/mp.2010.52

PubMed Abstract | CrossRef Full Text | Google Scholar

Shen, T. C., Lin, C. L., Liao, C. H., Wei, C. C., Sung, F. C., and Kao, C. H. (2017). Major depressive disorder is associated with subsequent adult-onset asthma: a population-based cohort study. Epidemiol. Psychiatr. Sci. 26, 664–671. doi: 10.1017/S2045796016000664

PubMed Abstract | CrossRef Full Text | Google Scholar

Silverman, M. N., and Sternberg, E. M. (2012). Glucocorticoid regulation of inflammation and its functional correlates: from HPA axis to glucocorticoid receptor dysfunction. Ann. N. Y. Acad. Sci. 1261, 55–63. doi: 10.1111/j.1749-6632.2012.06633.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Simard, M., and Nedergaard, M. (2004). The neurobiology of glia in the context of water and ion homeostasis. Neuroscience 129, 877–896.

PubMed Abstract | Google Scholar

Simpson, E., Worm, M., Soong, W., Blauvelt, A., Eckert, L., Wu, R., et al. (2015). Dupilumab improves patient-reported outcomes (PROs) in a Phase 2 study in adults with moderate-to-severe atopic dermatitis. J. Allergy Clin. Immunol. 135:AB167.

Google Scholar

Slavich, G. M., and Irwin, M. R. (2014). From stress to inflammation and major depressive disorder: a social signal transduction theory of depression. Psychol. Bull. 140, 774–815. doi: 10.1037/a0035302

PubMed Abstract | CrossRef Full Text | Google Scholar

Sluzewska, A., Sobieska, M., and Rybakowski, J. K. (1997). Changes in acute-phase proteins during lithium potentiation of antidepressants in refractory depression. Neuropsychobiology 35, 123–127.

PubMed Abstract | Google Scholar

Sofroniew, M. V., and Vinters, H. V. (2010). Astrocytes: biology and pathology. Acta Neuropathol. 119, 7–35.

Google Scholar

Steiner, J., Bielau, H., Brisch, R., Danos, P., Ullrich, O., Mawrin, C., et al. (2008). Immunological aspects in the neurobiology of suicide: elevated microglial density in schizophrenia and depression is associated with suicide. J. Psychiatr. Res. 42, 151–157.

PubMed Abstract | Google Scholar

Steiner, J., Walter, M., Gos, T., Guillemin, G. J., Bernstein, H. G., Sarnyai, Z., et al. (2011). Severe depression is associated with increased microglial quinolinic acid in subregions of the anterior cingulate gyrus: evidence for an immune-modulated glutamatergic neurotransmission? J. Neuroinflammation 8:94. doi: 10.1186/1742-2094-8-94

PubMed Abstract | CrossRef Full Text | Google Scholar

Stetler, C., and Miller, G. E. (2011). Depression and hypothalamic-pituitary-adrenal activation: a quantitative summary of four decades of research. Psychosom. Med. 73, 114–126. doi: 10.1097/PSY.0b013e31820ad12b

PubMed Abstract | CrossRef Full Text | Google Scholar

Stoll, G., Jander, S., and Schroeter, M. (2002). Detrimental and beneficial effects of injury-induced inflammation and cytokine expression in the nervous system. Adv. Exp. Med. Biol. 513, 87–113.

Google Scholar

Sutcigil, L., Oktenli, C., Musabak, U., Bozkurt, A., Cansever, A., Uzun, O., et al. (2008). Pro-and anti-inflammatory cytokine balance in major depression: effect of sertraline therapy. Clin. Dev. Immunol. 2007:76396. doi: 10.1155/2007/76396

PubMed Abstract | CrossRef Full Text | Google Scholar

Tas, S. W., Vervoordeldonk, M. J., Hajji, N., Schuitemaker, J. H., van der Sluijs, K. F., May, M. J., et al. (2007). Noncanonical NF-kappaB signaling in dendritic cells is required for indoleamine 2,3-dioxygenase (IDO) induction and immune regulation. Blood 110, 1540–1549.

PubMed Abstract | Google Scholar

Tavares, R. G., Tasca, C. I., Santos, C. E., Alves, L. B., Porciuncula, L. O., Emanuelli, T., et al. (2002). Quinolinic acid stimulates synaptosomal glutamate release and inhibits glutamate uptake into astrocytes. Neurochem. Int. 40, 621–627.

PubMed Abstract | Google Scholar

Thayer, J. F., and Sternberg, E. M. (2010). Neural aspects of immunomodulation: focus on the vagus nerve. Brain Behav. Immun. 24, 1223–1228. doi: 10.1016/j.bbi.2010.07.247

PubMed Abstract | CrossRef Full Text | Google Scholar

Toben, C., and Baune, B. T. (2015). An act of balance between adaptive and maladaptive immunity in depression: a role for T lymphocytes. J. Neuroimmune Pharmacol. 10, 595–609. doi: 10.1007/s11481-015-9620-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Torres-Platas, S. G., Cruceanu, C., Chen, G. G., Turecki, G., and Mechawar, N. (2014). Evidence for increased microglial priming and macrophage recruitment in the dorsal anterior cingulate white matter of depressed suicides. Brain Behav. Immun. 42, 50–59. doi: 10.1016/j.bbi.2014.05.007

PubMed Abstract | CrossRef Full Text | Google Scholar

Tuglu, C., Kara, S. H., Caliyurt, O., Vardar, E., and Abay, E. (2003). Increased serum tumor necrosis factor-alpha levels and treatment response in major depressive disorder. Psychopharmacology 170, 429–433.

Google Scholar

Tyring, S., Bagel, J., Lynde, C., Klekotka, P., Thompson, E. H. Z., Gandra, S. R., et al. (2013). Patient-reported outcomes in moderate-to-severe plaque psoriasis with scalp involvement: results from a randomized, double-blind, placebo-controlled study of etanercept. J. Eur. Acad. Dermatol. Venereol. 27, 125–128.

PubMed Abstract | Google Scholar

Tyring, S., Gottlieb, A., Papp, K., Gordon, K., Leonardi, C., Wang, A., et al. (2006). Etanercept and clinical outcomes, fatigue, and depression in psoriasis: double-blind placebo-controlled randomised phase III trial. Lancet 367, 29–35.

PubMed Abstract | Google Scholar

Uher, R., Tansey, K. E., Dew, T., Maier, W., Mors, O., Hauser, J., et al. (2014). An inflammatory biomarker as a differential predictor of outcome of depression treatment with escitalopram and nortriptyline. Am. J. Psychiatry 171, 1278–1286. doi: 10.1176/appi.ajp.2014.14010094

PubMed Abstract | CrossRef Full Text | Google Scholar

Van Lieshout, R. J., Bienenstock, J., and MacQueen, G. M. (2009). A review of candidate pathways underlying the association between asthma and major depressive disorder. Psychosom. Med. 71, 187–195. doi: 10.1097/PSY.0b013e3181907012

PubMed Abstract | CrossRef Full Text | Google Scholar

Vezzani, A., and Viviani, B. (2015). Neuromodulatory properties of inflammatory cytokines and their impact on neuronal excitability. Neuropharmacology 96, 70–82. doi: 10.1016/j.neuropharm.2014.10.027

PubMed Abstract | CrossRef Full Text | Google Scholar

Wachholz, S., Esslinger, M., Plumper, J., Manitz, M. P., Juckel, G., and Friebe, A. (2016). Microglia activation is associated with IFN-alpha induced depressive-like behavior. Brain Behav. Immun. 55, 105–113. doi: 10.1016/j.bbi.2015.09.016

PubMed Abstract | CrossRef Full Text | Google Scholar

Wake, H., Moorhouse, A. J., Miyamoto, A., and Nabekura, J. (2013). Microglia: actively surveying and shaping neuronal circuit structure and function. Trends Neurosci. 36, 209–217. doi: 10.1016/j.tins.2012.11.007

PubMed Abstract | CrossRef Full Text | Google Scholar

Warner-Schmidt, J. L., Vanover, K. E., Chen, E. Y., Marshall, J. J., and Greengard, P. (2011). Antidepressant effects of selective serotonin reuptake inhibitors (SSRIs) are attenuated by antiinflammatory drugs in mice and humans. Proc. Natl. Acad. Sci. U.S.A. 108, 9262–9267. doi: 10.1073/pnas.1104836108

PubMed Abstract | CrossRef Full Text | Google Scholar

Weber, D. J., Allette, Y. M., Wilkes, D. S., and White, F. A. (2015). The HMGB1-RAGE inflammatory pathway: implications for brain injury-induced pulmonary dysfunction. Antioxid. Redox. Signal. 23, 1316–1328. doi: 10.1089/ars.2015.6299

PubMed Abstract | CrossRef Full Text | Google Scholar

Whitney, N. P., Eidem, T. M., Peng, H., Huang, Y., and Zheng, J. C. (2009). Inflammation mediates varying effects in neurogenesis: relevance to the pathogenesis of brain injury and neurodegenerative disorders. J. Neurochem. 108, 1343–1359. doi: 10.1111/j.1471-4159.2009.05886.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Wohleb, E. S., Hanke, M. L., Corona, A. W., Powell, N. D., Stiner, L. M., Bailey, M. T., et al. (2011). beta-Adrenergic receptor antagonism prevents anxiety-like behavior and microglial reactivity induced by repeated social defeat. J. Neurosci. 31, 6277–6288. doi: 10.1523/JNEUROSCI.0450-11.2011

PubMed Abstract | CrossRef Full Text | Google Scholar

Wohleb, E. S., Powell, N. D., Godbout, J. P., and Sheridan, J. F. (2013). Stress-induced recruitment of bone marrow-derived monocytes to the brain promotes anxiety-like behavior. J. Neurosci. 33, 13820–13833. doi: 10.1523/JNEUROSCI.1671-13.2013

CrossRef Full Text | Google Scholar

Wohleb, E. S., Terwilliger, R., Duman, C. H., and Duman, R. S. (2018). Stress-induced neuronal colony stimulating factor 1 provokes microglia-mediated neuronal remodeling and depressive-like behavior. Biol. Psychiatry 83, 38–49. doi: 10.1016/j.biopsych.2017.05.026

PubMed Abstract | CrossRef Full Text | Google Scholar

Wong, M. L., Inserra, A., Lewis, M. D., Mastronardi, C. A., Leong, L., Choo, J., et al. (2016). Inflammasome signaling affects anxiety- and depressive-like behavior and gut microbiome composition. Mol. Psychiatry 21, 797–805. doi: 10.1038/mp.2016.46

PubMed Abstract | CrossRef Full Text | Google Scholar

Wood, J. C., Jackson, J. S., Jakubs, K., Chapman, K. Z., Ekdahl, C. T., Kokaia, Z., et al. (2011). Functional integration of new hippocampal neurons following insults to the adult brain is determined by characteristics of pathological environment. Exp. Neurol. 229, 484–493. doi: 10.1016/j.expneurol.2011.03.019

PubMed Abstract | CrossRef Full Text | Google Scholar

Wu, T. Y., Liu, L., Zhang, W., Zhang, Y., Liu, Y. Z., Shen, X. L., et al. (2015). High-mobility group box-1 was released actively and involved in LPS induced depressive-like behavior. J. Psychiatr. Res. 64, 99–106. doi: 10.1016/j.jpsychires.2015.02.016

PubMed Abstract | CrossRef Full Text | Google Scholar

Yirmiya, R., Pollak, Y., Barak, O., Avitsur, R., Ovadia, H., Bette, M., et al. (2001). Effects of antidepressant drugs on the behavioral and physiological responses to lipopolysaccharide (LPS) in rodents. Neuropsychopharmacology 24, 531–544.

Google Scholar

Zhang, Y., Welzig, C. M., Picard, K. L., Du, C., Wang, B., Pan, J. Q., et al. (2014). Glycogen synthase kinase-3beta inhibition ameliorates cardiac parasympathetic dysfunction in type 1 diabetic Akita mice. Diabetes Metab. Res. Rev. 63, 2097–2113. doi: 10.2337/db12-1459

PubMed Abstract | CrossRef Full Text | Google Scholar

Zheng, X., Ma, S., Kang, A., Wu, M., Wang, L., Wang, Q., et al. (2016). Chemical dampening of Ly6C(hi) monocytes in the periphery produces anti-depressant effects in mice. Sci. Rep. 6:19406. doi: 10.1038/srep19406

PubMed Abstract | CrossRef Full Text | Google Scholar

Ziv, Y., and Schwartz, M. (2008). Orchestrating brain-cell renewal: the role of immune cells in adult neurogenesis in health and disease. Trends Mol. Med. 14, 471–478. doi: 10.1016/j.molmed.2008.09.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Zorrilla, E. P., Luborsky, L., McKay, J. R., Rosenthal, R., Houldin, A., Tax, A., et al. (2001). The relationship of depression and stressors to immunological assays: a meta-analytic review. Brain Behav. Immun. 15, 199–226.

PubMed Abstract | Google Scholar

Keywords: GSK3, innate immune response, adaptive immune response, depression, NF-κB

Citation: Medina-Rodriguez EM, Lowell JA, Worthen RJ, Syed SA and Beurel E (2018) Involvement of Innate and Adaptive Immune Systems Alterations in the Pathophysiology and Treatment of Depression. Front. Neurosci. 12:547. doi: 10.3389/fnins.2018.00547

Received: 02 May 2018; Accepted: 20 July 2018;
Published: 17 August 2018.

Edited by:

Abed N. Azab, Ben-Gurion University of the Negev, Israel

Reviewed by:

Manish Kumar Jha, University of Texas Southwestern Medical Center, United States
Carla Mucignat, Università degli Studi di Padova, Italy

Copyright © 2018 Medina-Rodriguez, Lowell, Worthen, Syed and Beurel. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Eléonore Beurel, ebeurel@miami.edu

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.