The American Psychiatric Association (APA) has updated its Privacy Policy and Terms of Use, including with new information specifically addressed to individuals in the European Economic Area. As described in the Privacy Policy and Terms of Use, this website utilizes cookies, including for the purpose of offering an optimal online experience and services tailored to your preferences.

Please read the entire Privacy Policy and Terms of Use. By closing this message, browsing this website, continuing the navigation, or otherwise continuing to use the APA's websites, you confirm that you understand and accept the terms of the Privacy Policy and Terms of Use, including the utilization of cookies.

×

Abstract

This article will review the evidence for the “inflammatory hypothesis of depression.” The authors summarize the literature suggesting that immune-mediated changes contribute to the pathophysiology of MDD; describe potential underlying mechanisms; and discuss translational targets, including proinflammatory cytokines and cytokine-signaling pathways, for the development of novel antidepressants.

Introduction

For centuries, the view that the mind can shape our susceptibility to illness has captured the artistic imagination. Robert Dantzer, a pioneer in the field of psychoneuroimmunology, points to the following illustrative quote from the author, Franz Kafka, who suffered from tuberculosis (1): “It is all my mind that is ill; the affection of my lungs is nothing else than the spillover of my ill mind” (2). The reverse view, however, that the immune system may modulate the brain—and its emotional and cognitive functioning—has only recently been explored. In the 1990s, a research group in the Department of Psychiatry at the Stuivenberg in Antwerp, Belgium, found evidence of immune activation in major depressive disorder (MDD), including elevated levels of proinflammatory cytokines and increased numbers of peripheral leukocytes, monocytes, and T cells (3). Their findings have since been replicated and expanded by multiple independent investigators (4), giving rise to the hypothesis that “inflammation has a role in at least some significant sub-population of depressed patients” (5).

Evidence linking mood and immune systems

A brief review of the inflammatory response

When pathogens invade the body, our first line of defense consists of phagocytic cells (monocytes, tissue macrophages, and liver Kupffer cells). Phagocytes possess Toll-like receptors, an ancient family of receptors, which recognize molecules that are shared by pathogens but are distinguishable from host cells (referred to as pathogen-associated molecular patterns [PAMPs]). Activated phagocytes will ingest pathogens for internal degradation; Toll-like receptors will also initiate cellular activation through increased DNA binding to the transcription factor nuclear factor-κB (NF-κB), which upregulates the expression of proinflammatory cytokines such as interferon 1 (IF-1) and tumor necrosis factor alpha (TNF-alpha).

Cytokines then mediate diverse components of the inflammatory response: cytokines, known as chemokines, attract lymphocytes to the area of infection and facilitate the flow of antigen-presenting cells to nearby lymph nodes; other cytokines trigger fever, increase blood vessel permeability, and enhance nociception. This response underlies the cardinal features of infection: dolor (pain), calor (heat), rubor (redness), and tumor (swelling) (6).

Sickness behavior

In addition to recruiting local and systemic inflammatory responses, proinflammatory cytokines can also act on the brain to produce dramatic changes in behavior, termed “sickness behavior.” Symptoms of sickness behavior—malaise, anhedonia, anorexia, impaired concentration, and sleep disturbance—are easy to recognize from our own experience of sickness, and closely resemble the neurovegetative features of depression. In rodent models, administration of proinflammatory cytokines or cytokine inducers have been associated with depression-like behavioral changes, including increased immobility time on the forced swim test, anhedonia, sleep disruption, and anorexia (4). These symptoms can be reversed through acute treatment with an anti-inflammatory cytokine (IL-10) or cytokine antagonist (IL-1RA) or chronic treatment with a serotonergic antidepressant (7). Additionally, mice lacking the enzyme, caspase 1, required to synthesize IL-1, exhibit reduced sickness behavior (8), and deletion of genes for the TNF-alpha receptors has been associated with antidepressant effects (9).

In humans, both biological inducers, such as microbial pathogens, and psychosocial stressors can activate the inflammatory response (7). For example, in healthy volunteers, public speaking and mental arithmetic tasks have been found to increase DNA binding to the inflammatory transcription factor NF-κB (10). In a study by Brydon and colleagues (11), healthy participants were injected with either typhoid vaccine (previously shown to induce negative mood states in healthy subjects and to increase circulating levels of inflammatory cytokines) or placebo and then exposed to rest or stress conditions. IL-6 levels were highest in the vaccine/stress group; these levels also correlated with higher levels of negative mood, suggesting a potential synergy between psychosocial and immune stressors. It is not uncommon to see in clinical practice patients who experience transient worsenings of mood in the context of upper respiratory infections and an otherwise stable response to continued antidepressant therapy. Such worsenings are typically transient and end once the infection remits.

Interferon model

Perhaps the most compelling data linking sickness behavior and depression stems from studies of cytokine therapy. Approximately 20% to 50% of patients treated with the proinflammatory cytokine, interferon-alpha, for hepatitis C or malignant melanoma will develop a clinical depression (4), and these rates can be dramatically reduced (approximately fourfold) following pretreatment with the antidepressant paroxetine (12).

Further analysis of patients receiving interferon-alpha therapy has revealed an acute phase neurovegetative syndrome, characterized by prominent malaise, sleep disturbance, and psychomotor slowing, followed by a mood and cognitive syndrome (typically 1–3 months post treatment initiation), characterized by anxiety, depressed mood, and disturbances in concentration and memory (see Figure 1) (13). Compared with the neurovegetative syndrome, the mood and cognitive syndrome is more responsive to antidepressants and more prevalent in patients with intrinsic vulnerability factors. For example, using the Montgomery-Asberg Depression Rating Scale (MADRS), Capuron and Ravaud found that baseline mood ratings in patients with cancer could be used to predict the intensity of depressive symptoms induced by interferon-alpha (14); the same group also found that hypothalamic pituitary adrenal (HPA) axis hyperresponsiveness following the initial injection of interferon-alpha was associated with increased risk of developing depression during subsequent treatment (15). Additionally, researchers have now identified depressive subtypes among different cohorts receiving interferon-alpha therapy. A recent study found that patients treated with interferon-alpha for hepatitis C were more likely to develop depression with mixed features, including prominent irritability, anxiety, and dysphoria (16). This finding, perhaps, adds ecological validity to the interferon-induction model of depression, recapitulating the pleomorphic phenomenology of MDD found in clinical settings (17).

Figure 1. Neuropsychiatric Symptoms Seen With Interferon-Alpha Therapy

Adapted from Capuron L, Miller AH: Immune system to brain signaling: neuropsychopharmacological implications. Pharmacol Ther 2011; 130:226-238. Copyright © 2011 Elsevier Inc. Reprinted with permission.

MDD and biomarkers of inflammation

The behavioral similarities between sickness behavior and depression have catalyzed research exploring potential immunological mediators of MDD. Proinflammatory cytokines, acute phase reactant proteins, chemokines, and adhesion molecules have been shown to be increased in individuals with MDD (1820) or risk factors for MDD (stress, medical illness, obesity, sedentary lifestyle, diet, insomnia, social isolation, low socioeconomic status) (21) compared with healthy comparison subjects. High levels of inflammatory biomarkers have also been positively correlated with depression severity and negatively correlated with treatment resolution (22). There is also increasing evidence that elevations of proinflammatory cytokines represent a common factor underlying the bidirectional influence between major depressive disorder and chronic inflammatory disorders, such as cardiovascular disease, metabolic syndrome, and obesity (23, 24).

Mechanisms

The following section will review how inflammatory processes in the central nervous system (CNS) may affect the pathophysiology of MDD.

Peripheral cytokine signals can access the brain

Historically, the CNS was viewed as an immunologically privileged site, sequestered from the immune system by the blood-brain barrier (25). However, in 1984, Ericsson discovered that peripheral IL-1 can stimulate catecholamine neurons projecting from the medulla to the paraventricular nucleus of the hypothalamus; additionally, the vagus nerve was shown to transmit peripheral cytokines signals directly to the brain (referred to as the neural pathway) (26). Several other neuroimmune communication pathways have also been elucidated, including a dynamic active transport system for cytokines to cross the blood-brain barrier (27, 28) (the cellular pathway); and the possibility that peripheral cytokines can penetrate circumventricular organs, leaky regions of the blood-brain barrier, and then stimulate neural elements projecting to the CNS (the humoral pathway) (29). Within the CNS, cytokine networks (consisting of microglia, astrocytes, and neurons) can influence behavior and mood through multiple mechanisms, including the modulation of neurotransmitter systems, neurotrophic factors, neuroendocrine activity, and mood-related neurocircuitry.

Neurotransmitters

Serotonin

Beginning with research performed in the 1960s by Coppen, who found that MAOI efficacy could be enhanced by the addition of the serotonin precursor tryptophan, and later Ashcroft, who found that patients with MDD had lower CSF levels of the serotonin metabolite 5-hydroxyindoleacetic acid, serotonin has been viewed as a central neurotransmitter in the pathophysiology and treatment of MDD (30). In recent decades, the serotonin deficit model of MDD (31) has evolved to include abnormalities of serotonin receptors, particularly 5HT-1A (32), and dysregulation of the molecular and cellular mechanisms downstream of serotonin binding (33). Additionally, a novel class of antidepressants (e.g. vilazodone), termed multimodal serotonergic agents, has emerged to target specific serotonin subsystems (34).

There is now strong evidence that high levels of proinflammatory cytokines are associated with decreased serotonergic neurotransmission. Proinflammatory cytokines can induce the enzyme, indoleamine 2,3-dioxygenase (IDO), which shunts tryptophan metabolism toward the production of kynurenine and quinolinic acid rather than serotonin (35). Similar to the serotonin depletion paradigm (36), studies of patients receiving interferon-alpha—an inducer of IDO—reveal a positive correlation between the development of depression and decreases in serum tryptophan, increases in kynurenine, and increases in the kynurenine to tryptophan ratio (a proxy for low CSF levels of serotonin) (37, 38). Interferon-alpha also stimulates p38 mitogen-activated protein kinase (MAPK) pathways, increasing expression and activity of the serotonin transporter, and, therefore, decreasing extracellular levels of serotonin (39, 40). Finally, two independent research groups have found that a polymorphism of the serotonin transporter gene (5-HTTLPR) is associated with interferon-alpha-induced depression severity (41, 42).

Dopamine

There is increasing evidence that the pathophysiology of depression involves abnormal functioning of the dopaminergic, cortico-basal ganglia reward circuitry (43). Findings of dopamine dysregulation in MDD populations include reduced concentrations of the dopamine metabolite, homovanillic acid (HVA), in cerebrospinal fluid (44, 45), reduced l-dopa uptake across the blood-brain barrier (46), reduced density of striatal dopamine transporters (47), and increased striatal binding to D2/D3 receptors (48, 49), with several conflicting studies also reported (50, 51).

Multiple animal studies have now shown that proinflammatory cytokines can decrease concentrations of dopamine (DA) and its precursor, tetrahydrobiopterin (BH4), in limbic regions of the brain (5254). In a remarkable study using a mouse model, Qin et al. demonstrated that single exposure to systemic inflammation increased TNF-alpha production both peripherally and centrally, creating a cycle of neuroinflammation that persisted for 10 months and caused degeneration of nearly half (47%) of dopaminergic neurons in the substantia nigra (55). Interferon-alpha induced activation of the p38 MAPK pathway has also been linked to the severity of depressive episodes during cytokine therapy (56) and a putative mechanism of action, since activation of MAPK causes increased dopamine uptake, dropping synaptic levels (57).

Glutamate

Glutamate is an amino acid neurotransmitter, highly abundant in the central nervous system, which primarily functions to modulate excitatory neurotransmission. In 1990, Trullas and Skolnick first proposed the glutamate hypothesis of depression based on their findings that antagonism of NMDA receptors produced antidepressant-like effects and reversed behavioral deficits in mice exposed to inescapable stressors (58). Over the past two decades, this hypothesis has expanded from a focus on long-term potentiation and excitotoxicity to include multiple morphological and molecular mechanisms of synaptic plasticity, including neurogenesis (59). Increased levels of glutamate in the frontal cortex (60) and increased glutamatergic neurotransmission in the mesolimbic system (59) have recently been implicated in the pathophysiology of MDD.

Proinflammatory cytokines can increase glutamatergic transmission through multiple mechanisms: TNF-alpha has been shown to suppress glutamate transport and to reduce expression of the glutamate transporters, EAAT-1 and EAAT-2, leading to higher extracellular levels of glutamate (61); inflammatory cytokines also stimulate astrocytes to release glutamate (62); and quinolinic acid, a metabolite formed by activation of the IDO pathway (above) functions as a strong agonist of the glutamatergic N-methyl-d-aspartate receptor (63). Collectively, these changes increase net glutamate transmission, possibly contributing to excitotoxicity and loss of glial elements (64) relevant to the pathophysiology of MDD (65).

Neurogenesis

A growing body of data suggests that changes in synaptic plasticity, defined as changes in the number of synapses and signal transmission through a synapse, and impairments in neurogenesis contribute to the pathophysiology of MDD. In animal models, brain-derived neurotrophic factor (BDNF) signaling in the hippocampus decreases under chronic stress conditions, and increases following ECT, or sustained treatment with several pharmacodynamically distinct antidepressants (SSRIs, NERIs, MAOIs) (66). Additionally, direct infusion of BDNF into the rodent hippocampus has been shown to induce neurogenesis and produce antidepressant-like effects (67, 68). Recently, Tfilin and colleagues demonstrated that intracerebral infusion of mesenchymal stem cells also resulted in differentiation of new hippocampal neurons, and reduced behavioral measures of depression, such as immobility time on the forced swim test (69). In humans, atrophy of the hippocampus and other forebrain regions in patients with MDD have been linked to decrements in neurotrophic factors that also regulate homeostatic plasticity (66, 70).

Cytokines have a physiologic role in modulating neural development and synaptic plasticity in humans (64). However, pathological elevation of proinflammatory cytokines has been shown to disrupt these processes and to produce depression-like behaviors (71). In a particularly novel study, Ben Menachem-Zidon and colleagues established that chronically isolated mice demonstrated a significant elevation in hippocampal IL-1 associated with weight loss, cognitive impairment, and decreased hippocampal neurogenesis. They then delivered an IL-1 receptor antagonist (IL-1ra) into the brain, using transplantation of neural precursor cells (NPCs), obtained from neonatal mice with transgenic overexpression of IL-1ra; they found that, after 4 weeks, the transplanted mice demonstrated restored hippocampal neurogenesis and reversal of depressive-like symptoms, compared with isolated mice transplanted with wild type cells or sham operated (72). Similar results have been obtained by independent investigators, who found that blockade of the IL-1β receptor (also using IL-1ra in a mouse model) reverses the antineurogenic and anhedonic effects of stress (73). Additionally, a series of animal studies have demonstrated that IL-1 effects on synaptic plasticity follow an inverted U shaped curve: constitutive low levels of IL-1 are necessary for memory formation and maintenance of synaptic strength; however, the elevated levels found in proinflammatory states decrease cell proliferation in the hippocampus and impair memory formation (74, 75). Finally, as described in the glutamate section above, induction of IDO by proinflammatory cytokines increases concentrations of the glutamate agonist, quinolinic acid, which can inhibit BDNF expression via stimulation of extrasynaptic NMDA receptors (76).

Neuroendocrine function

Abnormalities of the HPA axis have been consistently observed in subsets of depressed patients (77). The physiologic stress response begins with activation of the sympathetic nervous system, which (within seconds) releases the catecholamines epinephrine and norepinephrine into the portal circulation. A slower activation of the HPA axis then follows, with secretion of corticotrophin-releasing hormone (CRH) by the hypothalamus stimulating the pituitary to secrete adrenocorticotropic hormone (ACTH), which activates the adrenals to release glucocorticoids (cortisol) (78). Unbound cortisol can cross the blood-brain barrier and bind to the glucocorticoid receptor, leading to changes in gene transcription that can acutely upregulate immune function prior to pathogen exposure (79). MDD, however, is characterized by both HPA hyperactivity and glucocorticoid receptor insensitivity, which results in impaired negative feedback inhibition of the stress response (77).

Consistent with the stress-diathesis model of MDD (80), proinflammatory cytokines may preferentially activate the HPA axis in patients vulnerable to depression. For example, in a study of patients receiving IFN-alpha therapy for malignant melanoma, Capuron and colleagues demonstrated that patients who developed MDD during treatment exhibited a higher production of ACTH and cortisol following initial infusion of IFN-alpha compared with those who did not develop a major depressive episode (15). In contrast, chronic exposure to elevated levels of proinflammatory cytokines has been associated with depressed mood and HPA axis abnormalities characteristic of MDD, including a blunted ACTH response to intravenous administration of CRH, flattened circadian cortisol variation, and reduced glucocorticoid receptor sensitivity (8184). Glucocorticoid resistance in MDD has specifically been correlated with high levels of IL-1 (85) and TNF-alpha (86). It is possible that this relationship is mediated by increased activity of the cytokine signaling pathways, p38 MAPK and NF-κB, which can reduce glucocorticoid receptor activity through multiple intracellular mechanisms (87).

Mood neurocircuitry

Neuroimaging studies of depressed patients have reported reductions in gray matter volume and glial density in the prefrontal cortex and the hippocampus (88); correlations between increased activity within the amygdala and subgenual cingulate cortex and dysphoric emotions (89); and abnormal functioning of the dopaminergic cortico-basal ganglia reward circuitry (43). Additionally, deep brain stimulation of the subgenual cingulate cortex has been shown to produce antidepressant effects in patients with treatment-resistant depression (90). Unfortunately, many published neuroimaging findings have been inconsistent (91) and limited by nonoverlapping study methodologies, heterogeneous study populations, and failure to establish causation (33).

There is now an emerging neuroimaging literature relating proinflammatory cytokines to specific changes in neurocircuitry (5). Two studies suggest that IFN-alpha induces hyperactivity in the basal ganglia (thought to represent increased oscillatory bursts from depleted dopamine neurons), which parallels the metabolic changes found in Parkinson's disease (92, 93). Remarkably, patients treated with interferon-alpha have been found to develop Parkinson-like symptoms, which were relieved by treatment with levodopa (94). Additionally, multiple studies have shown increased activation of the dorsal anterior cingulate cortex (dACC) associated with elevated levels of proinflammatory cytokines (9597); the dACC has been hypothesized to function as a “neural alarm system” for threatening social stimuli (98), and has also been found to be hyperactive in patients with high ratings of neuroticism (99), a robust risk factor for MDD (100).

Emerging treatments

Pharmacological interventions

A number of antidepressant medications have shown anti-inflammatory properties: for example bupropion lowers production of TNF-alpha and interferon-gamma in mice (101), hypericum extracts have produced antiinflammatory activity in the rat paw edema test (102), and S-adenosylmethionine suppresses TNF-alpha production (103). In addition, there is growing evidence that immune-modulating medications may translate into a novel class of antidepressants. In a study of patients with psoriasis, Tyring and colleagues found that individuals randomly assigned to treatment with the TNF-alpha antagonist, etanercept, had significantly greater improvements in depressive symptoms (measured by the HAM-D-17) compared with controls; changes in core features of depression were only weakly correlated with objective measures of skin clearance and joint pain (104). In a proof-of-concept study, Müller and colleagues found that depressed patients randomly assigned to treatment with reboxetine combined with the COX-2 inhibitor celecoxib showed greater improvements in depression (measured by the HAM-D-17) compared with controls treated with reboxetine alone (105). Finally, a large randomized controlled trial of the TNF-alpha antagonist infliximab for treatment-resistant depression has recently been completed, suggesting that TNF antagonism may have preferential antidepressant effects for patients with high baseline levels of proinflammatory cytokines (106). Other potentially promising pharmacological agents include inhibitors of the inflammatory cytokines, IL-1, IL-6, and interferon-alpha; modulators of the cytokine signaling pathways NF-kB, p38 MAPK, and STAT5; and augmenters or agonists of the anti-inflammatory cytokines IL-10 and TGF (5).

Possible preferential responders

Treatments specifically targeting inflammation may not be appropriate for every patient with depression. However, MDD patients with elevated inflammatory biomarkers (106) and/or comorbid chronic inflammatory diseases, such as autoimmune diseases, cardiovascular diseases, and obesity, may be more likely to respond to immune-modulating agents. For example, Howren and colleagues found positive correlations between body mass index (BMI) and circulating levels of IL-6 and CRP (107); and a recent study by Capuron et al. suggests the higher prevalence of neurovegative symptoms in patients with metabolic syndrome is associated with increased inflammation (108). Depressed patients experiencing acute psychosocial stress, and/or those with a history of child abuse, both of which are associated with hypersensitized innate immunity, may constitute additional subgroups of preferential responders (109, 110). Finally, insomniacs and men who are socially isolated also tend to have higher levels of inflammation and could, therefore, potentially derive greater mood benefits from CNS targeted anti-inflammatory compounds (111113).

Summary

This article is intended to provide a framework for understanding the inflammatory hypothesis of depression. We have covered the following key points:

•. 

The acute immune response is mediated by diverse cytokines and cytokine signaling pathways

•. 

Proinflammatory cytokines can also act acutely on the brain to produce “sickness behavior,” characterized by symptoms that closely resemble the neurovegetative features of MDD

•. 

The prevalence and phenomenology of depression associated with interferon-alpha therapy for medical illnesses have provided a human model of cytokine-induced MDD; research suggests that pretreatment with SSRIs has protective effects

•. 

Increased inflammatory biomarkers are associated with MDD and may contribute to the high comorbidity between MDD and chronic inflammatory diseases

•. 

Proinflammatory cytokines may contribute to the pathophysiology of MDD, and have been shown to modulate neurotransmitter systems, neurotrophic factors, neuroendocrine activity, and mood-related neurocircuitry

•. 

Immunological agents targeting proinflammatory cytokines and cytokine signaling pathways may represent a novel class of antidepressants

Address correspondence to Dr. Soskin, Depression Clinical and Research Program, Massachusetts General Hospital, One Bowdoin Square – 6th Floor, Boston, MA, 02114; e-mail:

Author Information and CME Disclosure

David P. Soskin, M.D., Depression Clinical and Research Program, Massachusetts General Hospital, Harvard Medical School, Boston, Mass.

Clair Cassiello, B.A., Depression Clinical and Research Program, Massachusetts General Hospital, Harvard Medical School, Boston, Mass.

Oren Isacoff, B.A., Perelman School of Medicine at the University of Pennsylvania, Philadelphia, Pa.

Maurizio Fava, M.D., Depression Clinical and Research Program, Massachusetts General Hospital, Harvard Medical School, Boston, Mass.

David Soskin, Clair Cassiello, and Oren Isacoff report no competing interests.

Dr. Fava reports the following lifetime disclosures: Research Support: Abbott, Alkermes, Aspect Medical Systems, AstraZeneca, BioResearch, BrainCells, Bristol-Myers Squibb, CeNeRx BioPharma, Cephalon, Clinical Trials Solutions, Clintara, Covance, Covidien, Eli Lilly, ElMindA, EnVivo, Euthymics Bioscience, Forest, Ganeden Biotech, GlaxoSmithKiine, Icon Clinical Research, i3 Innovus/ Ingenix, Johnson & Johnson, Lichtwer Pharma, Lorex, NARSAD, NCCAM, NIDA, NIMH, Novartis, Organon, PamLab, Pfizer, Pharmavite, Photothera, Roche, RCT Logic, Sanofi-Aventis, Shire, Solvay, Synthelabo, Wyeth-Ayerst; Advisory/Consulting: Abbott, Affectis, Alkermes, Amarin Pharma, Aspect Medical Systems, AstraZeneca, Auspex, Bayer, Best Practice Project Management, BioMarin Pharmaceuticals, Biovail Corporation, BrainCells Inc, Bristol-Myers Squibb, CeNeRx BioPharma, Cephalon, Clinical Trials Solutions, CNS Response, Compellis, Cypress DiagnoSearch Life Sciences, Dinippon Sumitomo, Edgemont, Eisai , Eli Lilly, EnVivo, ePharmaSolutions, EPIX, Euthymics Bioscience, Fabre-Kramer, Forest, GenOmind, GlaxoSmithKiine, Grunenthal GmbH, i3 lnnovus/lngenis, Janssen,Jazz, Johnson & Johnson, Knoll, Labopharm, Lorex, Lundbeck, MedAvante, Inc., Merck, MSI Methylation Sciences, Naurex, Neuronetics, NextWave Pharmaceuticals, Novartis, Nutrition 21, Orexigen Therapeutics, Organon, Otsuka, PamLab, Pfizer, PharmaStar, Pharmavite, PharmoRx Therapeutics, Precision Human Biolaboratory, Prexa Pharmaceuticals, Puretech Ventures, PsychoGenies, Psylin Neurosciences, Rexahn Pharmaceuticals, Ridge Diagnostics, Inc., Roche, RCT Logic, Sanofi Aventis, Sepracor, Servier Laboratories, Schering-Plough, Solvay, Somaxon, Somerset, Sunovion, Supernus, Synthelabo, Takeda, Tal Medical, Tetragenex, TransForm, Transcept, Vanda. Speaking/Publishing: Adamed, Advanced Meeting Partners, American Psychiatric Association, American Society of Clinical Psychopharmacology, AstraZeneca, Belvoir Media Group, Boehringer lngelheim, Bristol-Myers Squibb, Cephalon, CME Institute/Physicians Postgraduate Press, Eli Lilly, Forest, GlaxoSmithKiine, lmedex, MGH Psychiatry Academy/Primedia, MGH Psychiatry Academy/Reed Elsevier, Novartis, Organon, Pfizer, PharmaStar, United BioSource,Corp., Wyeth-Ayerst Laboratories; Equity holdings: Compellis; Other income: patent for Sequential Parallel Comparison Design, patent application for a combination of azapirones and bupropion in major depressive disorder; copyright royalties for the MGH Cognitive & Physical Functioning Questionnaire, Sexual Functioning Inventory, Antidepressant Treatment Response Questionnaire, Discontinuation-Emergent Signs & Symptoms, and SAFER; patent for research and licensing of SPCD with RCT Logic, Lippincott, Wolkers Kluwer, and World Scientific Publishing.

References

1 Dantzer R: Depression and inflammation: an intricate relationship. Biol Psychiatry 2012; 71:4–5CrossrefGoogle Scholar

2 Kafka F, Haas W, Jesenská M: Letters to Milena. New York, Schocken Books, 1954Google Scholar

3 Maes M, Smith R, Scharpe S: The monocyte-T-lymphocyte hypothesis of major depression. Psychoneuroendocrinology 1995; 20:111–116CrossrefGoogle Scholar

4 Miller AH, Maletic V, Raison CL: Inflammation and its discontents: the role of cytokines in the pathophysiology of major depression. Biol Psychiatry 2009; 65:732–741CrossrefGoogle Scholar

5 Haroon E, Raison CL, Miller AH: Psychoneuroimmunology meets neuropsychopharmacology: translational implications of the impact of inflammation on behavior. Neuropsychopharmacology 2012; 37:137–162CrossrefGoogle Scholar

6 Alberts B: Molecular biology of the cell, 5th ed. New York, Garland Science, 2008Google Scholar

7 Dantzer R, O’Connor JC, Freund GG, Johnson RW, Kelley KW: From inflammation to sickness and depression: when the immune system subjugates the brain. Nat Rev Neurosci 2008; 9:46–56CrossrefGoogle Scholar

8 Mastronardi C, Whelan F, Yildiz OA, Hannestad J, Elashoff D, McCann SM, Licinio J, Wong ML: Caspase 1 deficiency reduces inflammation-induced brain transcription. Proc Natl Acad Sci USA 2007; 104:7205–7210CrossrefGoogle Scholar

9 Simen BB, Duman CH, Simen AA, Duman RS: TNFalpha signaling in depression and anxiety: behavioral consequences of individual receptor targeting. Biol Psychiatry 2006; 59:775–785CrossrefGoogle Scholar

10 Bierhaus A, Wolf J, Andrassy M, Rohleder N, Humpert PM, Petrov D, Ferstl R, von Eynatten M, Wendt T, Rudofsky G, Joswig M, Morcos M, Schwaninger M, McEwen B, Kirschbaum C, Nawroth PP: A mechanism converting psychosocial stress into mononuclear cell activation. Proc Natl Acad Sci USA 2003; 100:1920–1925CrossrefGoogle Scholar

11 Brydon L, Walker C, Wawrzyniak A, Whitehead D, Okamura H, Yajima J, Tsuda A, Steptoe A: Synergistic effects of psychological and immune stressors on inflammatory cytokine and sickness responses in humans. Brain Behav Immun 2009; 23:217–224CrossrefGoogle Scholar

12 Musselman DL, Lawson DH, Gumnick JF, Manatunga AK, Penna S, Goodkin RS, Greiner K, Nemeroff CB, Miller AH: Paroxetine for the prevention of depression induced by high-dose interferon alfa. N Engl J Med 2001; 344:961–966CrossrefGoogle Scholar

13 Capuron L, Gumnick JF, Musselman DL, Lawson DH, Reemsnyder A, Nemeroff CB, Miller AH: Neurobehavioral effects of interferon-alpha in cancer patients: phenomenology and paroxetine responsiveness of symptom dimensions. Neuropsychopharmacology 2002; 26:643–652CrossrefGoogle Scholar

14 Capuron L, Ravaud A: Prediction of the depressive effects of interferon alfa therapy by the patient’s initial affective state. N Engl J Med 1999; 340:1370CrossrefGoogle Scholar

15 Capuron L, Raison CL, Musselman DL, Lawson DH, Nemeroff CB, Miller AH: Association of exaggerated HPA axis response to the initial injection of interferon-alpha with development of depression during interferon-alpha therapy. Am J Psychiatry 2003; 160:1342–1345CrossrefGoogle Scholar

16 Constant A, Castera L, Dantzer R, Couzigou P, de Ledinghen V, Demotes-Mainard J, Henry C: Mood alterations during interferon-alfa therapy in patients with chronic hepatitis C: evidence for an overlap between manic/hypomanic and depressive symptoms. J Clin Psychiatry 2005; 66:1050–1057CrossrefGoogle Scholar

17 Soskin DP, Carl JR, Alpert J, Fava M: Antidepressant effects on emotional temperament: toward a biobehavioral research paradigm for major depressive disorder. CNS Neurosci Ther 2012; 18:441–451CrossrefGoogle Scholar

18 Zorrilla EP, Luborsky L, McKay JR, Rosenthal R, Houldin A, Tax A, McCorkle R, Seligman DA, Schmidt K: The relationship of depression and stressors to immunological assays: a meta-analytic review. Brain Behav Immun 2001; 15:199–226CrossrefGoogle Scholar

19 Mössner R, Mikova O, Koutsilieri E, Saoud M, Ehlis AC, Müller N, Fallgatter AJ, Riederer P: Consensus paper of the WFSBP Task Force on Biological Markers: biological markers in depression. World J Biol Psychiatry 2007; 8:141–174CrossrefGoogle Scholar

20 Simon NM, McNamara K, Chow CW, Maser RS, Papakostas GI, Pollack MH, Nierenberg AA, Fava M, Wong KK: A detailed examination of cytokine abnormalities in Major Depressive Disorder. Eur Neuropsychopharmacol 2008; 18:230–233CrossrefGoogle Scholar

21 Raison CL, Lowry CA, Rook GA: Inflammation, sanitation, and consternation: loss of contact with coevolved, tolerogenic microorganisms and the pathophysiology and treatment of major depression. Arch Gen Psychiatry 2010; 67:1211–1224CrossrefGoogle Scholar

22 Maes M: Depression is an inflammatory disease, but cell-mediated immune activation is the key component of depression. Prog Neuropsychopharmacol Biol Psychiatry 2011; 35:664–675CrossrefGoogle Scholar

23 Evans DL, Charney DS, Lewis L, Golden RN, Gorman JM, Krishnan KR, Nemeroff CB, Bremner JD, Carney RM, Coyne JC, Delong MR, Frasure-Smith N, Glassman AH, Gold PW, Grant I, Gwyther L, Ironson G, Johnson RL, Kanner AM, Katon WJ, Kaufmann PG, Keefe FJ, Ketter T, Laughren TP, Leserman J, Lyketsos CG, McDonald WM, McEwen BS, Miller AH, Musselman D, O’Connor C, Petitto JM, Pollock BG, Robinson RG, Roose SP, Rowland J, Sheline Y, Sheps DS, Simon G, Spiegel D, Stunkard A, Sunderland T, Tibbits P, Valvo WJ: Mood disorders in the medically ill: scientific review and recommendations. Biol Psychiatry 2005; 58:175–189CrossrefGoogle Scholar

24 Szczepanska-Sadowska E, Cudnoch-Jedrzejewska A, Ufnal M, Zera T: Brain and cardiovascular diseases: common neurogenic background of cardiovascular, metabolic and inflammatory diseases. J Physiol Pharmacol 2010; 61:509–521Google Scholar

25 Quan N, Banks WA: Brain-immune communication pathways. Brain Behav Immun 2007; 21:727–735CrossrefGoogle Scholar

26 Bluthé RM, Walter V, Parnet P, Layé S, Lestage J, Verrier D, Poole S, Stenning BE, Kelley KW, Dantzer R: Lipopolysaccharide induces sickness behaviour in rats by a vagal mediated mechanism. C R Acad Sci III 1994; 317:499–503Google Scholar

27 Pan W, Kastin AJ: TNFalpha transport across the blood-brain barrier is abolished in receptor knockout mice. Exp Neurol 2002; 174:193–200CrossrefGoogle Scholar

28 D’Mello C, Le T, Swain MG: Cerebral microglia recruit monocytes into the brain in response to tumor necrosis factoralpha signaling during peripheral organ inflammation. J Neurosci 2009; 29:2089–2102CrossrefGoogle Scholar

29 Komaki G, Arimura A, Koves K: Effect of intravenous injection of IL-1 beta on PGE2 levels in several brain areas as determined by microdialysis. Am J Physiol 1992; 262:E246–E251Google Scholar

30 Healy D: The antidepressant era. Cambridge, Mass., Harvard University Press, 1997Google Scholar

31 Maas JW: Biogenic amines and depression. Biochemical and pharmacological separation of two types of depression. Arch Gen Psychiatry 1975; 32:1357–1361CrossrefGoogle Scholar

32 Pitchot W, Hansenne M, Pinto E, Reggers J, Fuchs S, Ansseau M: 5-Hydroxytryptamine 1A receptors, major depression, and suicidal behavior. Biol Psychiatry 2005; 58:854–858CrossrefGoogle Scholar

33 Krishnan V, Nestler EJ: The molecular neurobiology of depression. Nature 2008; 455:894–902CrossrefGoogle Scholar

34 Chang T, Fava M: The future of psychopharmacology of depression. J Clin Psychiatry 2010; 71:971–975CrossrefGoogle Scholar

35 Lestage J, Verrier D, Palin K, Dantzer R: The enzyme indoleamine 2,3-dioxygenase is induced in the mouse brain in response to peripheral administration of lipopolysaccharide and superantigen. Brain Behav Immun 2002; 16:596–601CrossrefGoogle Scholar

36 Ruhé HG, Mason NS, Schene AH: Mood is indirectly related to serotonin, norepinephrine and dopamine levels in humans: a meta-analysis of monoamine depletion studies. Mol Psychiatry 2007; 12:331–359CrossrefGoogle Scholar

37 Bonaccorso S, Marino V, Puzella A, Pasquini M, Biondi M, Artini M, Almerighi C, Verkerk R, Meltzer H, Maes M: Increased depressive ratings in patients with hepatitis C receiving interferon-alpha-based immunotherapy are related to interferon-alpha-induced changes in the serotonergic system. J Clin Psychopharmacol 2002; 22:86–90CrossrefGoogle Scholar

38 Capuron L, Neurauter G, Musselman DL, Lawson DH, Nemeroff CB, Fuchs D, Miller AH: Interferon-alpha-induced changes in tryptophan metabolism. relationship to depression and paroxetine treatment. Biol Psychiatry 2003; 54:906–914CrossrefGoogle Scholar

39 Zhu CB, Blakely RD, Hewlett WA: The proinflammatory cytokines interleukin-1beta and tumor necrosis factor-alpha activate serotonin transporters. Neuropsychopharmacology 2006; 31:2121–2131CrossrefGoogle Scholar

40 Zhu CB, Lindler KM, Owens AW, Daws LC, Blakely RD, Hewlett WA: Interleukin-1 receptor activation by systemic lipopolysaccharide induces behavioral despair linked to MAPK regulation of CNS serotonin transporters. Neuropsychopharmacology 2010; 35:2510–2520CrossrefGoogle Scholar

41 Bull SJ, Huezo-Diaz P, Binder EB, Cubells JF, Ranjith G, Maddock C, Miyazaki C, Alexander N, Hotopf M, Cleare AJ, Norris S, Cassidy E, Aitchison KJ, Miller AH, Pariante CM: Functional polymorphisms in the interleukin-6 and serotonin transporter genes, and depression and fatigue induced by interferon-alpha and ribavirin treatment. Mol Psychiatry 2009; 14:1095–1104CrossrefGoogle Scholar

42 Lotrich FE, Ferrell RE, Rabinovitz M, Pollock BG: Risk for depression during interferon-alpha treatment is affected by the serotonin transporter polymorphism. Biol Psychiatry 2009; 65:344–348CrossrefGoogle Scholar

43 Nestler EJ, Carlezon WA: The mesolimbic dopamine reward circuit in depression. Biol Psychiatry 2006; 59:1151–1159CrossrefGoogle Scholar

44 Roy A, Agren H, Pickar D, Linnoila M, Doran AR, Cutler NR, Paul SM: Reduced CSF concentrations of homovanillic acid and homovanillic acid to 5-hydroxyindoleacetic acid ratios in depressed patients: relationship to suicidal behavior and dexamethasone nonsuppression. Am J Psychiatry 1986; 143:1539–1545CrossrefGoogle Scholar

45 Lambert G, Johansson M, Agren H, Friberg P: Reduced brain norepinephrine and dopamine release in treatment-refractory depressive illness: evidence in support of the catecholamine hypothesis of mood disorders. Arch Gen Psychiatry 2000; 57:787–793CrossrefGoogle Scholar

46 Agren H, Reibring L: PET studies of presynaptic monoamine metabolism in depressed patients and healthy volunteers. Pharmacopsychiatry 1994; 27:2–6CrossrefGoogle Scholar

47 Klimek V, Schenck JE, Han H, Stockmeier CA, Ordway GA: Dopaminergic abnormalities in amygdaloid nuclei in major depression: a postmortem study. Biol Psychiatry 2002; 52:740–748CrossrefGoogle Scholar

48 Di Mascio M, Di Giovanni G, Di Matteo V, Prisco S, Esposito E: Selective serotonin reuptake inhibitors reduce the spontaneous activity of dopaminergic neurons in the ventral tegmental area. Brain Res Bull 1998; 46:547–554CrossrefGoogle Scholar

49 Meyer JH, McNeely HE, Sagrati S, Boovariwala A, Martin K, Verhoeff NP, Wilson AA, Houle S: Elevated putamen D(2) receptor binding potential in major depression with motor retardation: an [11C]raclopride positron emission tomography study. Am J Psychiatry 2006; 163:1594–1602CrossrefGoogle Scholar

50 Parsey RV, Oquendo MA, Zea-Ponce Y, Rodenhiser J, Kegeles LS, Pratap M, Cooper TB, Van Heertum R, Mann JJ, Laruelle M: Dopamine D(2) receptor availability and amphetamine-induced dopamine release in unipolar depression. Biol Psychiatry 2001; 50:313–322CrossrefGoogle Scholar

51 Hirvonen J, Karlsson H, Kajander J, Markkula J, Rasi-Hakala H, Någren K, Salminen JK, Hietala J: Striatal dopamine D2 receptors in medication-naive patients with major depressive disorder as assessed with [11C]raclopride PET. Psychopharmacology (Berl) 2008; 197:581–590CrossrefGoogle Scholar

52 Kumai T, Tateishi T, Tanaka M, Watanabe M, Shimizu H, Kobayashi S: Effect of interferon-alpha on tyrosine hydroxylase and catecholamine levels in the brain of rats. Life Sci 2000; 67:663–669CrossrefGoogle Scholar

53 Shuto H, Kataoka Y, Horikawa T, Fujihara N, Oishi R: Repeated interferon-alpha administration inhibits dopaminergic neural activity in the mouse brain. Brain Res 1997; 747:348–351CrossrefGoogle Scholar

54 Kitagami T, Yamada K, Miura H, Hashimoto R, Nabeshima T, Ohta T: Mechanism of systemically injected interferon-alpha impeding monoamine biosynthesis in rats: role of nitric oxide as a signal crossing the blood-brain barrier. Brain Res 2003; 978:104–114CrossrefGoogle Scholar

55 Qin L, Wu X, Block ML, Liu Y, Breese GR, Hong JS, Knapp DJ, Crews FT: Systemic LPS causes chronic neuroinflammation and progressive neurodegeneration. Glia 2007; 55:453–462CrossrefGoogle Scholar

56 Felger JC, Alagbe O, Pace TW, Woolwine BJ, Hu F, Raison CL, Miller AH: Early activation of p38 mitogen activated protein kinase is associated with interferon-alpha-induced depression and fatigue. Brain Behav Immun 2011; 25:1094–1098CrossrefGoogle Scholar

57 Morón JA, Zakharova I, Ferrer JV, Merrill GA, Hope B, Lafer EM, Lin ZC, Wang JB, Javitch JA, Galli A, Shippenberg TS: Mitogen-activated protein kinase regulates dopamine transporter surface expression and dopamine transport capacity. J Neurosci 2003; 23:8480–8488Google Scholar

58 Trullas R, Skolnick P: Functional antagonists at the NMDA receptor complex exhibit antidepressant actions. Eur J Pharmacol 1990; 185:1–10CrossrefGoogle Scholar

59 Pittenger C, Duman RS: Stress, depression, and neuroplasticity: a convergence of mechanisms. Neuropsychopharmacology 2008; 33:88–109CrossrefGoogle Scholar

60 Hashimoto K, Sawa A, Iyo M: Increased levels of glutamate in brains from patients with mood disorders. Biol Psychiatry 2007; 62:1310–1316CrossrefGoogle Scholar

61 Pitt D, Nagelmeier IE, Wilson HC, Raine CS: Glutamate uptake by oligodendrocytes: Implications for excitotoxicity in multiple sclerosis. Neurology 2003; 61:1113–1120CrossrefGoogle Scholar

62 Volterra A, Meldolesi J: Astrocytes, from brain glue to communication elements: the revolution continues. Nat Rev Neurosci 2005; 6:626–640CrossrefGoogle Scholar

63 Müller N, Schwarz MJ: The immune-mediated alteration of serotonin and glutamate: towards an integrated view of depression. Mol Psychiatry 2007; 12:988–1000CrossrefGoogle Scholar

64 Khairova RA, Machado-Vieira R, Du J, Manji HK: A potential role for pro-inflammatory cytokines in regulating synaptic plasticity in major depressive disorder. Int J Neuropsychopharmacol 2009; 12:561–578CrossrefGoogle Scholar

65 Rajkowska G, Miguel-Hidalgo JJ: Gliogenesis and glial pathology in depression. CNS Neurol Disord Drug Targets 2007; 6:219–233CrossrefGoogle Scholar

66 Duman RS, Monteggia LM: A neurotrophic model for stress-related mood disorders. Biol Psychiatry 2006; 59:1116–1127CrossrefGoogle Scholar

67 Santarelli L, Saxe M, Gross C, Surget A, Battaglia F, Dulawa S, Weisstaub N, Lee J, Duman R, Arancio O, Belzung C, Hen R: Requirement of hippocampal neurogenesis for the behavioral effects of antidepressants. Science 2003; 301:805–809CrossrefGoogle Scholar

68 Shirayama Y, Chen AC, Nakagawa S, Russell DS, Duman RS: Brain-derived neurotrophic factor produces antidepressant effects in behavioral models of depression. J Neurosci 2002; 22:3251–3261CrossrefGoogle Scholar

69 Tfilin M, Sudai E, Merenlender A, Gispan I, Yadid G, Turgeman G: Mesenchymal stem cells increase hippocampal neurogenesis and counteract depressive-like behavior. Mol Psychiatry 2010; 15:1164–1175CrossrefGoogle Scholar

70 Monteggia LM, Barrot M, Powell CM, Berton O, Galanis V, Gemelli T, Meuth S, Nagy A, Greene RW, Nestler EJ: Essential role of brain-derived neurotrophic factor in adult hippocampal function. Proc Natl Acad Sci USA 2004; 101:10827–10832CrossrefGoogle Scholar

71 Capuron L, Miller AH: Immune system to brain signaling: neuropsychopharmacological implications. Pharmacol Ther 2011; 130:226–238CrossrefGoogle Scholar

72 Ben Menachem-Zidon O, Goshen I, Kreisel T, Ben Menahem Y, Reinhartz E, Ben Hur T, Yirmiya R: Intrahippocampal transplantation of transgenic neural precursor cells overexpressing interleukin-1 receptor antagonist blocks chronic isolation-induced impairment in memory and neurogenesis. Neuropsychopharmacology 2008; 33:2251–2262CrossrefGoogle Scholar

73 Koo JW, Duman RS: IL-1beta is an essential mediator of the antineurogenic and anhedonic effects of stress. Proc Natl Acad Sci USA 2008; 105:751–756CrossrefGoogle Scholar

74 Goshen I, Kreisel T, Ounallah-Saad H, Renbaum P, Zalzstein Y, Ben-Hur T, Levy-Lahad E, Yirmiya R: A dual role for interleukin-1 in hippocampal-dependent memory processes. Psychoneuroendocrinology 2007; 32:1106–1115CrossrefGoogle Scholar

75 Goshen I, Kreisel T, Ben-Menachem-Zidon O, Licht T, Weidenfeld J, Ben-Hur T, Yirmiya R: Brain interleukin-1 mediates chronic stress-induced depression in mice via adrenocortical activation and hippocampal neurogenesis suppression. Mol Psychiatry 2008; 13:717–728CrossrefGoogle Scholar

76 Hardingham GE, Fukunaga Y, Bading H: Extrasynaptic NMDARs oppose synaptic NMDARs by triggering CREB shut-off and cell death pathways. Nat Neurosci 2002; 5:405–414CrossrefGoogle Scholar

77 Pariante CM, Miller AH: Glucocorticoid receptors in major depression: relevance to pathophysiology and treatment. Biol Psychiatry 2001; 49:391–404CrossrefGoogle Scholar

78 Sorrells SF, Caso JR, Munhoz CD, Sapolsky RM: The stressed CNS: when glucocorticoids aggravate inflammation. Neuron 2009; 64:33–39CrossrefGoogle Scholar

79 Johnson JD, O’Connor KA, Deak T, Stark M, Watkins LR, Maier SF: Prior stressor exposure sensitizes LPS-induced cytokine production. Brain Behav Immun 2002; 16:461–476CrossrefGoogle Scholar

80 Caspi A, Sugden K, Moffitt TE, Taylor A, Craig IW, Harrington H, McClay J, Mill J, Martin J, Braithwaite A, Poulton R: Influence of life stress on depression: moderation by a polymorphism in the 5-HTT gene. Science 2003; 301:386–389CrossrefGoogle Scholar

81 Pace TW, Hu F, Miller AH: Cytokine-effects on glucocorticoid receptor function: relevance to glucocorticoid resistance and the pathophysiology and treatment of major depression. Brain Behav Immun 2007; 21:9–19CrossrefGoogle Scholar

82 Hu F, Pace TW, Miller AH: Interferon-alpha inhibits glucocorticoid receptor-mediated gene transcription via STAT5 activation in mouse HT22 cells. Brain Behav Immun 2009; 23:455–463CrossrefGoogle Scholar

83 Smoak KA, Cidlowski JA: Mechanisms of glucocorticoid receptor signaling during inflammation. Mech Ageing Dev 2004; 125:697–706CrossrefGoogle Scholar

84 Raison CL, Borisov AS, Woolwine BJ, Massung B, Vogt G, Miller AH: Interferon-alpha effects on diurnal hypothalamic-pituitary-adrenal axis activity: relationship with proinflammatory cytokines and behavior. Mol Psychiatry 2010; 15:535–547CrossrefGoogle Scholar

85 Maes M, Bosmans E, Meltzer HY, Scharpé S, Suy E: Interleukin-1 beta: a putative mediator of HPA axis hyperactivity in major depression? Am J Psychiatry 1993; 150:1189–1193CrossrefGoogle Scholar

86 Fitzgerald P, O’Brien SM, Scully P, Rijkers K, Scott LV, Dinan TG: Cutaneous glucocorticoid receptor sensitivity and pro-inflammatory cytokine levels in antidepressant-resistant depression. Psychol Med 2006; 36:37–43CrossrefGoogle Scholar

87 Pace TW, Miller AH: Cytokines and glucocorticoid receptor signaling. Relevance to major depression. Ann N Y Acad Sci 2009; 1179:86–105CrossrefGoogle Scholar

88 Harrison PJ: The neuropathology of primary mood disorder. Brain 2002; 125:1428–1449CrossrefGoogle Scholar

89 Ressler KJ, Mayberg HS: Targeting abnormal neural circuits in mood and anxiety disorders: from the laboratory to the clinic. Nat Neurosci 2007; 10:1116–1124CrossrefGoogle Scholar

90 Mayberg HS, Lozano AM, Voon V, McNeely HE, Seminowicz D, Hamani C, Schwalb JM, Kennedy SH: Deep brain stimulation for treatment-resistant depression. Neuron 2005; 45:651–660CrossrefGoogle Scholar

91 Hasler G: Pathophysiology of depression: do we have any solid evidence of interest to clinicians? World Psychiatry 2010; 9:155–161CrossrefGoogle Scholar

92 Juengling FD, Ebert D, Gut O, Engelbrecht MA, Rasenack J, Nitzsche EU, Bauer J, Lieb K: Prefrontal cortical hypometabolism during low-dose interferon alpha treatment. Psychopharmacology (Berl) 2000; 152:383–389CrossrefGoogle Scholar

93 Capuron L, Pagnoni G, Demetrashvili MF, Lawson DH, Fornwalt FB, Woolwine B, Berns GS, Nemeroff CB, Miller AH: Basal ganglia hypermetabolism and symptoms of fatigue during interferon-alpha therapy. Neuropsychopharmacology 2007; 32:2384–2392CrossrefGoogle Scholar

94 Bersano A, Aghemo A, Rumi MG, Ballabio E, Candelise L, Colombo M: Recovery after L-DOPA treatment in peginterferon and ribavirin induced parkinsonism. Eur J Intern Med 2008; 19:370–371CrossrefGoogle Scholar

95 Capuron L, Pagnoni G, Demetrashvili M, Woolwine BJ, Nemeroff CB, Berns GS, Miller AH: Anterior cingulate activation and error processing during interferon-alpha treatment. Biol Psychiatry 2005; 58:190–196CrossrefGoogle Scholar

96 Harrison NA, Brydon L, Walker C, Gray MA, Steptoe A, Critchley HD: Inflammation causes mood changes through alterations in subgenual cingulate activity and mesolimbic connectivity. Biol Psychiatry 2009; 66:407–414CrossrefGoogle Scholar

97 Harrison NA, Brydon L, Walker C, Gray MA, Steptoe A, Dolan RJ, Critchley HD: Neural origins of human sickness in interoceptive responses to inflammation. Biol Psychiatry 2009; 66:415–422CrossrefGoogle Scholar

98 Eisenberger NI, Lieberman MD: Why rejection hurts: a common neural alarm system for physical and social pain. Trends Cogn Sci 2004; 8:294–300CrossrefGoogle Scholar

99 Eisenberger NI, Lieberman MD, Satpute AB: Personality from a controlled processing perspective: an fMRI study of neuroticism, extraversion, and self-consciousness. Cogn Affect Behav Neurosci 2005; 5:169–181CrossrefGoogle Scholar

100 Kendler KS, Gatz M, Gardner CO, Pedersen NL: Personality and major depression: a Swedish longitudinal, population-based twin study. Arch Gen Psychiatry 2006; 63:1113–1120CrossrefGoogle Scholar

101 Brustolim D, Ribeiro-dos-Santos R, Kast RE, Altschuler EL, Soares MB: A new chapter opens in anti-inflammatory treatments: the antidepressant bupropion lowers production of tumor necrosis factor-alpha and interferon-gamma in mice. Int Immunopharmacol 2006; 6:903–907CrossrefGoogle Scholar

102 Savikin K, Dobrić S, Tadić V, Zdunić G: Antiinflammatory activity of ethanol extracts of Hypericum perforatum L., H. barbatum Jacq., H. hirsutum L., H. richeri Vill. and H. androsaemum L. in rats. Phytother Res 2007; 21:176–180CrossrefGoogle Scholar

103 Yu J, Sauter S, Parlesak A: Suppression of TNF-alpha production by S-adenosylmethionine in human mononuclear leukocytes is not mediated by polyamines. Biol Chem 2006; 387:1619–1627Google Scholar

104 Tyring S, Gottlieb A, Papp K, Gordon K, Leonardi C, Wang A, Lalla D, Woolley M, Jahreis A, Zitnik R, Cella D, Krishnan R: Etanercept and clinical outcomes, fatigue, and depression in psoriasis: double-blind placebo-controlled randomised phase III trial. Lancet 2006; 367:29–35CrossrefGoogle Scholar

105 Müller N, Schwarz MJ, Dehning S, Douhe A, Cerovecki A, Goldstein-Müller B, Spellmann I, Hetzel G, Maino K, Kleindienst N, Möller HJ, Arolt V, Riedel M: The cyclooxygenase-2 inhibitor celecoxib has therapeutic effects in major depression: results of a double-blind, randomized, placebo controlled, add-on pilot study to reboxetine. Mol Psychiatry 2006; 11:680–684CrossrefGoogle Scholar

106 Raison CL, Rutherford RE, Woolwine BJ, Shuo C, Schettler P, Drake DF, Haroon E, Miller AH: A randomized controlled trial of the tumor necrosis factor antagonist Infliximab for treatment-resistant depression: the role of baseline inflammatory biomarkers. Arch Gen Psychiatry 2012; 10:1–11Google Scholar

107 Howren MB, Lamkin DM, Suls J: Associations of depression with C-reactive protein, IL-1, and IL-6: a meta-analysis. Psychosom Med 2009; 71:171–186CrossrefGoogle Scholar

108 Capuron L, Su S, Miller AH, Bremner JD, Goldberg J, Vogt GJ, Maisano C, Jones L, Murrah NV, Vaccarino V: Depressive symptoms and metabolic syndrome: is inflammation the underlying link? Biol Psychiatry 2008; 64:896–900CrossrefGoogle Scholar

109 Pace TW, Mletzko TC, Alagbe O, Musselman DL, Nemeroff CB, Miller AH, Heim CM: Increased stress-induced inflammatory responses in male patients with major depression and increased early life stress. Am J Psychiatry 2006; 163:1630–1633LinkGoogle Scholar

110 Rooks C, Veledar E, Goldberg J, Bremner JD, Vaccarino V: Early trauma and inflammation: role of familial factors in a study of twins. Psychosom Med 2012; 74:146–152CrossrefGoogle Scholar

111 Motivala SJ: Sleep and inflammation: psychoneuroimmunology in the context of cardiovascular disease. Ann Behav Med 2011; 42:141–152CrossrefGoogle Scholar

112 Häfner S, Emeny RT, Lacruz ME, Baumert J, Herder C, Koenig W, Thorand B, Ladwig KHKORA Study Investigators: Association between social isolation and inflammatory markers in depressed and non-depressed individuals: results from the MONICA/KORA study. Brain Behav Immun 2011; 25:1701–1707CrossrefGoogle Scholar

113 Friedman EM: Sleep quality, social well-being, gender, and inflammation: an integrative analysis in a national sample. Ann N Y Acad Sci 2011; 1231:23–34CrossrefGoogle Scholar