You are on page 1of 29

HHS Public Access

Author manuscript
J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Author Manuscript

Published in final edited form as:


J Affect Disord Rep. 2021 December ; 6: . doi:10.1016/j.jadr.2021.100177.

A chicken and egg scenario in psychoneuroimmunology:


Bidirectional mechanisms linking cytokines and depression
Manivel Rengasamya,*, Anna Marslandb, Meredith Spadaa, Kimberly Hsiungc, Tessa
Kovatsa, Rebecca B. Pricea,b
aDepartment of Psychiatry, University of Pittsburgh, Pittsburgh, PA, USA
bDepartment of Psychology, University of Pittsburgh, Pittsburgh, PA, USA
Author Manuscript

cUniversity of Pittsburgh School of Medicine, University of Pittsburgh, Pittsburgh, PA, USA

Abstract
Background: Cytokines are an important part of the immune system. Certain cytokines, such
as interleukin-6 (IL-6) and tumor necrosis factor alpha (TNFα), have well-described associations
with depression. Various mechanisms exist that may explain bidirectional effects of cytokines
on depression and vice versa. No recent reviews to our knowledge have comprehensively
characterized both these mechanisms and the interaction of these mechanisms using evidence
from the molecular level to the clinical level. The goal of this review is to both evaluate the present
knowledge base and identify knowledge gaps to help guide future research.

Methods: We conducted an extensive bibliographic search across multiple databases, using both
Author Manuscript

general (e.g. “cytokine”) and topic-specific (e.g. “kynurenine”) keywords.

Results: We describe the most recent evidence outlining these mechanisms, including the role of
the hypothalamic pituitary axis, the kynurenine pathway, and neural circuitry. For relevant topics,
we outline the pathways by which cytokine activation may lead to depressive symptoms, and
how depressive symptomology may lead to elevations in cytokines. We also identify key areas
for future research, including the need for longitudinal clinical studies to examine causality in
pertinent mechanisms and modulating factors in the cytokine-depression interaction.

Limitations: Given the numerous potential mechanisms associating cytokines and depressions,
this review paper solely focuses on the most commonly described mechanisms at a basic level.

Conclusions: Bidirectional evidence exists for several mechanisms in the relationship between
Author Manuscript

cytokines and depression. However, more work is required to further elucidate the role of these
mechanisms in specific clinical populations.

This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/)


*
Corresponding author at: Western Psychiatric Institute and Clinic, 3811 O’Hara St., Pittsburgh, PA 15213, USA.
rengasamym@upmc.edu (M. Rengasamy).
Declaration of Competing Interest
There are no conflicts of interest associated with any author.
Supplementary materials
Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.jadr.2021.100177.
Rengasamy et al. Page 2

Keywords
Author Manuscript

Cytokines; Depression; Major depressive disorder; Inflammation; Interleukin-6;


Psychoneuroimmunology

1. Introduction
Major Depressive Disorder (MDD) ranks among one of the most disabling illnesses
worldwide, with significant personal and societal costs (Richards, 2011). More than 50%
of suicides occur in individuals with MDD, and thus MDD is specifically associated with
significant mortality (Henriksson et al., 1993). Although effective psychotropic medication
and psychotherapy treatments exist for individuals with MDD, not all depressed patients
are able to achieve remission of their symptoms even with treatment, leading to continued
Author Manuscript

risk of depression-associated morbidity and mortality (Fekadu et al., 2009). The lack of
objective biomarkers and symptom heterogeneity make treatment of MDD difficult. Thus,
understanding of the varied neurobiological mechanisms of depression may be important in
developing efficacious and targeted treatments for depression.

The “inflammatory” hypothesis of depression has received increasing attention in recent


years and postulates that immune system activation and altered cytokine regulation are
associated with depressive disorders. Cytokines are small proteins released by immune
system cells in reaction to certain stimuli (e.g. bacteria or proteins associated with cell
death). From peripheral circulation, these cytokines can cross the blood brain barrier or
have effects on the blood brain barrier to affect brain neuron function, as described in prior
reviews (Supplemental Table 1) (Dantzer, 2001). The inflammatory hypothesis suggests that,
at least in a subset of depressed patients, inflammatory mechanisms related to cytokine
Author Manuscript

activation (either in the periphery or the central nervous system (CNS)) are associated with
depression (Dantzer et al., 2008). Furthermore, the relationship between cytokine activation
and depression is likely bidirectional, with depression contributing to cytokine elevation
and vice versa (Amodeo et al., 2017; Young et al., 2014). Of note, the characterization of
depression as being a distinct concept from cytokine activation is part of a widely-used
conceptual framework within the fields of psychiatry and neuroscience, with the aim of
better understanding the pathophysiology of depression as well as the relationship between
the patient’s experience of depression symptoms and future inflammation. This perspective
is supported by studies that suggest a temporal distinction between the two processes, such
as those identifying that depressive symptoms at one point in time predict future cytokine
activation. However, an important alternative perspective is that cytokine activation and
depression may be part of a unitary syndrome that bridges both biological and psychological
Author Manuscript

aspects simultaneously (Glannon, 2020; Miresco and Kirmayer, 2006).

The mechanisms for the bidirectional cytokine-depression relationship have previously been
partially reviewed by other authors, although often with limited specificity or attention to
the theoretical, molecular, preclinical, or clinical evidence for such mechanisms (Felger
and Lotrich, 2013; Jeon and Kim, 2016). Major mechanisms involved in this relationship
include neurotransmitters, neurocircuitry, the hypothalamic pituitary adrenal axis (HPA),

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 3

the sympathetic nervous system (SNS), and the kynurenine pathway, among others. In this
Author Manuscript

review examining the latest evidence related to depression and cytokines (in both the blood
and brain), we hope to expound upon mechanisms of effects of cytokines on depression and
of depression on cytokines, ultimately to identify needed areas of future research.

2. Methods
We completed a structured narrative review by searching Google Scholar and PubMed using
a combination of keywords, including general keywords “depression,” “cytokines”, “major
depressive disorder,” “mechanism,” and “MDD”, and also topic-specific keywords (e.g.
“kynurenine” for the kynurenine pathway) when relevant. We examined at minimum the
top fifty most relevant peer-reviewed manuscripts (based on the search engine algorithm)
for each search or keyword. Topic and topic specific keywords were determined by
qualitative examination of review papers identified by general keyword search. We only
Author Manuscript

included papers that contributed novel information related to mechanisms not described in
prior manuscripts that we had included, which was determined by review of the potential
manuscript by the primary author MR. As noted by our keyword selection, we examined
both MDD and depressive symptomatology. We choose to focus on non-genetic mechanistic
factors given the inability to cover both genetic and other mechanistic factors in only one
review. Lastly, we would emphasize to readers that concepts have necessarily been reduced
to general key points given the intrinsic complexity of the referenced subjects and the wide
breadth of information.

3. Evidence for bidirectional causal relationship between cytokine


activation and depression
Author Manuscript

As described earlier, evidence exists explaining contribution of depression to cytokine


elevation, contribution of cytokine elevation to depression, and bidirectional interaction
of cytokine activation and depression, with respect to cytokines both in the periphery and
the CNS. Of note, peripheral cytokine elevations in depression are subclinical, typically
presenting without constituent symptoms of high-grade inflammation (e.g. fever) and lower
than peripheral cytokine elevations in immune or pathogenic illnesses.

In terms of cross-sectional evidence suggesting depression is associated with cytokine


elevation, several meta-analyses confirm elevations of cytokines in peripheral circulation
in patients with MDD compared to healthy controls, including elevated plasma interleukin-6
(IL-6), C-reactive protein (CRP)/high-sensitivity CRP (which detects lower levels of CRP),
and tumor-necrosis-factor-alpha (TNFα) (Dowlati et al., 2010). Additionally, less robust
Author Manuscript

evidence demonstrates elevations in plasma soluble IL-2R, IL-1β, IL-10, and IL-1 receptor
antagonist (IL-1RA) (Howren et al., 2009; Köhler et al., 2018; Yuan et al., 2019).
Concordantly, clinical cerebrospinal fluid (CSF) and brain autopsy studies in depressed and
suicidal individuals identify elevations in IL-1β, IL-6, and TNFα as being associated with
depressive symptomatology, although some mixed findings and small samples characterize
these studies (Carpenter et al., 2004; Dean et al., 2010; Felger et al., 2018; Levine et al.,
1999; Pandey et al., 2012; Sasayama et al., 2013; Tonelli et al., 2008). Most patients with
MDD do not show frank signs of clinical infection, such as fever or elevations in white

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 4

blood cell counts, suggesting lack of an exogenous source of cytokine activation. Taken
Author Manuscript

together, these findings suggest that depressive symptomatology contributes to immune


system activation.

The research evidence also shows that inflammation and peripheral cytokine activation
are associated with and contribute to depressive symptomatology, based on cross-sectional
studies and studies of individuals with comorbid medical illnesses. Models in rats and mice
show that depressive symptoms develop following the administration of lipopolysaccharide
or other inflammatory stimuli (Shen et al., 1999). In humans, cytokine treatments (e.g.
interferon alpha (IFNα) or IL-2) for certain illnesses, bacterial infection, viral infection,
and vaccines have all been associated with cytokine-mediated induction of depressive
symptoms (Capuron et al., 2001). Similarly, high comorbidity of autoimmune diseases
and depression have suggested a role of peripheral and central cytokines in depression,
with general identification of IL-1β, IL-6, and TNFα as key participants (Dantzer et al.,
Author Manuscript

2008; Kubera et al., 2011; Raison et al., 2006). Meanwhile, in individuals with comorbid
medical illness, anti-inflammatory treatments such as nonsteroidal anti-inflammatory drugs
(NSAIDs), statins, and cytokine inhibitors are associated with reduction in depressive
symptoms, often independently of improvement in comorbid medical illness (Kappelmann
et al., 2018; Köhler et al., 2014; Salagre et al., 2016). These findings suggest that immune
system activation may lead to depressive symptomology.

To further bolster the findings of cross-sectional studies, longitudinal clinical studies


identify a bidirectional relationship of depression and immune system alterations (primarily
peripheral) in depressed individuals, generally within timeframes between a few months to
less than five years. One meta-analysis and a few large cohort studies suggest that increases
in peripheral inflammatory indices (primarily CRP, IL-6, and IL-8) precede depressive
Author Manuscript

symptomatology, while one smaller study implicated CSF IL-1β in the development of
depression (Baune et al., 2012b; Gimeno et al., 2009; Khandaker et al., 2014; Lamers
et al., 2019; Milaneschi et al., 2009; Miller et al., 2019; Smith et al., 2018; Valkanova
et al., 2013). For instance, findings from the Dunedin cohort demonstrated that elevated
serum IL-6 in childhood was associated with depression in early adulthood (Khandaker et
al., 2014). Suggesting that depressive symptoms contribute to pro-inflammatory cytokine
activation, other studies find that general reduction of depressive symptoms leads to a
reduction in peripheral inflammatory markers, or that depressive symptoms predict future
peripheral IL-6, CRP or other cytokine elevation (Supplemental Table 2) (Dahl et al.,
2016; Glaus et al., 2018; Stewart et al., 2009). Moreover, consistent with these findings
which demonstrate a bidirectional relationship between inflammation and depression, a
recent meta-analysis identified that elevated CRP and IL-6 (measured via saliva, blood
Author Manuscript

or lumbar puncture) predicted higher future depressive symptoms, and vice versa (Mac
Giollabhui et al., 2020). In terms of individual studies examining bidirectional relationships,
a study of depression following IFNα treatment used time-lagged analyses to show that
elevated peripheral IL-6 predicted increased depressive symptomatology and vice versa
(Prather et al., 2009). Similarly, a recent smaller study examining male twins found
that elevated plasma IL-6 predicted future depressive symptomatology, while elevated
depressive symptoms prospectively predicted elevated plasma CRP levels (Huang et al.,

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 5

2018). Together, the limited research suggests that depressive symptomatology leads to
Author Manuscript

elevations in cytokines, and vice versa, at least in the short term.

4. Mechanisms for cytokine mediated depression: cytokine activation


leading to depression
4.1. Overview
The most prominent mechanisms for cytokine-induced depression involve neurotransmitter
alterations, disruption of neurogenesis, hypothalamic pituitary axis activation, kynurenine
pathway activation, and the gut-brain axis (Felger and Lotrich, 2013; Jeon and Kim, 2016).
We will examine the evidence for each of these mechanisms, at the theoretical, molecular,
preclinical, and clinical level (Felger and Lotrich, 2013).
Author Manuscript

4.2. Neurotransmitter alterations


CNS neurotransmitter alterations are well studied in MDD, with ongoing debate focused
on the relative primacy and contributions of various neurotransmitters alterations in
depression, including low synaptic monoamines (Capuron and Miller, 2011). Lower synaptic
levels of serotonin, dopamine, norepinephrine, and gamma-aminobutyric acid (GABA)
and higher levels of glutamate are likely present in certain regions of the CNS of
depressed patients. Effects of cytokines (prominently IL-1β, TNFα, IFNα, or IL-6) on
neurotransmitter alterations have primarily been demonstrated in in-vitro models, murine
models, and models with exogenous immune stimulus administration (Felger and Lotrich,
2013; Miller et al., 2013). In these studies, proinflammatory cytokines have been shown to
reduce monoamine synaptic activity, increase monoamine reuptake, or decrease monoamine
release (Supplemental Figure 1) (Hodes et al., 2015; Miller et al., 2013). Indirectly, they
Author Manuscript

may affect monoamine activity via several mechanisms including monoamine pathway
cofactors (e.g. tetrahydrobiopterin, BH4), direct enzymatic effects, modulation of the
vagal nerve, or oxidative stress (Miller et al., 2013; Miller et al., 2009). Specifically,
IL-1β, IFNα, and TNFα can increase activity of monoamine enzymes, such as the
serotonin reuptake transporter enzyme (SERT) (Zhu et al., 2006). Similarly, IFNα has
been implicated in extracellular dopamine regulation (Felger and Lotrich, 2013). These
cytokine-induced monoamine alterations may occur in depression-relevant brain regions
such as the hippocampus, hypothalamus, and prefrontal cortex (PFC) (Zalcman et al., 1994).
For instance, one clinical study of patients treated with IFNα found a negative correlation
between CSF 5-Hydroxyindoleacetic acid (5-HIAA) and CSF IL-6 (Raison et al., 2009).
However, a recent post-hoc study of treatment-resistant MDD patients found no association
between plasma IL-6 and CSF monoamine metabolites or BH4, suggesting further research
Author Manuscript

is warranted (Rengasamy et al., 2018).

Apart from the monoamine pathway, cytokines (primarily TNFα, IL-1, IL-6, and IL-18)
have been associated with changes in glutamate signaling (Supplemental Figure 2), via
enzymatic effects and the kynurenine pathway (described later) (Haroon and Miller, 2016).
These changes include effects on glutamate release, glutamate reuptake, and enhancement
of signal transducers of glutamate receptors (Haroon and Miller, 2016). For example, TNFα
may increase glutamate release by microglia through upregulation of glutaminase (which

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 6

converts glutamine to glutamate) and to block astroglial reuptake of glutamate, leading


Author Manuscript

to glutamate-mediated neurotoxicity (Takeuchi et al., 2006). Similarly, in neurons, IL-1β


was shown to activate tyrosine kinase to enhance NMDA-glutamate-mediated neurotoxicity
(Viviani et al., 2003). Concurrently, some cytokines (e.g. TNFα) may reduce the anti-
excitatory neurotransmitter GABA’s activity via GABA receptor endocytosis, thereby
upgregulating glutamate activity (Stellwagen et al., 2005). However, this relationship with
glutamate is not consistent, as IL-1β may potentially enhance GABA receptor function
via ion channel activation (Miller et al., 1991; Vezzani and Viviani, 2015). IL-6 has been
associated with decreases in both glutamate and GABA signaling via both direct receptor
effects and reduction of GABA cell populations (Dugan et al., 2009; Vezzani and Viviani,
2015). Cytokines (primarily IL-1β) also limit effects of cannabinoid receptors, which are
involved in blockade of cytokine effects on other neurotransmitter receptors and reduction of
excitotoxicity (Vezzani and Viviani, 2015). Clearly, cytokine effects on neurotransmitter
Author Manuscript

activity exist but the mechanisms by which they occur are complex, requiring further
specificity to identify state and region-specific effects.

4.3. Dysfunctional neurogenesis


Dysfunctional neurogenesis at the molecular level, particularly in the hippocampus, has
been associated with depressive-like behaviors in animal models with potential cytokine
involvement (Krishnan and Nestler, 2008). Notably, proinflammatory cytokine (e.g. IL-6)
activation has been associated with both decreases in neurogenesis and increases in
neurogenesis, with effects partially dependent on the temporality of activation and level
of cytokine elevation (Erta et al., 2012). Two factors implicated in cytokine-related
dysfunctional neurogenesis are apoptosis and neurotrophic factors deficits (Supplemental
Figure 3). IL-1β and TNFα have been associated with decreased neurogenesis via activation
Author Manuscript

of apoptotic pathways or reduction of cell differentiation into neurons (Kohman and Rhodes,
2013), generally in murine or progenitor cell models (Borsini et al., 2015; Downen et
al., 1999). Regarding neurotrophic factors, brain-derived neurotrophic factor (BDNF) is a
neurotrophic factor present in the hippocampus, and deficits of BDNF have been associated
with depressive symptomatology (Molendijk et al., 2014). Some cytokines, such as IL-1β,
have been associated with reduction in hippocampal BDNF mRNA expression in a murine
studies (Numakawa et al., 2014). Early clinical studies do not bear out such associations,
finding positive associations of plasma IL-6 with BDNF and null/positive relationship of
TNFα and BDNF (Neupane et al., 2015; Patas et al., 2014), suggesting the need for
future clinical studies to help understand the reasons for differences between clinical and
preclinical studies.

4.4. Neuroanatomical and neurocircuitry alterations


Author Manuscript

In conjunction with effects on neurotransmitters and neurogenesis, peripheral cytokine


activation has been associated with abnormalities in certain brain regions, neural circuits,
and neural networks. Meta-analyses find that patients with MDD have reductions in
volumes of their hippocampus, anterior cingulate cortex (ACC), and orbitofrontal cortex,
with less replicated findings for volumetric reduction of the basal ganglia and amygdala
(Hamilton et al., 2008; Lorenzetti et al., 2009; Schmaal et al., 2016; Wise et al.,
2017). Consistently, studies in depressed individuals suggest associations of increased

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 7

pro-inflammatory cytokines (e.g. IL-6, TNFα) with smaller hippocampal volume (Baune
Author Manuscript

et al., 2012a; Frodl and Amico, 2014). Additionally, peripheral cytokines such as IL-6
are implicated in development of white matter hyperintensities (WMH), which indicate
demyelination or white matter loss and are associated with depressive symptomology
(Herrmann et al., 2008; Satizabal et al., 2012). Ischemic lesions in the white matter (along
with WMH), potentially due to downstream effects of cytokine activation through effects on
vascular homeostasis, are also theorized to contribute to depressive symptoms as described
in the vascular depression hypothesis (Taylor et al., 2013). Lastly, pre-clinical models
suggest that cytokine effects may lead to structural breakdown of the blood brain barrier
(BBB) (Banks, 2005). This BBB breakdown may lead to increased cytokine presence in
the brain, particularly in regions controlling mood and emotion. However, no robust clinical
evidence exists demonstrating evidence of BBB breakdown in conditions of low-grade
inflammation.
Author Manuscript

In regards to brain activation patterns, a recent fMRI meta-analysis found activations of


limbic (amygdala, hippocampus, hypothalamus), striatal, brainstem, and cortical (e.g. dorsal
anterior cingulate cortex (dACC), dorsomedial prefrontal cortex (dmPFC)) regions that were
associated with peripheral inflammation, including cytokines such as IL-1β, IL-6, IL-1RA,
or soluble TNFαR2 (Kraynak et al., 2018). Cytokines such as IL-1 may cross the blood-
brain barrier to activate neurons in brain regions central to depression, such as the amygdala
(Konsman et al., 2002). Given the concordance in inflammatory and depression-associated
brain regions, inflammation may lead to these changes in functional brain activation, as
seen in studies implicating the dACC, ventral striatum, and subgenual ACC (sgACC) in
mediating IL-6 or TNFαR2 associated mood changes (Frodl and Amico, 2014).

In depression, neurocircuitry changes are also characterized by different connections


Author Manuscript

between areas of the brain as compared to healthy controls (Kaiser et al., 2015). These
neurocircuitry changes are associated with difficulty with emotional regulation and other
symptoms characterizing and preceding depression. For instance, individuals with MDD
may have abnormalities in the connections of the limbic-medial prefrontal circuit, leading to
a decrease in medial prefrontal cortical activity and excess amygdala activation (Price and
Drevets, 2010). Increased plasma CRP, a marker of cytokine activation, similarly has been
associated with decreased amygdala-ventromedial PFC connectivity (Mehta et al., 2018).
Two neurocircuits associated with cytokine activation, largely on a theoretical basis, are the
default mode network (DMN) and reward circuit, described below. However, other studies in
non-depressed populations suggest peripheral cytokine associations (IL-6, IFNα) with motor
neurocircuitry (Miller et al., 2013).
Author Manuscript

Reward circuits involve projections of dopamine neurons from the ventral tegmental area to
the ventral striatum and nucleus accumbens. Pro-inflammatory cytokines (e.g. IL-6, IFNα,
TNFα) may act through mechanisms including kynurenine pathway activation, reduction in
dopamine neurotransmission, or direct neurotoxicity as in the case of IL-1 RA (Swardfager
et al., 2016) to disrupt reward circuitry. For instance, one study in depressed patients found
that increased peripheral CRP, IL-6, IL-1β, and IL-1RA were associated with decreased
striatal-ventromedial PFC connectivity, a component of the reward pathway (Felger et
al., 2016). Animal and human studies suggest these alterations of reward circuitry that

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 8

follow activation of inflammatory processes results in symptoms of anhedonia. In another


Author Manuscript

study, increased plasma CRP (in conjunction with basal ganglia glutamate) was associated
with increased anhedonia in depressed patients, with association of anhedonia and reward
processing regions (Haroon et al., 2018). Ultimately, cytokine-induced anhedonia may
progressively lead to other symptoms of depression.

The DMN is theoretically associated with rumination or attention deficits in depression,


and has primary nodes in the posterior cingulate cortex and ventromedial PFC (Hamilton
et al., 2015). One study in healthy adults found that plasma IL-6 was associated positively
with sgACC connectivity in the DMN and negatively with dmPFC connectivity in the DMN
(Marsland et al., 2017a). Similarly, lipopolysaccharide-induced elevations of plasma IL-6 in
healthy humans has been associated with increased connectivity between the left thalamus
and right posterior cingulate cortex (Labrenz et al., 2016). Thus, cytokine effects may induce
changes in the DMN that relate to depressive symptoms.
Author Manuscript

Although the extant research has identified associations with both the DMN and reward
pathways with cytokines in depression, forthcoming studies are needed help identify if such
findings are specific to these pathways and the mechanism of action of such pathways.

4.5. Hypothalamus pituitary adrenal (HPA) axis activation


Overactivation of the HPA axis has been demonstrated in depressed patients for many
years, consistent with significantly elevated rates of depression in hypercortisolemic
states such as Cushing’s disease (Pariante and Lightman, 2008; Sonino and Fava, 1998).
The HPA axis involves the hypothalamus, which produces cortisol releasing hormone
(CRH), which then interacts with pituitary cells to release adrenocorticotrophic hormone
(ACTH), which stimulates cortisol production in the adrenal glands (Pariante and Lightman,
Author Manuscript

2008). In depression, numerous studies demonstrate abnormal dexamethasone cortisol


response (e.g. decreased suppression of cortisol after administration of dexamethasone)
or hypercortisolemia (Stetler and Miller, 2011). However, these findings are not consistent
in all depressed patients and may vary depending on factors such as cortisol source, age,
gender, type of depression, severity of depression, or history of trauma (Burke et al., 2005;
Knorr et al., 2010; Kunugi et al., 2015; Pariante and Lightman, 2008; Zorn et al., 2017).

As demonstrated primarily in murine and in vitro models, cytokines such as IL-6 and IL-1β
centrally activate the hypothalamus, likely at parvocellular paraventricular cortisol-releasing
hormone (PVN-CRH) neurons (Bethin et al., 2000; Ferri and Ferguson, 2003). Cytokine
effects at PVN-CRH neurons may be mediated via a transcriptional effects at the CRH
gene, intermediary reactive oxygen species, or activation of pathways involved in neuronal
Author Manuscript

depolarization (Schrader et al., 2006; Shi et al., 2010; Vallières and Rivest, 1999). Via PVN-
CRH neuron release of CRH and CRH-mediated ACTH release at the median eminence of
the pituitary, cortisol is ultimately released from the adrenal cortex. Concurrently, cytokines
such as IL-1β and IL-6 activate the transcription factor nuclear factor kappa beta (NF-kβ) in
a variety of cell types including macrophages (Miller et al., 2009). Such activation leads to
reduction of glucocorticoid receptor mRNA production, increase in the inert glucocorticoid-
binding glucocorticoid receptor beta (GR-β), and disruption of activated glucocorticoid
receptor function via translocation blockade and direct protein-protein interactions (Pace and

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 9

Miller, 2009; Pariante et al., 1999). These changes to glucocorticoid effects on certain cell
Author Manuscript

types may limit negative feedback of cortisol on the HPA axis. Given that the hippocampus
provides negative feedback to the HPA axis, cytokine-mediated hippocampal damage may
result in further activation of the HPA axis (Schloesser et al., 2009). Overall, the theoretical
and preclinical evidence indicates that cytokine activation leads to HPA axis activation
(Supplemental Figure 4) (Kim et al., 2016).

In depressed patients, various cortisol measures (e.g. plasma cortisol) demonstrate either
no association between plasma cytokines (e.g. IL-1 and IL-2) or positive association
with cytokines (e.g. IL-6) (Anisman et al., 1999; Kaestner et al., 2005; Karlovic et al.,
2012; Wichers et al., 2007). Such varied results may be due to multiple issues in the
reliable measurement of cortisol levels, lack of robust methodology, depression subtype
heterogeneity, or the cross-sectional nature of such analyses. Thus, the existing clinical
research does not demonstrate a clear association between cytokines and HPA axis activation
Author Manuscript

in depression, despite evidence of acute cytokine-induced HPA axis activation and increased
cytokine activity secondary to cortisol resistance. Upcoming research should clarify these
results in clinical samples using more robust methodology, particularly focusing on the
temporality of cytokine and HPA axis changes in relation to depressive symptoms.

4.6. Kynurenine pathway activation


More recently implicated in depressive pathology, the kynurenine pathway involves the
degradation of tryptophan into kynurenine (Supplemental Figure 5). Kynurenine accounts
for greater than 90% of tryptophan breakdown products, with serotonin being the other
tryptophan metabolite (Badawy, 2017). Tryptophan and kynurenine are transported across
the blood-brain barrier, with about 40% of CNS kynurenine produced from CNS tryptophan
(Badawy, 2017; Cuartero et al., 2016; Fukui et al., 1991). Kynurenine then is converted
Author Manuscript

into kynurenic acid and quinolinic acid. This conversion primarily occurs in the CNS by
indoleamine 2,3-dioxygenase (IDO) and peripherally by tryptophan 2,3-dioxygenase (TDO),
with greater activation of IDO under times of significant inflammatory activation. IDO is
induced by cytokines (most prominently IFNγ but also IFNα, IL-1β, IL-6, and TNFα) and
inhibited by nitric oxide, anti-inflammatory cytokines (e.g. IL-10, IL-4), and tryptophan.
In states without significant inflammatory activation, TDO is not induced by cytokines
but rather is induced by cortisol, heme cofactors, and nicotinamide adenine dinucleotide
phosphate (NADP) (Badawy, 2017).

Theoretically, kynurenine (KYN) may be converted preferentially into the neurotoxic


NMDA-glutamate agonist quinolinic acid (QA) as opposed to the neuroprotective NMDA
glutamate antagonist kynurenic acid (KA). Excess QA activation of NMDA receptors
Author Manuscript

may lead to cell death in depression-relevant brain regions via calcium mediated toxicity,
failure of synaptogenesis via mechanistic target of rapamycin (mTOR) or BDNF pathways
(Supplemental Figure 6) (Zanos and Gould, 2018). KA also acts as an alpha 7 nicotinic
acetylcholine receptor (α7nAChR) antagonist (Badawy, 2017). Antagonism of α7nAChR
has shown direct antidepressant properties in murine models, while pharmacological
blockade of nicotinic acetylcholine receptors has decreased depressive symptoms in humans
in small preliminary clinical studies, suggesting another neuroprotective mechanism for KA

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 10

(Philip et al., 2010). Other kynurenine pathway metabolites such as 3-hydroxykynurenine


Author Manuscript

and picolinic acid may affect ATP synthesis or induce oxidative stress and consequently
reduce neuronal survival in depression-relevant brain regions (Réus et al., 2015). Cytokines
(e.g. IL-6 or TNFα) mediate IDO activation in a number of suggested brain regions,
including the hippocampus (Réus et al., 2015). Preliminary findings in humans suggest that
kynurenine pathway activation may reduce volumes of the hippocampus or prefrontal cortex
(Meier et al., 2015a; Meier et al., 2015b). Lastly, a few studies demonstrate that kynurenine
pathway activation has effects on oxidative stress and dopamine regulation (Miller et al.,
2009).

Although stimuli-induced cytokine activation (e.g. IFNα) in depressed patients and pre-
clinical models has been shown to result in peripheral kynurenine pathway activation,
a recent meta-analysis surprisingly found decreased kynurenine pathway activation (e.g.
decreased KA and KYN measured in blood, urine, CSF, or brain tissue) in depressed
Author Manuscript

patients compared to healthy controls, although antidepressant-free depressed patients had


elevated QA levels compared to controls and blood levels of QA were elevated in depressed
patients compared to controls(O’connor et al., 2009; Ogyu et al., 2018). However, some of
these conflicting findings may be explained by differential regulation of peripheral TDO
from central IDO or differential brain-region regulation of the kynurenine pathway (Chen
et al., 2017). In fact, studies of CSF in suicidal and depressed patients more consistently
find elevations of QA, suggesting the kynurenine pathway may be more validly studied in
the CSF (Bryleva and Brundin, 2017; Raison et al., 2010). These studies typically do not
demonstrate reduction in CNS tryptophan and serotonin concentrations with kynurenine
pathway activation, consistent with studies showing dichotomy of kynurenine pathway
activation and CNS tryptophan depletion (Hughes et al., 2012; Raison et al., 2010; Wichers
et al., 2005). Thus, despite excellent pre-clinical and stimuli-induced models of kynurenine
Author Manuscript

pathway activation in depression, clinical studies of the peripheral kynurenine pathway


activation have yet to fully bear this out, and further studies are required to confirm results
from CSF kynurenine pathway studies.

4.7. Gut-brain axis activation and the microbiota


The gut microbiota is a dynamic intestinal reservoir of 1000+ bacterial species. Some
microbiota species have positive effects on health, while others are maladaptive (Rooks and
Garrett, 2016). A putative “gut-brain axis” exists, wherein alterations in gut microbiota are
linked to changes in behavior due to central nervous system effects (Foster and Neufeld,
2013). Evidence for gut-brain axis activity in depression is largely pre-clinical, though a few
clinical studies exist. Mice with absence of gut microbiota, termed “germ free mice,” have
abnormal stress responses, decreased central BDNF levels, differential central monoamine
Author Manuscript

metabolism, impairments in microglia activity, myelin protein dysregulation, and alterations


of cytokines such as IL-1 and IL-6 (Hoban et al., 2015; Liu et al., 2016; Sherwin et al.,
2016). Similarly, probiotic administration reduces depressive symptomatology in murine
models, while vagotomy eliminates these beneficial effects (Foster and Neufeld, 2013).
In depressed humans, clinical studies have found depressive symptoms associated with
presence of certain gut bacterial species such as Enterobacteriaceae (Jiang et al., 2015;
Naseribafrouei et al., 2014). In RCTs and observational studies in humans, preliminary

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 11

findings of efficacy of antidepressants for bacterial infections, efficacy of antibiotics for


Author Manuscript

depression, and efficacy of probiotics for depression also hint that microbiota have a role in
clinical depression (Huang et al., 2016; Macedo et al., 2017).

In murine and primate models, both acute and chronic stress lead to alterations in the
microbiota diversity and increase leakiness of the intestinal wall (Cryan and Dinan, 2012;
Foster and Neufeld, 2013; Sherwin et al., 2016). Then, this gut microbiota dysbiosis
leads to alteration-specific changes in bacterial species that may cause different changes
in immune cell populations, such as Th17 or regulatory T cells (Treg), and results in
variations in microbiota metabolites (Rooks and Garrett, 2016). Such changes in gut immune
cell populations may lead to systemic pro-inflammatory cytokine activation, including
IFNγ and TNFα, and vice versa (Rooks and Garrett, 2016). Microbiota metabolites
include neurotransmitters (e.g. 5-HT or tryptophan) and short chain fatty acids, and these
metabolites activate enteroendocrine cells in the gut, which interact directly with vagal
Author Manuscript

afferents via serotonin 5-HT3 receptors (Bonaz et al., 2018; Morrison and Preston, 2016).
Vagal afferent fibers project to the nucleus tractus solitarus (NTS) in the brainstem and
its downstream projections such as the hypothalamus and amygdala (Bonaz et al., 2018).
Ultimately, signals from vagal afferents may translate into depressive behavior and cytokine
release through effects on the HPA axis (via the hypothalamus) or fronto-limbic circuitry
(via the amygdala). Effects of microbiota metabolites on tryptophan catabolism may also
influence depressive symptomatology via the aforementioned kynurenine pathway.

One smaller quasi-clinical study found decreases in microbiome diversity in rats with
transplanted fecal microbiota from depressed human patients, although there were no
changes in rat plasma cytokine levels after transplantation (Kelly et al., 2016). One larger
clinical study and several preclinical studies suggest increases in gut permeability and
Author Manuscript

bacterial translocation based on peripheral immunoglobulin levels against gut bacteria in


depressed patients, although studies have not examined direct effects on cytokines (Maes
et al., 2012). Thus, based primarily on preclinical and theoretical evidence, microbiota
dysbiosis may lead to effects on other putative mechanistic pathways involved in cytokine
activation (e.g. HPA axis) via microbiota metabolites or lead to direct effects on cytokine
expression. Forthcoming research should examine temporal associations of microbiota
dysbiosis and cytokine changes in clinical samples, and ultimately if these are reflected
in clinical samples.

4.8. Reactive oxygen species (ROS) and reactive nitrogen species (RNS)
Reactive Oxygen Species (ROS; O2−, OH− ) and Reactive Nitrogen Species (RNS) are
short-lived molecules that are involved in numerous cell death pathways, and are commonly
Author Manuscript

released by macrophages via inducible nitric oxide synthase (iNOS) activation, cytokine
signals, and other cell death signals (Predonzani et al., 2015; Redza-Dutordoir and Averill-
Bates, 2016). ROS and RNS can directly lead to cell death through damage to cell
proteins, mitochondrial damage, damage to DNA, axonal damage, endoplasmic reticulum
stress protein damage, or lipid peroxidation (Redza-Dutordoir and Averill-Bates, 2016).
Meta-analyses have found that peripheral markers of oxidative damage are elevated in
depressed patients, particularly 8-hydroxy-2’-deoxyguanosine, malondialdehyde (MDA),

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 12

and F2-isoprostanes (Black et al., 2015; Lopresti et al., 2014). Concurrently, depressed
Author Manuscript

patients have been found to have a lower anti-oxidant “status” (e.g. having lower levels of
anti-oxidants such as zinc) that reflects higher levels of systemic ROS (Palta et al., 2014;
Swardfager et al., 2013).

RNS/ROS are bidirectionally associated with cytokine activation at a molecular level.


Cytokines such as TNFα, INFγ, and IL-1 may directly interfere with the electron transport
chain, induce iNOS or induce other pro-oxidant enzymes via NF-kβ activation (Chao et al.,
1996; Iborra et al., 2011; Papageorgiou et al., 2016). Additionally, TNFα binds to TNFαR1,
which activates the enzyme NAPDH Oxidase (NOX) through protein-protein interactions,
resulting in direct production of ROS (Blaser et al., 2016). Alternatively, oxidation may
induce the kynurenine pathway via direct activation of IDO, serotonin receptors, or NF-kβ
(Bakunina et al., 2015). Damage associated molecular pattern molecules (DAMPs) are
intracellular proteins that are activated via the RNS/ROS damaging effects on proteins
Author Manuscript

and cells, leading to cytokine activation when detected by macrophages. Some pre-clinical
studies have found mice with knockout of antioxidant-associated genes (e.g. Nrf2) have
both increased depressive symptoms and peripheral cytokine activation (e.g. TNFα, IL-6,
IL-10) while administration of certain anti-inflammatory agents in rats results in decreased
depressive symptoms, increased antioxidant parameters, and decreased peripheral cytokine
levels (e.g. TNFα, IL-6) (Thakare et al., 2017; Yao et al., 2016). However, few pre-clinical
or clinical studies have concurrently examined the existence of specific cytokine associations
with oxidative stress markers in depression, with unclear findings in regard to the association
of such cytokines and oxidative stress markers over the course of depression (Lindqvist et
al., 2017; Oglodek et al., 2017).

5 Mechanisms of cytokine induction by depression: where does the


Author Manuscript

inflammation come from?


In contrast to the prior section in which we explored mechanisms of cytokine-induced
depression, less clinical evidence exists for the other half of this relationship involving
depression-induced cytokine activation. As described earlier, these connections are likely
multifactorial and not attributable to one cause. We will focus on a few biological
factors given the larger evidence base for these factors, although other behavioral and
medical factors, including obesity, diet, pain, sleep, medical comorbidities, substance
use, and trauma, may be as important (see Supplemental Table 3). One limitation of
the conceptualization of depression as effecting change on cytokine activation is that
depression and cytokine activation are frequently characterized within these literatures as
dual concepts, whereas depression and cytokine activation may be part of a unitary construct
Author Manuscript

(Glannon, 2020; Miresco and Kirmayer, 2006). This alternative perspective highlights the
importance of research examining whether casual relationships between depression and
cytokine activation can be identified, e.g., where one variable is manipulated experimentally
and downstream changes in the other are then observed.

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 13

5.1. Chronic stress and HPA axis activation


Author Manuscript

Chronic stress provides a foundation to understanding the interactions of HPA axis


activation, cytokines and depression, and we summarize this linkage while recognizing
the caveats of simplification of complex neural mechanisms and the incomplete knowledge
base. Stress is broadly defined as a factor resulting in physical or mental tension, and
has been strongly linked with depressive symptomatology (Eisenberger and Cole, 2012;
Raison and Miller, 2013). Generally, the level of stress experienced in response to any
given context or environment may be related to sensory stimuli, ruminative processes,
cognitions, deficits in reward processes, or emotional memory (Belzung et al., 2015; Disner
et al., 2011; LaBar and Cabeza, 2006; Steimer, 2002; Treadway and Zald, 2011). In one
basic hypothesized model of stress, threat detection or stressor presence increases neuronal
activity in the cingulate gyrus, hippocampus, or amygdala and decreases neuronal activity in
the dorsolateral prefrontal cortex (dlPFC) (Disner et al., 2011). In this model of depression,
Author Manuscript

chronic activation of these pathways leads to complex remodeling of neural circuitry that
contribute to the changes in mood and cognition that characterize depressive disorders.
While there is evidence that depression results in direct activation of the HPA axis without
mediation via cytokines, acute stress models are also clearly associated with HPA axis
activation and peripheral cytokine elevations (primarily IL-6, IL-10, TNFα, and IL1β)
(Marsland et al., 2017b). Acutely, HPA axis activation may lead anti-inflammatory effects of
the HPA axis’ hormone cortisol which is well known to suppress cytokine activity (Padgett
and Glaser, 2003). However, chronic activation of HPA pathway via the chronic stressors
may lead to the changes in cytokines noted in MDD, with the evidence for this being largely
theoretical and preclinical (Supplemental Figure 7) (Calcia et al., 2016; Marsland et al.,
2017b; Tian et al., 2014).

Established by the described chronic stress-related neurocircuitry changes, PFC,


Author Manuscript

hippocampus, and amygdala engagement likely indirectly stimulate hypothalamic PVN cells
(Dedovic et al., 2009). These effects perhaps occur via the bed nucleus stria terminalis
(BNST) and other subcortical brain regions (e.g. dorsomedial hypothalamus and nucleus
tractus solitarius), although limited direct PVN stimulation may be present (Herman,
2012; Jankord and Herman, 2008; Lebow and Chen, 2016). As described previously,
stimulation of the PVN leads to HPA axis activation and cortisol release, which has been
well-characterized in MDD. In MDD, chronic HPA axis activation (secondary to stressors
present in depression) results in hypercortisolemia and glucocorticoid receptor resistance
(Pariante and Miller, 2001). Cortisol is important in suppressing pro-inflammatory cytokine
expression via increasing transcription of anti-inflammatory proteins such as IL-10 and
suppressor of cytokine signaling 1 (SOCS), while blocking transcription of inflammatory
transcription factors such as NF-kβ and activator protein 1 complexes (AP-1) (Desmet and
Author Manuscript

De Bosscher, 2017). In the post-translational milieu, glucocorticoids (such as cortisol) also


directly reduce TNFα cytokine mRNA stability and interfere with NF-kβ protein activity
via protein-protein interactions (Desmet and De Bosscher, 2017). With the glucocorticoid
receptor resistance present in MDD, cortisol may fail to have its typical anti-inflammatory
effects on macrophages and other immune cells, leading to cytokine overexpression
in these cells. Notably, contrary to findings of hypercortisolemia in some depressed
individuals, individuals with other subtypes of depression (e.g. atypical depression) have

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 14

been hypothesized to have hypocortisolemia, for which the above mechanisms may not be as
Author Manuscript

relevant (Lojko and Rybakowski, 2017).

Paradoxically, low but elevated glucocorticoid levels without concurrent glucocorticoid


resistance in certain cells may also lead to increased cytokine activation (Sorrells and
Sapolsky, 2007). Such effects may occur based on the immune cell/cytokine milieu,
neuronal cell type, or level of cortisol elevation. Preclinical studies often find less elevated
levels of glucocorticoids to have stimulatory effects on microglial cells or cytokines
(including IL-1β and macrophage migration inhibitory factor), while more elevated levels
of glucocorticoids tend to be immunosuppressive (Cain and Cidlowski, 2017; Sapolsky et
al., 2000; Tanaka et al., 1997). Additionally, differences in murine models with respect to
glucocorticoid receptor abundance and stress-mediated glucocorticoid receptor production
exist between and within brain regions such as the prefrontal cortex, hippocampus, and
hypothalamus (Mizoguchi et al., 2003; Roque et al., 2016; Zhang et al., 2017). For instance,
Author Manuscript

one murine study found that chronic stress resulted in elevations in cytosolic and nuclear
glucocorticoid receptor levels in the hippocampus but only resulted in decreases in cytosolic
glucocorticoid receptors in the prefrontal cortex (Mizoguchi et al., 2003). In regions of
the brain (e.g. hippocampus) with elevated cortisol, glucocorticoid activity may lead to
increased activation of cortisol receptors, followed by activation of endonucleases, DNA
cleavage, failure of protein synthesis, and ultimately cell death (Reagan and McEwen,
1997). Apoptosis and cell death lead to release of DAMPs, which are then recognized
by antigen presenting cells such as macrophages, which subsequently release a variety
of cytokines (Felger and Lotrich, 2013). Such effects are consistent with reports noting
cortisol-mediated hippocampal damage, as previously described.

In summary, hypercortisolemia may lead to cytokine activation (e.g. IL-6 and TNFα) via
Author Manuscript

failure of glucocorticoids’ anti-inflammatory activity in certain cells, and overactivity of


glucocorticoids in other cells. As described previously in section 4.5, few clinical studies
have addressed the cytokine-HPA axis interaction which would help clearly delineate this
relationship.

5.2. Stress and sympathetic nervous system activation


As the counterpart to the HPA axis, the sympathetic nervous system (SNS) is the
“fight or flight” part of the nervous system that is activated in response to a stressor,
predominantly releasing norepinephrine and epinephrine. In some hypothesized models
of depression, the chronic prefrontal and limbic neural circuitry changes in depression
(described previously) may activate SNS components including the locus coeruleus in the
brainstem, the hypothalamic paraventricular nucleus, and the rostral ventrolateral medulla
Author Manuscript

(rvML) (Aston-Jones and Waterhouse, 2016; Bellinger and Lorton, 2014; Dum et al., 2016).
Sympathetic preganglionic neurons originating from these activated brain regions then may
stimulate spinal cord neurons, which project to the adrenal medulla, leading to adrenal
chromaffin cell release of epinephrine and norepinephrine (Brown et al., 1982; Engeland
and Arnhold, 2005; Morrison and Cao, 2000). Given that clinical studies generally have
not examined the cytokine-SNS relationship in depressed patients, we will primarily review

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 15

theoretical and pre-clinical evidence, and future studies should attempt to examine clinical
Author Manuscript

associations of SNS activation with cytokine activity.

At both systemic and more localized zones of inflammation, acute activation of the SNS
results in suppression of cytokines such as IL-1β and TNFα via transcription factors
such as NF-kβ in microglia and peripheral macrophages, likely mediated by β2 adrenergic
receptor binding per preclinical models (Hasko and Szab o, 1998; Nance and Sanders,
2007; Scanzano and Cosentino, 2015). However, chronic activation of the SNS may be
associated with β adrenergic mediated transcription of pro-inflammatory cytokines such as
IL-6 (Ramirez et al., 2016; Rohan Walker et al., 2013). Additionally, low level α2 adrenergic
receptor activation (as opposed to greater adrenergic receptor activation) on monocytes or
macrophages is associated with release of certain cytokines (e.g. TNFα) (Chavan et al.,
2017; Pongratz and Straub, 2014).
Author Manuscript

The SNS also has indirect effects on the immune system that may predominate over
direct leukocyte adrenergic receptor effects. The SNS may increase peripheral or central
cytokine levels via chronic glucocorticoid release (due to adrenal cortex stimulation) or
vagal nerve hypoactivity (due to downregulation of parasympathetic activity) (Barnes et
al., 2015; Ulrich-Lai and Herman, 2009). Within the SNS, β1 adrenergic receptors on the
renal juxtaglomerular apparatus (JGA) activate the renin-angiotensin system (RAS). RAS
activates ROS/NOS pathways and binding of angiotensin II on AT1 receptors in CNS cells
such as microglia, which results in NF-kβ mediated cytokine transcription (primarily TNFα)
(Labandeira-Garcia et al., 2017). The involvement of this pathway in depression is buoyed
by preliminary observational clinical studies noting efficacy of angiotensin converting
enzyme inhibitors in reducing depressive symptoms (Vian et al., 2017).
Author Manuscript

5.3. Vagal nerve hypoactivity


As a key component of the parasympathetic nervous system (PNS), the vagal nerve may also
play a role in cytokine activation in depression. Depressed individuals appear to have altered
parasympathetic activity with possible vagal nerve hypoactivity (Kemp et al., 2010; Marvel
et al., 2004). Vagal nerve hypoactivity may be secondary to direct effects from depression
or from downstream effects of depression activating the SNS, which can suppress the PNS.
Consistent with that hypothesis, vagal nerve stimulation has been suggested as a treatment
for depression, with partial support from clinical trial data (Martin and Martin-Sanchez,
2012).

In recent years, the association between peripheral afferent and efferent vagal nerve
activation and reduction in cytokine production (primarily TNFα) in peripheral organs (e.g.
Author Manuscript

liver and spleen) as well as in macrophages, described as the “inflammatory reflex,” has
been elucidated (Supplemental Figure 8) (Tracey, 2002). Concurrent with strong reported
associations between gut pathology and the CNS, afferent vagal nerves in abdominal organs,
the digestive tract, and heart project to the NTS after sensing cytokines (e.g. IL-1) in the
periphery (Johnston and Webster, 2009). Efferent vagal nerves from the NTS project to
the thalamus and the dorsal motor vagal nucleus (DMVN) of the medulla, which also
receives projections from the amygdala and hypothalamus (Yuan and Silberstein, 2016).
The DMVN then projects parasympathetic nerve fibers to a variety of gastrointestinal and

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 16

lymphoid organs, including the liver and spleen. These vagal nerve efferents may synapse
Author Manuscript

with sympathetic postganglionic noradrenergic nerves in the spleen. In these nerves, splenic
norepinephrine release recruits T cells which then release AcH onto α7nAChR receptors
on macrophages to downregulate cytokine release (Dantzer, 2017). Alternatively, DMVN
derived parasympathetic fibers may send cholinergic signals directly to peripheral organs to
affect circulating monocytes or recruit immunosuppressive T cells. Molecularly, macrophage
α7nAChR receptor activation may lead to phosphorylation and consequent downregulation
of STAT3, which is a major transcription factor for both IL-6 and TNFα (Pavlov and Tracey,
2012).

Such findings are largely supported by research on endotoxemia in murine models, and
extant studies suggest that these pathways are likely present in humans (Dantzer, 2017;
Yuan and Silberstein, 2016). One study found association of parasympathetic dysregulation
and in vitro LPS-stimulated blood cytokine levels of IL-6 and TNFα in healthy controls,
Author Manuscript

although other studies of depressed patients suggest both positive and negative covariation of
vagal nerve activity with such peripheral cytokines (Corcoran et al., 2005; Hu et al., 2018;
Marsland et al., 2007). However, more directed studies are needed clinically to examine
naturalistic clinical associations of vagal nerve activity with cytokine changes in depression,
using reliable measures of vagal nerve activity.

5.4. Omega 3 polyunsaturated fatty acid deficits


Omega-3 polyunsaturated fatty acids (ω-3 PUFAs) are natural anti-inflammatory fatty
acids present in foods such as fish and nuts. ω-3 PUFAs include docosahexaenoic acid
(DHA) and eicosapentaenoic acid (EPA). Decreased intake of ω-3 PUFAs in diet may be
secondary to appetite disturbances commonly seen in depression and has been associated
with a higher risk of development of depression (Grosso et al., 2016). In preclinical
Author Manuscript

studies, the peripheral anti-inflammatory mechanisms of ω-3 PUFAs are well described,
with theoretically similar effects in the CNS (Supplemental Figure 9). ω-3 PUFAs may
bind to cellular G protein receptors such as G-protein coupled receptor (GPR120), which
blocks NF-kβ and c-Jun N-terminal kinase (JNK) pathways, resulting in a decrease in
canonical NF-kB dependent cytokines (e.g. IL-6 and TNFα) (Calder, 2010; Oh et al.,
2010). Alternatively, ω-3 PUFAs may activate peroxisome proliferator-activated receptors
(PPARs), which bind and inactivate cytokine transcription factors such as NF-kβ, STAT, or
AP-1 (Calder, 2010; Tyagi et al., 2011). Lastly, ω-3 PUFAs upregulate anti-inflammatory
proteins such as resolvins and protectins, which mediate effects potentially via transcription
factors (e.g. NF-kβ) or micro RNA (miRNA) pathways that have demonstrated reduction of
LPS-induced proinflammatory cytokine expression (e.g. IL-6, TNF-α, or IL-1β) (Rey et al.,
2016).
Author Manuscript

In a few meta-analyses, ω-3 PUFA supplementation demonstrated beneficial effects as an


adjunctive pharmacologic treatment for MDD, although studies tended to be heterogenous,
smaller, and demonstrated greater benefits for EPA as compared to DHA (Mocking et al.,
2017). Such benefits appear greater for those with elevated cytokine activation. For example,
a recent trial found that in patients with MDD and higher relative levels of IL-1RA and
hs-CRP, individuals experienced greater reductions in depression severity when treated with

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 17

ω-3 PUFAs compared to patients with MDD with lower relative levels of plasma IL-1RA
Author Manuscript

and hs-CRP (Rapaport et al., 2016). Extant studies in non-depressed populations suggest
that higher plasma ω-3 PUFAs are associated with lower levels of plasma IL-6, IL-1RA,
and TNFα and higher levels of plasma IL-10 (Ferrucci et al., 2006). Although preliminary
clinical observations of depressed patients similarly suggest association of lower ω-3 PUFA
plasma levels with elevations in plasma TNFα and IL-6, studies have not sufficiently
replicated such ω-3 PUFA and pro-inflammatory cytokine associations in depressed cohorts
(Kiecolt-Glaser et al., 2007; Lai et al., 2016).

6. Discussion, areas of future research, and limitations


In summary, cytokine related induction of depressive symptomology may directly occur
due to the effects of cytokines on neurotransmitters, brain regions, or neural networks
(Supplemental Figure 10). Indirectly, cytokines may also activate large physiological
Author Manuscript

networks such as the HPA axis and the gut-brain axis to trigger depression pathways.
From a molecular perspective, cytokines may also activate the kynurenine pathway or
ROS/NOS pathways, leading to accumulation of neurotoxic metabolites. At the same time,
depressive symptoms may lead to activation of the HPA axis or SNS overactivity (and
concurrent vagal nerve hypoactivity), which lead to increase in pro-inflammatory cytokine
levels, enhanced by deficits in omega 3 fatty acids. However, other factors associated
with depression, such as obesity, diet, pain, sleep, medical comorbidities, substance use,
and trauma, may play a role in cytokine activation but require more research. The
mechanism with the strongest evidence for cytokine-induced depression likely relates to the
effects of cytokines on neurotransmitters and neural networks; fair preclinical and clinical
evidence exists as well for the role of cytokines on the HPA axis. As one feature in a
multifactorial stress-diathesis model, depressive symptoms may lead to pro-inflammatory
Author Manuscript

cytokine activation, which in turn, may contribute to worsening depressive symptoms. In


conjunction, pro-inflammatory cytokine activation may contribute to depressive symptoms
through the described mechanistic pathways.

Although recent research has helped to elucidate the mechanistic interactions between
depression and cytokine activation, important areas of future research exist. First,
methodologically robust clinical research examining these mechanisms in human
populations is needed, given that clinical depression is typically heterogeneous, episodic,
and not always treatment-responsive, as compared to preclinical models. Given the
heterogeneity of depression, multiple mechanisms may be contributing to cytokine
involvement in depression. Noting the numerous pre-clinical studies and relative dearth
of large cross-sectional or longitudinal clinical studies, future studies would best focus on
Author Manuscript

clinical populations with careful attention to the weight of various moderating factors on
cytokine activation and specificity of different cytokines. Long term clinical research could
identify aberrations in inflammatory/cytokine networks in depressed individuals, rather than
just individual cytokine levels, given that cytokines are constantly interacting with each
other and other parts of the immune system. Ultimately, such research could help guide
identification of treatments for depression. Further exploration is required to investigate
modulating factors of cytokine activity in depression, including unique cytokine effects,
duration of cytokine alterations, cellular targets of cytokines, and brain region-dependent

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 18

variations in cytokine effects. Clarification of the time, cell and location dependent effects of
Author Manuscript

cytokines is critical to recognizing moderating factors of cytokine activation. Given cytokine


alterations reported in other psychiatric disorders such as schizophrenia and autism, better
understanding of depression-specific cytokine associations is also needed. As previously
described, an important critique of the conceptualization in contemporary frameworks of
depression and cytokine activation is that they are dualistic in nature. If biological and
psychological aspects of depression are part of a unitary construct experienced by some
humans (Glannon, 2020; Miresco and Kirmayer, 2006), then any debate regarding which is
cause and which is effect is nonsensical and can’t be resolved.

7. Conclusions
As the inflammatory hypothesis of depression approaches its 30th anniversary, the
importance of understanding the contribution of cytokines to depressive symptomology
Author Manuscript

continues to grow. Researchers have concretely identified elevations of plasma IL-6, CRP,
and TNFα in depressed patients as compared to healthy patients, and the relationship
between cytokines and depression appears to be bidirectional. As described, numerous
mechanisms exist that may mediate interactions between depression and cytokine activation.
Next steps for future research involve methodologically rigorous observational studies and
clinical trials examining these relationships in close detail with attention to confounders and
focusing on the specificity of cytokine effects.

Supplementary Material
Refer to Web version on PubMed Central for supplementary material.
Author Manuscript

References
Amodeo G, Trusso MA, Fagiolini A, 2017. Depression and inflammation: disentangling a clear yet
complex and multifaceted link. Neuropsychiatry 7, 448–457.
Anisman H, Ravindran AV, Griffiths J, Merali Z, 1999. Endocrine and cytokine correlates of major
depression and dysthymia with typical or atypical features. Mol. Psychiatry 4, 182–188. [PubMed:
10208451]
Aston-Jones G, Waterhouse B, 2016. Locus coeruleus: From global projection system to adaptive
regulation of behavior. Brain Res. 1645, 75–78. [PubMed: 26969408]
Badawy AA, 2017. Kynurenine Pathway of Tryptophan Metabolism: Regulatory and Functional
Aspects. Int. J. Tryptophan Res 10, 1178646917691938.
Bakunina N, Pariante CM, Zunszain PA, 2015. Immune mechanisms linked to depression via oxidative
stress and neuroprogression. Immunology 144, 365–373. [PubMed: 25580634]
Banks WA, 2005. Blood-brain barrier transport of cytokines: a mechanism for neuropathology. Curr.
Pharm. Des 11, 973–984. [PubMed: 15777248]
Author Manuscript

Barnes MA, Carson MJ, Nair MG, 2015. Non-traditional cytokines: How catecholamines and
adipokines influence macrophages in immunity, metabolism and the central nervous system.
Cytokine 72, 210–219. [PubMed: 25703786]
Baune BT, Konrad C, Grotegerd D, Suslow T, Ohrmann P, Bauer J, Arolt V, Heindel W, Domschke
K, Schoning S, Rauch AV, Sehlmeyer C, Kugel H, Dannlowski U, 2012a. Tumor necrosis factor
gene variation predicts hippocampus volume in healthy individuals. Biol. Psychiatry 72, 655–662.
[PubMed: 22554453]
Baune BT, Smith E, Reppermund S, Air T, Samaras K, Lux O, Brodaty H, Sachdev P, Trollor
JN, 2012b. Inflammatory biomarkers predict depressive, but not anxiety symptoms during aging:

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 19

the prospective Sydney Memory and Aging Study. Psychoneuroendocrinology 37, 1521–1530.
[PubMed: 22406002]
Author Manuscript

Bellinger DL, Lorton D, 2014. Autonomic regulation of cellular immune function. Auton. Neurosci
182, 15–41. [PubMed: 24685093]
Belzung C, Willner P, Philippot P, 2015. Depression: from psychopathology to pathophysiology. Curr.
Opin. Neurobiol 30, 24–30. [PubMed: 25218233]
Bethin KE, Vogt SK, Muglia LJ, 2000. Interleukin-6 is an essential, corticotropin-releasing hormone-
independent stimulator of the adrenal axis during immune system activation, 97. Proceedings of
the National Academy of Sciences of the United States of America, pp. 9317–9322. [PubMed:
10922080]
Black CN, Bot M, Scheffer PG, Cuijpers P, Penninx BW, 2015. Is depression associated with increased
oxidative stress? A systematic review and meta-analysis. Psychoneuroendocrinology 51, 164–175.
[PubMed: 25462890]
Blaser H, Dostert C, Mak TW, Brenner D, 2016. TNF and ROS Crosstalk in Inflammation. Trends Cell
Biol. 26, 249–261. [PubMed: 26791157]
Bonaz B, Bazin T, Pellissier S, 2018. The Vagus Nerve at the Interface of the Microbiota-Gut-Brain
Author Manuscript

Axis. Front Neurosci 12, 49. [PubMed: 29467611]


Borsini A, Zunszain PA, Thuret S, Pariante CM, 2015. The role of inflammatory cytokines as key
modulators of neurogenesis. Trends Neurosci. 38, 145–157. [PubMed: 25579391]
Brown MR, Fisher LA, Spiess J, Rivier C, Rivier J, Vale W, 1982. Corticotropin-releasing factor:
actions on the sympathetic nervous system and metabolism. Endocrinology 111, 928–931.
[PubMed: 7049676]
Bryleva EY, Brundin L, 2017. Kynurenine pathway metabolites and suicidality. Neuropharmacology
112, 324–330. [PubMed: 26820800]
Burke HM, Davis MC, Otte C, Mohr DC, 2005. Depression and cortisol responses to psychological
stress: A meta-analysis. Psychoneuroendocrinology 30, 846–856. [PubMed: 15961250]
Cain DW, Cidlowski JA, 2017. Immune regulation by glucocorticoids. Nat. Rev. Immunol 17, 233–
247. [PubMed: 28192415]
Calcia MA, Bonsall DR, Bloomfield PS, Selvaraj S, Barichello T, Howes OD, 2016. Stress and
neuroinflammation: a systematic review of the effects of stress on microglia and the implications
Author Manuscript

for mental illness. Psychopharmacology (Berl.) 233, 1637–1650. [PubMed: 26847047]


Calder PC, 2010. Omega-3 fatty acids and inflammatory processes. Nutrients 2, 355–374. [PubMed:
22254027]
Capuron L, Miller AH, 2011. Immune system to brain signaling: neuropsychopharmacological
implications. Pharmacol. Ther 130, 226–238. [PubMed: 21334376]
Capuron L, Ravaud A, Gualde N, Bosmans E, Dantzer R, Maes M, Neveu PJ, 2001. Association
between immune activation and early depressive symptoms in cancer patients treated with
interleukin-2-based therapy. Psychoneuroendocrinology 26, 797–808. [PubMed: 11585680]
Carpenter LL, Heninger GR, Malison RT, Tyrka AR, Price LH, 2004. Cerebrospinal fluid interleukin
(IL)-6 in unipolar major depression. J. Affect. Disord 79, 285–289. [PubMed: 15023509]
Chao CC, Hu S, Sheng WS, Bu D, Bukrinsky MI, Peterson PK, 1996. Cytokine-stimulated astrocytes
damage human neurons via a nitric oxide mechanism. Glia 16, 276–284. [PubMed: 8833198]
Chavan SS, Pavlov VA, Tracey KJ, 2017. Mechanisms and Therapeutic Relevance of Neuro-immune
Communication. Immunity 46, 927–942. [PubMed: 28636960]
Author Manuscript

Chen Y, Xie Z, Xiao C, Zhang M, Li Z, Xie J, Zhang Y, Zhao X, Zeng P, Mo L, 2017. Peripheral
kynurenine/tryptophan ratio is not a reliable marker of systemic indoleamine 2, 3-dioxygenase: A
lesson drawn from patients on hemodialysis. Oncotarget 8, 25261. [PubMed: 28445957]
Corcoran C, Connor TJ, O’Keane V, Garland MR, 2005. The effects of vagus nerve stimulation on
pro-and anti-inflammatory cytokines in humans: a preliminary report. NeuroImmuno-Modulation
12, 307–309.
Cryan JF, Dinan TG, 2012. Mind-altering microorganisms: the impact of the gut microbiota on brain
and behaviour. Nat. Rev. Neurosci 13, 701–712. [PubMed: 22968153]

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 20

Cuartero MI, de la Parra J, Garcia-Culebras A, Ballesteros I, Lizasoain I, Moro MA, 2016. The
Kynurenine Pathway in the Acute and Chronic Phases of Cerebral Ischemia. Curr. Pharm. Des 22,
Author Manuscript

1060–1073. [PubMed: 25248805]


Dahl J, Ormstad H, Aass H, Sandvik L, Malt U, Andreassen O, 2016. Recovery from major depressive
disorder episode after non-pharmacological treatment is associated with normalized cytokine
levels. Acta Psychiatr. Scand 134, 40–47. [PubMed: 27028967]
Dantzer R, 2001. Cytokine-induced sickness behavior: mechanisms and implications. Ann. N. Y. Acad.
Sci 933, 222–234. [PubMed: 12000023]
Dantzer R, 2017. Neuroimmune interactions: from the brain to the immune system and vice versa.
Physiol. Rev 98, 477–504.
Dantzer R, O’Connor JC, Freund GG, Johnson RW, Kelley KW, 2008. From inflammation to sickness
and depression: when the immune system subjugates the brain. Nat. Rev. Neurosci 9, 46–56.
[PubMed: 18073775]
Dean B, Tawadros N, Scarr E, Gibbons AS, 2010. Regionally-specific changes in levels of tumour
necrosis factor in the dorsolateral prefrontal cortex obtained postmortem from subjects with major
depressive disorder. J. Affect. Disord 120, 245–248. [PubMed: 19446343]
Author Manuscript

Dedovic K, Duchesne A, Andrews J, Engert V, Pruessner JC, 2009. The brain and the stress axis: the
neural correlates of cortisol regulation in response to stress. Neuroimage 47, 864–871. [PubMed:
19500680]
Desmet SJ, De Bosscher K, 2017. Glucocorticoid receptors: finding the middle ground. J. Clin. Invest
127, 1136–1145. [PubMed: 28319043]
Disner SG, Beevers CG, Haigh EA, Beck AT, 2011. Neural mechanisms of the cognitive model of
depression. Nat. Rev. Neurosci 12, 467–477. [PubMed: 21731066]
Dowlati Y, Herrmann N, Swardfager W, Liu H, Sham L, Reim EK, Lanctot KL, 2010. A meta-analysis
of cytokines in major depression. Biol. Psychiatry 67, 446–457. [PubMed: 20015486]
Downen M, Amaral TD, Hua LL, Zhao ML, Lee SC, 1999. Neuronal death in cytokine-activated
primary human brain cell culture: role of tumor necrosis factor-alpha. Glia 28, 114–127. [PubMed:
10533055]
Dugan LL, Ali SS, Shekhtman G, Roberts AJ, Lucero J, Quick KL, Behrens MM, 2009. IL-6 mediated
degeneration of forebrain GABAergic interneurons and cognitive impairment in aged mice through
Author Manuscript

activation of neuronal NADPH oxidase. PLoS One 4, e5518. [PubMed: 19436757]


Dum RP, Levinthal DJ, Strick PL, 2016. Motor, cognitive, and affective areas of the cerebral cortex
influence the adrenal medulla. In: Proceedings of the National Academy of Sciences of the United
States of America, 113, pp. 9922–9927. [PubMed: 27528671]
Eisenberger NI, Cole SW, 2012. Social neuroscience and health: neurophysiological mechanisms
linking social ties with physical health. Nat. Neurosci 15, 669–674. [PubMed: 22504347]
Engeland WC, Arnhold MM, 2005. Neural circuitry in the regulation of adrenal corticosterone
rhythmicity. Endocrine 28, 325–332. [PubMed: 16388123]
Erta M, Quintana A, Hidalgo J, 2012. Interleukin-6, a major cytokine in the central nervous system.
Int. J. Biol. Sci 8, 1254–1266.
Fekadu A, Wooderson SC, Markopoulo K, Donaldson C, Papadopoulos A, Cleare AJ, 2009. What
happens to patients with treatment-resistant depression? A systematic review of medium to long
term outcome studies. J. Affect. Disord 116, 4–11. [PubMed: 19007996]
Felger JC, Haroon E, Patel TA, Goldsmith DR, Wommack EC, Woolwine BJ, Le N-A, Feinberg
R, Tansey MG, Miller AH, 2018. What does plasma CRP tell us about peripheral and central
Author Manuscript

inflammation in depression? Molecular Psychiatry.


Felger JC, Li Z, Haroon E, Woolwine BJ, Jung MY, Hu X, Miller AH, 2016. Inflammation
is associated with decreased functional connectivity within corticostriatal reward circuitry in
depression. Mol. Psychiatry 21, 1358. [PubMed: 26552591]
Felger JC, Lotrich FE, 2013. Inflammatory cytokines in depression: neurobiological mechanisms and
therapeutic implications. Neuroscience 246, 199–229. [PubMed: 23644052]
Ferri CC, Ferguson AV, 2003. Interleukin-1 beta depolarizes paraventricular nucleus parvocellular
neurones. J. Neuroendocrinol 15, 126–133. [PubMed: 12535154]

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 21

Ferrucci L, Cherubini A, Bandinelli S, Bartali B, Corsi A, Lauretani F, Martin A, Andres-Lacueva


C, Senin U, Guralnik JM, 2006. Relationship of plasma polyunsaturated fatty acids to circulating
Author Manuscript

inflammatory markers. J. Clin. Endocrinol. Metab 91, 439–446. [PubMed: 16234304]


Foster JA, Neufeld K-AM, 2013. Gut–brain axis: how the microbiome influences anxiety and
depression. Trends Neurosci. 36, 305–312. [PubMed: 23384445]
Frodl T, Amico F, 2014. Is there an association between peripheral immune markers and structural/
functional neuroimaging findings? Prog. Neuropsychopharmacol. Biol. Psychiatry 48, 295–303.
[PubMed: 23313563]
Fukui S, Schwarcz R, Rapoport SI, Takada Y, Smith QR, 1991. Blood-brain barrier transport of
kynurenines: implications for brain synthesis and metabolism. J. Neurochem 56, 2007–2017.
[PubMed: 1827495]
Gimeno D, Kivimaki M, Brunner EJ, Elovainio M, De Vogli R, Steptoe A, Kumari M, Lowe GDO,
Rumley A, Marmot MG, Ferrie JE, 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. [PubMed: 18533059]
Glannon W, 2020. Mind-brain dualism in psychiatry: ethical implications. Front. Psychiatry 11.
Author Manuscript

Glaus J, von Kanel R, Lasserre AM, Strippoli MF, Vandeleur CL, Castelao E, Gholam-Rezaee M,
Marangoni C, Wagner EN, Marques-Vidal P, Waeber G, Vollenweider P, Preisig M, Merikangas
KR, 2018. Mood disorders and circulating levels of inflammatory markers in a longitudinal
population-based study. Psychol. Med 48, 961–973. [PubMed: 28929992]
Grosso G, Micek A, Marventano S, Castellano S, Mistretta A, Pajak A, Galvano F, 2016. Dietary n-3
PUFA, fish consumption and depression: A systematic review and meta-analysis of observational
studies. J. Affect. Disord 205, 269–281. [PubMed: 27544316]
Hamilton JP, Farmer M, Fogelman P, Gotlib IH, 2015. Depressive Rumination, the Default-Mode
Network, and the Dark Matter of Clinical Neuroscience. Biol. Psychiatry 78, 224–230. [PubMed:
25861700]
Hamilton JP, Siemer M, Gotlib IH, 2008. Amygdala volume in major depressive disorder: a meta-
analysis of magnetic resonance imaging studies. Mol. Psychiatry 13, 993–1000. [PubMed:
18504424]
Haroon E, Chen X, Li Z, Patel T, Woolwine BJ, Hu XP, Felger JC, Miller AH, 2018. Increased
inflammation and brain glutamate define a subtype of depression with decreased regional
Author Manuscript

homogeneity, impaired network integrity, and anhedonia. Transl. Psychiatry 8, 189. [PubMed:
30202011]
Haroon E, Miller AH, 2016. Inflammation effects on brain glutamate in depression: mechanistic
considerations and treatment implications, Inflammation-Associated Depression: Evidence,
Mechanisms and Implications. Springer, pp. 173–198.
Hasko G, Sza bo C, 1998. Regulation of cytokine and chemokine production by transmitters and
co-transmitters of the autonomic nervous system. Biochem. Pharmacol 56, 1079–1087. [PubMed:
9802316]
Henriksson MM, Aro HM, Marttunen MJ, Heikkinen ME, Isometsa ET, Kuoppasalmi KI, Lonnqvist
JK, 1993. Mental disorders and comorbidity in suicide. Am. J. Psychiatry 150, 935–940. [PubMed:
8494072]
Herman JP, 2012. Neural pathways of stress integration: relevance to alcohol abuse. Alcohol Res. 34,
441–447. [PubMed: 23584110]
Herrmann LL, Le Masurier M, Ebmeier KP, 2008. White matter hyperintensities in late life depression:
Author Manuscript

a systematic review. J. Neurol. Neurosurg. Psychiatry 79, 619–624. [PubMed: 17717021]


Hoban A, Stilling R, Desbonnet L, Shanahan F, Dinan T, Cryan J, Clarke G, 2015. Regulation of
myelination in the prefrontal cortex by the gut microbiota: implications for health and disease.
FASEB J. 29, 672, 674.
Hodes GE, Kana V, Menard C, Merad M, Russo SJ, 2015. Neuroimmune mechanisms of depression.
Nat. Neurosci 18, 1386–1393. [PubMed: 26404713]
Howren MB, Lamkin DM, Suls J, 2009. Associations of depression with C-reactive protein, IL-1, and
IL-6: a meta-analysis. Psychosom. Med 71, 171–186. [PubMed: 19188531]

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 22

Hu MX, Penninx B, de Geus EJC, Lamers F, Kuan DC, Wright AGC, Marsland AL, Muldoon MF,
Manuck SB, Gianaros PJ, 2018. Associations of immunometabolic risk factors with symptoms
Author Manuscript

of depression and anxiety: The role of cardiac vagal activity. Brain Behav. Immun 73, 493–503.
[PubMed: 29920329]
Huang M, Su S, Goldberg J, Miller AH, Levantsevych OM, Shallenberger L, Pimple P, Pearce
B, Bremner JD, Vaccarino V, 2018. Longitudinal association of inflammation with depressive
symptoms: A 7-year cross-lagged twin difference study. Brain Behav. Immun
Huang R, Wang K, Hu J, 2016. Effect of Probiotics on Depression: A Systematic Review and
Meta-Analysis of Randomized Controlled Trials. Nutrients 8, 483.
Hughes MM, Carballedo A, McLoughlin DM, Amico F, Harkin A, Frodl T, Connor TJ, 2012.
Tryptophan depletion in depressed patients occurs independent of kynurenine pathway activation.
Brain Behav. Immun 26, 979–987. [PubMed: 22683764]
Iborra M, Moret I, Rausell F, Bastida G, Aguas M, Cerrillo E, Nos P, Beltran B, 2011. Role of
oxidative stress and antioxidant enzymes in Crohn’s disease. Portland Press Limited.
Jankord R, Herman JP, 2008. Limbic regulation of hypothalamo-pituitary-adrenocortical function
during acute and chronic stress. Ann. N. Y. Acad. Sci 1148, 64–73. [PubMed: 19120092]
Author Manuscript

Jeon SW, Kim YK, 2016. Neuroinflammation and cytokine abnormality in major depression: cause or
consequence in that illness? World journal of psychiatry 6, 283. [PubMed: 27679767]
Jiang H, Ling Z, Zhang Y, Mao H, Ma Z, Yin Y, Wang W, Tang W, Tan Z, Shi J, Li L, Ruan B, 2015.
Altered fecal microbiota composition in patients with major depressive disorder. Brain Behav.
Immun 48, 186–194. [PubMed: 25882912]
Johnston GR, Webster NR, 2009. Cytokines and the immunomodulatory function of the vagus nerve.
Br. J. Anaesth 102, 453–462. [PubMed: 19258380]
Kaestner F, Hettich M, Peters M, Sibrowski W, Hetzel G, Ponath G, Arolt V, Cassens U, Rothermundt
M, 2005. Different activation patterns of proinflammatory cytokines in melancholic and non-
melancholic major depression are associated with HPA axis activity. J. Affect. Disord 87, 305–
311. [PubMed: 15951024]
Kaiser RH, Andrews-Hanna JR, Wager TD, Pizzagalli DA, 2015. Large-Scale Network Dysfunction
in Major Depressive Disorder: A Meta-analysis of Resting-State Functional Connectivity. JAMA
Psychiatry 72, 603–611. [PubMed: 25785575]
Author Manuscript

Kappelmann N, Lewis G, Dantzer R, Jones PB, Khandaker GM, 2018. Antidepressant activity of
anti-cytokine treatment: a systematic review and meta- analysis of clinical trials of chronic
inflammatory conditions. Mol. Psychiatry 23, 335–343. [PubMed: 27752078]
Karlovic D, Serretti A, Vrkic N, Martinac M, Marcinko D, 2012. Serum concentrations of CRP,
IL-6, TNF-alpha and cortisol in major depressive disorder with melancholic or atypical features.
Psychiatry Res. 198, 74–80. [PubMed: 22386567]
Kelly JR, Borre Y, C OB, Patterson E, El Aidy S, Deane J, Kennedy PJ, Beers S, Scott K, Moloney
G, Hoban AE, Scott L, Fitzgerald P, Ross P, Stanton C, Clarke G, Cryan JF, Dinan TG, 2016.
Transferring the blues: Depression-associated gut microbiota induces neurobehavioural changes in
the rat. J. Psychiatr. Res 82, 109–118. [PubMed: 27491067]
Kemp AH, Quintana DS, Gray MA, Felmingham KL, Brown K, Gatt JM, 2010. Impact of depression
and antidepressant treatment on heart rate variability: a review and meta-analysis. Biol. Psychiatry
67, 1067–1074. [PubMed: 20138254]
Khandaker GM, Pearson RM, Zammit S, Lewis G, Jones PB, 2014. Association of serum interleukin
6 and C-reactive protein in childhood with depression and psychosis in young adult life: a
Author Manuscript

population-based longitudinal study. JAMA Psychiatry 71, 1121–1128. [PubMed: 25133871]


Kiecolt-Glaser JK, Belury MA, Porter K, Beversdorf D, Lemeshow S, Glaser R, 2007. Depressive
symptoms, n-6: n-3 fatty acids, and inflammation in older adults. Psychosom. Med 69, 217.
[PubMed: 17401057]
Kim YK, Na KS, Myint AM, Leonard BE, 2016. The role of pro-inflammatory cytokines in
neuroinflammation, neurogenesis and the neuroendocrine system in major depression. Prog.
Neuropsychopharmacol. Biol. Psychiatry 64, 277–284. [PubMed: 26111720]

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 23

Knorr U, Vinberg M, Kessing LV, Wetterslev J, 2010. Salivary cortisol in depressed patients versus
control persons: a systematic review and meta-analysis. Psychoneuroendocrinology 35, 1275–
Author Manuscript

1286. [PubMed: 20447770]


Köhler CA, Freitas TH, Stubbs B, Maes M, Solmi M, Veronese N, de Andrade NQ, Morris G,
Fernandes BS, Brunoni AR, 2018. Peripheral alterations in cytokine and chemokine levels after
antidepressant drug treatment for major depressive disorder: systematic review and meta-analysis.
Mol. Neurobiol 55, 4195–4206. [PubMed: 28612257]
Köhler O, Benros ME, Nordentoft M, Farkouh ME, Iyengar RL, Mors O, Krogh J, 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. [PubMed: 25322082]
Kohman RA, Rhodes JS, 2013. Neurogenesis, inflammation and behavior. Brain Behav. Immun 27,
22–32. [PubMed: 22985767]
Konsman JP, Parnet P, Dantzer R, 2002. Cytokine-induced sickness behaviour: mechanisms and
implications. Trends Neurosci. 25, 154–159. [PubMed: 11852148]
Kraynak TE, Marsland AL, Wager TD, Gianaros PJ, 2018. Functional neuroanatomy of peripheral
Author Manuscript

inflammatory physiology: A meta-analysis of human neuroimaging studies. Neurosci. Biobehav.


Rev 94, 76–92. [PubMed: 30067939]
Krishnan V, Nestler EJ, 2008. The molecular neurobiology of depression. Nature 455, 894. [PubMed:
18923511]
Kubera M, Obuchowicz E, Goehler L, Brzeszcz J, Maes M, 2011. In animal models, psychosocial
stress-induced (neuro) inflammation, apoptosis and reduced neurogenesis are associated to the
onset of depression. Prog. Neuropsychopharmacol. Biol. Psychiatry 35, 744–759.
Kunugi H, Hori H, Ogawa S, 2015. Biochemical markers subtyping major depressive disorder.
Psychiatry Clin. Neurosci 69, 597–608. [PubMed: 25825158]
Labandeira-Garcia JL, Rodriguez-Perez AI, Garrido-Gil P, Rodriguez-Pallares J, Lanciego JL, Guerra
MJ, 2017. Brain Renin-Angiotensin System and Microglial Polarization: Implications for Aging
and Neurodegeneration. Front Aging Neurosci 9, 129. [PubMed: 28515690]
LaBar KS, Cabeza R, 2006. Cognitive neuroscience of emotional memory. Nat. Rev. Neurosci 7,
54–64. [PubMed: 16371950]
Author Manuscript

Labrenz F, Wrede K, Forsting M, Engler H, Schedlowski M, Elsenbruch S, Benson S, 2016.


Alterations in functional connectivity of resting state networks during experimental endotoxemia
-An exploratory study in healthy men. Brain Behav. Immun 54, 17–26. [PubMed: 26597151]
Lai JS, Oldmeadow C, Hure AJ, McEvoy M, Hiles SA, Boyle M, Attia J, 2016. Inflammation mediates
the association between fatty acid intake and depression in older men and women. Nutr. Res 36,
234–245. [PubMed: 26923510]
Lamers F, Milaneschi Y, Smit JH, Schoevers RA, Wittenberg G, Penninx BW, 2019. Longitudinal
Association Between Depression and Inflammatory Markers: Results From the Netherlands
Study of Depression and Anxiety. Biol. Psychiatry
Lebow M, Chen A, 2016. Overshadowed by the amygdala: the bed nucleus of the stria terminalis
emerges as key to psychiatric disorders. Mol. Psychiatry 21, 450. [PubMed: 26878891]
Levine J, Barak Y, Chengappa K, Rapoport A, Rebey M, Barak V, 1999. Cerebrospinal cytokine levels
in patients with acute depression. Neuropsychobiology 40, 171–176. [PubMed: 10559698]
Lindqvist D, Dhabhar FS, James SJ, Hough CM, Jain FA, Bersani FS, Réus VI, Verhoeven JE, Epel
ES, Mahan L, Rosser R, Wolkowitz OM, Mellon SH, 2017. Oxidative stress, inflammation
Author Manuscript

and treatment response in major depression. Psychoneuroendocrinology 76, 197–205. [PubMed:


27960139]
Liu WH, Chuang HL, Huang YT, Wu CC, Chou GT, Wang S, Tsai YC, 2016. Alteration of behavior
and monoamine levels attributable to Lactobacillus plantarum PS128 in germ-free mice. Behav.
Brain Res 298, 202–209. [PubMed: 26522841]
Lojko D, Rybakowski JK, 2017. Atypical depression: current perspectives. Neuropsychiatr. Dis. Treat
13, 2447–2456. [PubMed: 29033570]

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 24

Lopresti AL, Maker GL, Hood SD, Drummond PD, 2014. A review of peripheral biomarkers
in major depression: the potential of inflammatory and oxidative stress biomarkers. Prog.
Author Manuscript

Neuropsychopharmacol. Biol. Psychiatry 48, 102–111. [PubMed: 24104186]


Lorenzetti V, Allen NB, Fornito A, Yucel M, 2009. Structural brain abnormalities in major depressive
disorder: a selective review of recent MRI studies. J. Affect. Disord 117, 1–17. [PubMed:
19237202]
Mac Giollabhui N, Ng TH, Ellman LM, Alloy LB, 2020. The longitudinal associations of
inflammatory biomarkers and depression revisited: systematic review, meta-analysis, and meta-
regression. Mol. Psychiatry 1–13.
Macedo D, Filho A, Soares de Sousa CN, Quevedo J, Barichello T, Junior HVN, Freitas de Lucena
D, 2017. Antidepressants, antimicrobials or both? Gut microbiota dysbiosis in depression and
possible implications of the antimicrobial effects of antidepressant drugs for antidepressant
effectiveness. J. Affect. Disord 208, 22–32. [PubMed: 27744123]
Maes M, Kubera M, Leunis JC, Berk M, 2012. Increased IgA and IgM responses against gut
commensals in chronic depression: further evidence for increased bacterial translocation or leaky
gut. J. Affect. Disord 141, 55–62. [PubMed: 22410503]
Author Manuscript

Marsland AL, Gianaros PJ, Prather AA, Jennings JR, Neumann SA, Manuck SB, 2007. Stimulated
production of proinflammatory cytokines covaries inversely with heart rate variability.
Psychosom. Med 69, 709–716. [PubMed: 17942840]
Marsland AL, Kuan DC, Sheu LK, Krajina K, Kraynak TE, Manuck SB, Gianaros PJ, 2017a. Systemic
inflammation and resting state connectivity of the default mode network. Brain Behav. Immun
62, 162–170. [PubMed: 28126500]
Marsland AL, Walsh C, Lockwood K, John-Henderson NA, 2017b. The effects of acute psychological
stress on circulating and stimulated inflammatory markers: A systematic review and meta-
analysis. Brain Behav. Immun 64, 208–219. [PubMed: 28089638]
Martin JL, Martin-Sanchez E, 2012. Systematic review and meta-analysis of vagus nerve stimulation
in the treatment of depression: variable results based on study designs. Eur. Psychiatry 27, 147–
155. [PubMed: 22137776]
Marvel FA, Chen CC, Badr N, Gaykema RP, Goehler LE, 2004. Reversible inactivation of the dorsal
vagal complex blocks lipopolysaccharide-induced social withdrawal and c-Fos expression in
central autonomic nuclei. Brain Behav. Immun 18, 123–134. [PubMed: 14759590]
Author Manuscript

Mehta ND, Haroon E, Xu X, Woolwine BJ, Li Z, Felger JC, 2018. Inflammation negatively correlates
with amygdala-ventromedial prefrontal functional connectivity in association with anxiety in
patients with depression: Preliminary results. Brain Behav. Immun 73, 725–730. [PubMed:
30076980]
Meier T, Savitz J, Singh R, Teague TK, Bellgowan P, 2015a. Smaller Dentate Gyrus and CA
2 & 3 volumes are associated with kynurenine metabolites in collegiate football athletes. J.
Neurotrauma
Meier TB, Drevets WC, Wurfel BE, Ford BN, Morris HM, Victor TA, Bodurka J, Teague TK, Dantzer
R, Savitz J, 2015b. Relationship between neurotoxic kynurenine metabolites and reductions in
right medial prefrontal cortical thickness in major depressive disorder. Brain Behav. Immun
Milaneschi Y, Corsi AM, Penninx BW, Bandinelli S, Guralnik JM, Ferrucci L, 2009. Interleukin-1
Receptor Antagonist and Incident Depressive Symptoms Over 6 Years in Older Persons: The
InCHIANTI Study. Biol. Psychiatry 65, 973–978. [PubMed: 19111279]
Miller AH, Haroon E, Raison CL, Felger JC, 2013. Cytokine targets in the brain: impact on
Author Manuscript

neurotransmitters and neurocircuits. Depress. Anxiety 30, 297–306. [PubMed: 23468190]


Miller AH, Maletic V, Raison CL, 2009. Inflammation and its discontents: the role of cytokines in the
pathophysiology of major depression. Biol. Psychiatry 65, 732–741. [PubMed: 19150053]
Miller ES, Sakowicz A, Roy A, Yang A, Sullivan JT, Grobman WA, Wisner KL, 2019. Plasma and
cerebrospinal fluid inflammatory cytokines in perinatal depression. Am. J. Obstet. Gynecol 220,
271 e271–271 e210. [PubMed: 30557551]
Miller LG, Galpern WR, Dunlap K, Dinarello CA, Turner TJ, 1991. Interleukin-1 augments
gamma-aminobutyric acidA receptor function in brain. Mol. Pharmacol 39, 105–108. [PubMed:
1847488]

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 25

Miresco MJ, Kirmayer LJ, 2006. The persistence of mind-brain dualism in psychiatric reasoning about
clinical scenarios. Am. J. Psychiatry 163, 913–918. [PubMed: 16648335]
Author Manuscript

Mizoguchi K, Ishige A, Aburada M, Tabira T, 2003. Chronic stress attenuates glucocorticoid negative
feedback: involvement of the prefrontal cortex and hippocampus. Neuroscience 119, 887–897.
[PubMed: 12809708]
Mocking R, Harmsen I, Assies J, Koeter M, Ruhe H, Schene A, 2017. Meta-analysis and meta-
regression of omega-3 polyunsaturated fatty acid supplementation for major depressive disorder.
Transl. Psychiatry 6, e756.
Molendijk M, Spinhoven P, Polak M, Bus B, Penninx B, Elzinga B, 2014. Serum BDNF
concentrations as peripheral manifestations of depression: evidence from a systematic review and
meta-analyses on 179 associations (N= 9484). Mol. Psychiatry 19, 791. [PubMed: 23958957]
Morrison DJ, Preston T, 2016. Formation of short chain fatty acids by the gut microbiota and their
impact on human metabolism. Gut Microbes 7, 189–200. [PubMed: 26963409]
Morrison SF, Cao WH, 2000. Different adrenal sympathetic preganglionic neurons regulate
epinephrine and norepinephrine secretion. Am. J. Physiol. Regul. Integr. Comp. Physiol 279,
R1763–R1775. [PubMed: 11049860]
Author Manuscript

Nance DM, Sanders VM, 2007. Autonomic innervation and regulation of the immune system (1987–
2007). Brain Behav. Immun 21, 736–745. [PubMed: 17467231]
Naseribafrouei A, Hestad K, Avershina E, Sekelja M, Linlokken A, Wilson R, Rudi K, 2014.
Correlation between the human fecal microbiota and depression. Neurogastroenterology and
motility: the official journal of the. Eur. Gastrointest. Motil. Soc 26, 1155–1162.
Neupane SP, Lien L, Ueland T, Mollnes TE, Aukrust P, Bramness JG, 2015. Serum brain-derived
neurotrophic factor levels in relation to comorbid depression and cytokine levels in Nepalese men
with alcohol-use disorders. Alcohol 49, 471–478. [PubMed: 25873205]
Numakawa T, Richards M, Nakajima S, Adachi N, Furuta M, Odaka H, Kunugi H, 2014. The Role
of Brain-Derived Neurotrophic Factor in Comorbid Depression: Possible Linkage with Steroid
Hormones. Cytokines Nutrit. Front. Psychiatry 5.
O’connor J, Lawson M, Andre C, Moreau M, Lestage J, Castanon N, Kelley K, Dantzer R,
2009. Lipopolysaccharide-induced depressive-like behavior is mediated by indoleamine 2, 3-
dioxygenase activation in mice. Mol. Psychiatry 14, 511–522. [PubMed: 18195714]
Author Manuscript

Oglodek EA, Just MJ, Szromek AR, Araszkiewicz A, 2017. Assessing the serum concentration levels
of NT-4/5, GPX-1, TNF-alpha, and l-arginine as biomediators of depression severity in first
depressive episode patients with and without posttraumatic stress disorder. Pharmacol. Rep 69,
1049–1058. [PubMed: 28958613]
Ogyu K, Kubo K, Noda Y, Iwata Y, Tsugawa S, Omura Y, Wada M, Tarumi R, Plitman E, Moriguchi
S, Miyazaki T, Uchida H, Graff-Guerrero A, Mimura M, Nakajima S, 2018. Kynurenine pathway
in depression: A systematic review and meta-analysis. Neurosci. Biobehav. Rev 90, 16–25.
[PubMed: 29608993]
Oh DY, Talukdar S, Bae EJ, Imamura T, Morinaga H, Fan W, Li P, Lu WJ, Watkins SM, Olefsky
JM, 2010. GPR120 is an omega-3 fatty acid receptor mediating potent anti-inflammatory and
insulin-sensitizing effects. Cell 142, 687–698. [PubMed: 20813258]
Pace TWW, Miller AH, 2009. Cytokines and glucocorticoid receptor signaling. Relevance to major
depression. Ann. N. Y. Acad. Sci 1179, 86–105. [PubMed: 19906234]
Padgett DA, Glaser R, 2003. How stress influences the immune response. Trends Immunol. 24, 444–
448. [PubMed: 12909458]
Author Manuscript

Palta P, Samuel LJ, Miller ER 3rd, Szanton SL, 2014. Depression and oxidative stress: results from a
meta-analysis of observational studies. Psychosom. Med 76, 12–19. [PubMed: 24336428]
Pandey GN, Rizavi HS, Ren X, Fareed J, Hoppensteadt DA, Roberts RC, Conley RR, Dwivedi Y,
2012. Proinflammatory cytokines in the prefrontal cortex of teenage suicide victims. J. Psychiatr.
Res 46, 57–63. [PubMed: 21906753]
Papageorgiou IE, Lewen A, Galow LV, Cesetti T, Scheffel J, Regen T, Hanisch UK, Kann O, 2016.
TLR4-activated microglia require IFN-gamma to induce severe neuronal dysfunction and death
in situ. PNAS 113, 212–217. [PubMed: 26699475]

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 26

Pariante CM, Lightman SL, 2008. The HPA axis in major depression: classical theories and new
developments. Trends Neurosci. 31, 464–468. [PubMed: 18675469]
Author Manuscript

Pariante CM, Miller AH, 2001. Glucocorticoid receptors in major depression: relevance to
pathophysiology and treatment. Biol. Psychiatry 49, 391–404. [PubMed: 11274650]
Pariante CM, Pearce BD, Pisell TL, Sanchez CI, Po C, Su C, Miller AH, 1999. The
proinflammatory cytokine, interleukin-1α, reduces glucocorticoid receptor translocation and
function. Endocrinology 140, 4359–4366. [PubMed: 10465310]
Patas K, Penninx BW, Bus BA, Vogelzangs N, Molendijk ML, Elzinga BM, Bosker FJ, Voshaar RCO,
2014. Association between serum brain-derived neurotrophic factor and plasma interleukin-6 in
major depressive disorder with melancholic features. Brain Behav. Immun 36, 71–79. [PubMed:
24140302]
Pavlov VA, Tracey KJ, 2012. The vagus nerve and the inflammatory reflex—linking immunity and
metabolism. Nat. Rev. Endocrinol 8, 743. [PubMed: 23169440]
Philip NS, Carpenter LL, Tyrka AR, Price LH, 2010. Nicotinic acetylcholine receptors and depression:
a review of the preclinical and clinical literature. Psychopharmacology (Berl.) 212, 1–12.
[PubMed: 20614106]
Author Manuscript

Pongratz G, Straub RH, 2014. The sympathetic nervous response in inflammation. Arthritis Res. Ther
16, 504. [PubMed: 25789375]
Prather AA, Rabinovitz M, Pollock BG, Lotrich FE, 2009. Cytokine-induced depression during
IFN-alpha treatment: the role of IL-6 and sleep quality. Brain Behav. Immun 23, 1109–1116.
[PubMed: 19615438]
Predonzani A, Cali B, Agnellini AH, Molon B, 2015. Spotlights on immunological effects of reactive
nitrogen species: When inflammation says nitric oxide. World J Exp Med 5, 64–76.
Price JL, Drevets WC, 2010. Neurocircuitry of mood disorders. Neuropsychopharmacology 35, 192–
216. [PubMed: 19693001]
Raison CL, Borisov AS, Majer M, Drake DF, Pagnoni G, Woolwine BJ, Vogt GJ, Massung B, Miller
AH, 2009. Activation of central nervous system inflammatory pathways by interferon-alpha:
relationship to monoamines and depression. Biol. Psychiatry 65, 296–303. [PubMed: 18801471]
Raison CL, Capuron L, Miller AH, 2006. Cytokines sing the blues: inflammation and the pathogenesis
of depression. Trends Immunol. 27, 24–31. [PubMed: 16316783]
Author Manuscript

Raison CL, Dantzer R, Kelley KW, Lawson MA, Woolwine BJ, Vogt G, Spivey JR, Saito K, Miller
AH, 2010. CSF concentrations of brain tryptophan and kynurenines during immune stimulation
with IFN-alpha: relationship to CNS immune responses and depression. Mol. Psychiatry 15,
393–403. [PubMed: 19918244]
Raison CL, Miller AH, 2013. Do cytokines really sing the blues? Cerebrum: the Dana forum on brain
science. Dana Foundation
Ramirez K, Fornaguera-Trías J, Sheridan JF, 2016. Stress-induced microglia activation and monocyte
trafficking to the brain underlie the development of anxiety and depression, Inflammation-
Associated Depression: Evidence, Mechanisms and Implications. Springer, pp. 155–172.
Rapaport MH, Nierenberg AA, Schettler PJ, Kinkead B, Cardoos A, Walker R, Mischoulon D, 2016.
Inflammation as a predictive biomarker for response to omega-3 fatty acids in major depressive
disorder: a proof-of-concept study. Mol. Psychiatry 21, 71. [PubMed: 25802980]
Reagan LP, McEwen BS, 1997. Controversies surrounding glucocorticoid-mediated cell death in the
hippocampus. J. Chem. Neuroanat 13, 149–167. [PubMed: 9315966]
Redza-Dutordoir M, Averill-Bates DA, 2016. Activation of apoptosis signalling pathways by reactive
Author Manuscript

oxygen species. Biochim. Biophys. Acta 1863, 2977–2992. [PubMed: 27646922]


Rengasamy M, McClain L, Gandhi P, Segreti AM, Brent D, Peters D, Pan L, 2018. Associations
of plasma interleukin-6 with plasma and cerebrospinal fluid monoamine biosynthetic pathway
metabolites in treatment-resistant depression. Neurol. Psychiatry Brain Res 30, 39–46.
Réus GZ, Jansen K, Titus S, Carvalho AF, Gabbay V, Quevedo J, 2015. Kynurenine pathway
dysfunction in the pathophysiology and treatment of depression: evidences from animal and
human studies. J. Psychiatr. Res

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 27

Rey C, Nadjar A, Buaud B, Vaysse C, Aubert A, Pallet V, Laye S, Joffre C, 2016. Resolvin D1 and
E1 promote resolution of inflammation in microglial cells in vitro. Brain Behav. Immun 55,
Author Manuscript

249–259. [PubMed: 26718448]


Richards D, 2011. Prevalence and clinical course of depression: a review. Clin. Psychol. Rev 31,
1117–1125. [PubMed: 21820991]
Rohan Walker F, Nilsson M, Jones K, 2013. Acute and chronic stress-induced disturbances of
microglial plasticity, phenotype and function. Curr. Drug Targets 14, 1262–1276. [PubMed:
24020974]
Rooks MG, Garrett WS, 2016. Gut microbiota, metabolites and host immunity. Nat. Rev. Immunol 16,
341–352. [PubMed: 27231050]
Roque A, Ochoa-Zarzosa A, Torner L, 2016. Maternal separation activates microglial cells and induces
an inflammatory response in the hippocampus of male rat pups, independently of hypothalamic
and peripheral cytokine levels. Brain Behav. Immun 55, 39–48. [PubMed: 26431692]
Salagre E, Fernandes BS, Dodd S, Brownstein DJ, Berk M, 2016. Statins for the treatment of
depression: A meta-analysis of randomized, double-blind, placebo- controlled trials. J. Affect.
Disord 200, 235–242. [PubMed: 27148902]
Author Manuscript

Sapolsky RM, Romero LM, Munck AU, 2000. How do glucocorticoids influence stress responses?
Integrating permissive, suppressive, stimulatory, and preparative actions. Endocr. Rev 21, 55–89.
[PubMed: 10696570]
Sasayama D, Hattori K, Wakabayashi C, Teraishi T, Hori H, Ota M, Yoshida S, Arima K, Higuchi
T, Amano N, Kunugi H, 2013. Increased cerebrospinal fluid interleukin-6 levels in patients with
schizophrenia and those with major depressive disorder. J. Psychiatr. Res 47, 401–406. [PubMed:
23290488]
Satizabal CL, Zhu YC, Mazoyer B, Dufouil C, Tzourio C, 2012. Circulating IL-6 and CRP are
associated with MRI findings in the elderly: the 3C-Dijon Study. Neurology 78, 720–727.
[PubMed: 22357713]
Scanzano A, Cosentino M, 2015. Adrenergic regulation of innate immunity: a review. Front.
Pharmacol 6, 171. [PubMed: 26321956]
Schloesser RJ, Manji HK, Martinowich K, 2009. Suppression of adult neurogenesis leads to an
increased HPA axis response. Neuroreport 20, 553. [PubMed: 19322118]
Author Manuscript

Schmaal L, Veltman DJ, van Erp TG, Samann PG, Frodl T, Jahanshad N, Loehrer E, Tiemeier H,
Hofman A, Niessen WJ, Vernooij MW, Ikram MA, Wittfeld K, Grabe HJ, Block A, Hegenscheid
K, Volzke H, Hoehn D, Czisch M, Lagopoulos J, Hatton SN, Hickie IB, Goya-Maldonado R,
Kramer B, Gruber O, Couvy-Duchesne B, Renteria ME, Strike LT, Mills NT, de Zubicaray
GI, McMahon KL, Medland SE, Martin NG, Gillespie NA, Wright MJ, Hall GB, MacQueen
GM, Frey EM, Carballedo A, van Velzen LS, van Tol MJ, van der Wee NJ, Veer IM, Walter
H, Schnell K, Schramm E, Normann C, Schoepf D, Konrad C, Zurowski B, Nickson T,
McIntosh AM, Papmeyer M, Whalley HC, Sussmann JE, Godlewska BR, Cowen PJ, Fischer
FH, Rose M, Penninx BW, Thompson PM, Hibar DP, 2016. Subcortical brain alterations in major
depressive disorder: findings from the ENIGMA Major Depressive Disorder working group. Mol.
Psychiatry 21, 806–812. [PubMed: 26122586]
Schrader LA, Birnbaum SG, Nadin BM, Ren Y, Bui D, Anderson AE, Sweatt JD, 2006. ERK/MAPK
regulates the Kv4.2 potassium channel by direct phosphorylation of the pore-forming subunit.
Am. J. Physiol. Cell Physiol 290, C852–C861. [PubMed: 16251476]
Shen Y, Connor TJ, Nolan Y, Kelly JP, Leonard BE, 1999. Differential effect of chronic antidepressant
Author Manuscript

treatments on lipopolysaccharide-induced depressive-like behavioural symptoms in the rat. Life


Sci. 65, 1773–1786. [PubMed: 10576557]
Sherwin E, Rea K, Dinan TG, Cryan JF, 2016. A gut (microbiome) feeling about the brain. Curr. Opin.
Gastroenterol. 32, 96–102.
Shi P, Raizada MK, Sumners C, 2010. Brain cytokines as neuromodulators in cardiovascular control.
Clin. Exp. Pharmacol. Physiol 37, e52–e57. [PubMed: 19566837]
Smith KJ, Au B, Ollis L, Schmitz N, 2018. The association between C-reactive protein, Interleukin-6
and depression among older adults in the community: a systematic review and meta-analysis.
Exp. Gerontol 102, 109–132. [PubMed: 29237576]

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 28

Sonino N, Fava G, 1998. Psychosomatic aspects of Cushing’s disease. Psychother. Psychosom 67,
140–146. [PubMed: 9667061]
Author Manuscript

Sorrells SF, Sapolsky RM, 2007. An inflammatory review of glucocorticoid actions in the CNS. Brain
Behav. Immun 21, 259–272. [PubMed: 17194565]
Steimer T, 2002. The biology of fear-and anxiety-related behaviors. Dialogues Clin Neurosci 4, 231–
249. [PubMed: 22033741]
Stellwagen D, Beattie EC, Seo JY, Malenka RC, 2005. Differential regulation of AMPA receptor and
GABA receptor trafficking by tumor necrosis factor-alpha. J. Neurosci 25, 3219–3228. [PubMed:
15788779]
Stetler C, Miller GE, 2011. Depression and hypothalamic-pituitary-adrenal activation: a quantitative
summary of four decades of research. Psychosom. Med 73, 114–126. [PubMed: 21257974]
Stewart JC, Rand KL, Muldoon MF, Kamarck TW, 2009. A prospective evaluation of the directionality
of the depression–inflammation relationship. Brain Behav. Immun 23, 936–944. [PubMed:
19416750]
Swardfager W, Herrmann N, Mazereeuw G, Goldberger K, Harimoto T, Lanctot KL, 2013. Zinc in
depression: a meta-analysis. Biol. Psychiatry 74, 872–878. [PubMed: 23806573]
Author Manuscript

Swardfager W, Rosenblat JD, Benlamri M, McIntyre RS, 2016. Mapping inflammation onto
mood: Inflammatory mediators of anhedonia. Neurosci. Biobehav. Rev 64, 148–166. [PubMed:
26915929]
Takeuchi H, Jin S, Wang J, Zhang G, Kawanokuchi J, Kuno R, Sonobe Y, Mizuno T, Suzumura
A, 2006. Tumor necrosis factor-alpha induces neurotoxicity via glutamate release from
hemichannels of activated microglia in an autocrine manner. J. Biol. Chem 281, 21362–21368.
[PubMed: 16720574]
Tanaka J, Fujita H, Matsuda S, Toku K, Sakanaka M, Maeda N, 1997. Glucocorticoid-and
mineralocorticoid receptors in microglial cells: the two receptors mediate differential effects
of corticosteroids. Glia 20, 23–37. [PubMed: 9145302]
Taylor WD, Aizenstein HJ, Alexopoulos GS, 2013. The vascular depression hypothesis: mechanisms
linking vascular disease with depression. Mol. Psychiatry 18, 963–974. [PubMed: 23439482]
Thakare VN, Aswar MK, Kulkarni YP, Patil RR, Patel BM, 2017. Silymarin ameliorates
experimentally induced depressive like behavior in rats: Involvement of hippocampal BDNF
Author Manuscript

signaling, inflammatory cytokines and oxidative stress response. Physiol. Behav 179, 401–410.
[PubMed: 28711395]
Tian R, Hou G, Li D, Yuan T-F, 2014. A possible change process of inflammatory cytokines in the
prolonged chronic stress and its ultimate implications for health. Scientific World J. 2014.
Tonelli LH, Stiller J, Rujescu D, Giegling I, Schneider B, Maurer K, Schnabel A, Moller HJ, Chen
HH, Postolache TT, 2008. Elevated cytokine expression in the orbitofrontal cortex of victims of
suicide. Acta Psychiatr. Scand 117, 198–206. [PubMed: 18081924]
Tracey KJ, 2002. The inflammatory reflex. Nature 420, 853–859. [PubMed: 12490958]
Treadway MT, Zald DH, 2011. Reconsidering anhedonia in depression: lessons from translational
neuroscience. Neurosci. Biobehav. Rev 35, 537–555. [PubMed: 20603146]
Tyagi S, Gupta P, Saini AS, Kaushal C, Sharma S, 2011. The peroxisome proliferator-activated
receptor: A family of nuclear receptors role in various diseases. J. Adv. Pharm. Technol. Res 2,
236. [PubMed: 22247890]
Ulrich-Lai YM, Herman JP, 2009. Neural regulation of endocrine and autonomic stress responses. Nat.
Rev. Neurosci 10, 397–409. [PubMed: 19469025]
Author Manuscript

Valkanova V, Ebmeier KP, Allan CL, 2013. CRP, IL-6 and depression: a systematic review and
meta-analysis of longitudinal studies. J. Affect. Disord 150, 736–744. [PubMed: 23870425]
Vallieres L, Rivest S, 1999. Interleukin-6 is a needed proinflammatory cytokine in thè prolonged
neural activity and transcriptional activation of corticotropin-releasing factor during endotoxemia.
Endocrinology 140, 3890–3903. [PubMed: 10465257]
Vezzani A, Viviani B, 2015. Neuromodulatory properties of inflammatory cytokines and their impact
on neuronal excitability. Neuropharmacology 96, 70–82. [PubMed: 25445483]

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.
Rengasamy et al. Page 29

Vian J, Pereira C, Chavarria V, Köhler C, Stubbs B, Quevedo J, Kim S-W, Carvalho AF, Berk M,
Fernandes BS, 2017. The renin–angiotensin system: a possible new target for depression. BMC
Author Manuscript

Med. 15, 144. [PubMed: 28760142]


Viviani B, Bartesaghi S, Gardoni F, Vezzani A, Behrens M, Bartfai T, Binaglia M, Corsini E, Di
Luca M, Galli C, 2003. Interleukin-1β enhances NMDA receptor-mediated intracellular calcium
increase through activation of the Src family of kinases. J. Neurosci 23, 8692–8700. [PubMed:
14507968]
Wichers M, Koek G, Robaeys G, Verkerk R, Scharpe S, Maes M, 2005. IDO and interferon-α-induced
depressive symptoms: a shift in hypothesis from tryptophan depletion to neurotoxicity. Mol.
Psychiatry 10, 538–544. [PubMed: 15494706]
Wichers MC, Kenis G, Koek GH, Robaeys G, Nicolson NA, Maes M, 2007. Interferon-α-induced
depressive symptoms are related to changes in the cytokine network but not to cortisol. J.
Psychosom. Res 62, 207–214. [PubMed: 17270579]
Wise T, Radua J, Via E, Cardoner N, Abe O, Adams TM, Amico F, Cheng Y, Cole JH, de Azevedo
Marques Perico C, Dickstein DP, Farrow TFD, Frodl T, Wagner G, Gotlib IH, Gruber O, Ham
BJ, Job DE, Kempton MJ, Kim MJ, Koolschijn P, Malhi GS, Mataix-Cols D, McIntosh AM,
Author Manuscript

Nugent AC, O’Brien JT, Pezzoli S, Phillips ML, Sachdev PS, Salvadore G, Selvaraj S, Stanfield
AC, Thomas AJ, van Tol MJ, van der Wee NJA, Veltman DJ, Young AH, Fu CH, Cleare
AJ, Arnone D, 2017. Common and distinct patterns of grey-matter volume alteration in major
depression and bipolar disorder: evidence from voxel-based meta-analysis. Mol. Psychiatry 22,
1455–1463. [PubMed: 27217146]
Yao W, Zhang JC, Ishima T, Dong C, Yang C, Ren Q, Ma M, Han M, Wu J, Suganuma H,
Ushida Y, Yamamoto M, Hashimoto K, 2016. Role of Keap1-Nrf2 signaling in depression and
dietary intake of glucoraphanin confers stress resilience in mice. Sci. Rep 6, 30659. [PubMed:
27470577]
Young JJ, Bruno D, Pomara N, 2014. A review of the relationship between proinflammatory cytokines
and major depressive disorder. J. Affect. Disord 169, 15–20. [PubMed: 25128861]
Yuan H, Silberstein SD, 2016. Vagus Nerve and Vagus Nerve Stimulation, a Comprehensive Review:
Part I. Headache 56, 71–78. [PubMed: 26364692]
Yuan N, Chen Y, Xia Y, Dai J, Liu C, 2019. Inflammation-related biomarkers in major psychiatric
disorders: a cross-disorder assessment of reproducibility and specificity in 43 meta-analyses.
Author Manuscript

Transl. Psychiatry 9, 1–13. [PubMed: 30664621]


Zalcman S, Green-Johnson JM, Murray L, Nance DM, Dyck D, Anisman H, Greenberg AH, 1994.
Cytokine-specific central monoamine alterations induced by interleukin-1, −2 and −6. Brain Res.
643, 40–49. [PubMed: 7518332]
Zanos P, Gould TD, 2018. Mechanisms of ketamine action as an antidepressant. Mol. Psychiatry
Zhang J, Fan Y, Raza MU, Zhan Y, Du XD, Patel PD, Zhu MY, 2017. The regulation of corticosteroid
receptors in response to chronic social defeat. Neurochem. Int 108, 397–409. [PubMed:
28577990]
Zhu CB, Blakely RD, Hewlett WA, 2006. The proinflammatory cytokines interleukin-1beta and tumor
necrosis factor-alpha activate serotonin transporters. Neuropsychopharmacology 31, 2121–2131.
[PubMed: 16452991]
Zorn JV, Schür RR, Boks MP, Kahn RS, Joels M, Vinkers CH, 2017. Cortisol stress reactivity across
psychiatric disorders: A systematic review and meta-analysis. Psychoneuroendocrinology 77,
25–36. [PubMed: 28012291]
Author Manuscript

J Affect Disord Rep. Author manuscript; available in PMC 2022 December 01.

You might also like