ArticlePDF AvailableLiterature Review

Gut-microbiota-brain axis in depression: The role of neuroinflammation

Authors:

Abstract

Major depressive disorder (MDD) is a psychiatric condition that affects a large number of people in the world and the treatment existents do not work for all individuals affected. Thus, it is believed that other systems or pathways which regulate brain networks involved in mood regulation and cognition are associated with MDD pathogenesis. Studies in humans and animal models have been shown that in MDD there are increased levels of inflammatory mediators, including cytokines and chemokines in both periphery and central nervous system (CNS). In addition, microglial activation appears to be a key event that triggers changes in signaling cascades and gene expression that would be determinant for the onset of depressive symptoms. Recent researches also points out that changes in the gut microbiota would lead to a systemic inflammation that in different ways would reach the CNS modulating inflammatory pathways and especially the microglia, which could influence responses to treatments. Moreover, pre- and probiotics have shown antidepressant responses and anti-inflammatory effects. This review will focus on studies that show the relationship of inflammation with the gut-microbiota brain axis and its relation with MDD.
Eur J Neurosci. 2019;00:1–14. wileyonlinelibrary.com/journal/ejn
|
1
© 2019 Federation of European Neuroscience Societies
and John Wiley & Sons Ltd
Received: 14 August 2019
|
Revised: 18 November 2019
|
Accepted: 20 November 2019
DOI: 10.1111/ejn.14631
SPECIAL ISSUE REVIEW
Gut microbiota–brain axis in depression: The role of
neuroinflammation
Anelise S.Carlessi1
|
Laura A.Borba1
|
Alexandra I.Zugno1
|
JoãoQuevedo1,2,3,4
|
Gislaine Z.Réus1
Edited by Dr. Michel Barrot.
The peer review history for this article is available at https ://publo ns.com/publo n/10.1111/ejn.14631
Abbreviations: ACTH, adrenocorticotropic hormone; ANS, autonomic nervous system; ATP, adenosine triphosphate; BBB, blood–brain barrier; CNS, central
nervous system; CSF, cerebrospinal fluid; DAMPs, damage-associated molecular patterns; ENS, enteric nervous system; GABA, gamma-aminobutyric acid; GF,
germ-free; GOS, galactooligosaccharide; HAA, 3-hydroxyanthranilic acid; HPA, hypothalamic–pituitary–adrenal; IBS, irritable bowel syndrome; IDO, indoleam-
ine-2, 3-dioxygenase; IFN-γ, interferon gamma; IgA, immunoglobulin A; IL-10, interleukin-10; IL-5, interleukin-5; IL-6, interleukin-6; IL-7, interleukin-7; INF-γ,
Interferon gamma; KATs, kynurenine aminotransferases; KMO, quinurenine-3-mono-oxygenase; KP, kynurenine pathway; KYNA, kynurenic acid; LAC,
glycoprotein lactoferrin; LH, learned helplessness; LPS, lipopolysaccharide; MAOIs, monoamine oxidase inhibitors; MDD, major depressive disorder; MFGM, milk
fat globule membrane; MR16-1, anti-mouse IL-6 receptor antibody; NAD, nicotinamide adenine dinucleotide; NF-Κb, nuclear factor kappa B; NLRP3, receptor
family pyrin domain-containing 3; NMDA, antagonist of N-methyl-D-aspartate; NSAD, nonsteroidal anti-inflammatory drugs; OHK, 3-hydroxyquinurenine;
PAMPs, pathogen-associated molecular patterns; PDX, polydextrose; QUIN, quinolinic acid; SCFAs, short-chain fatty acids; SSRIs, selective serotonin reuptake
inhibitors; TCAs, tricyclic agents; TH1, T helper 1 effector cells; TLRs, Toll-like Receptors; TNF-α, tumor necrosis factor- α; TODO, tryptophan-2,3-dioxygenase.
1Translational Psychiatry Laboratory,
Graduate Program in Health Sciences,
University of Southern Santa Catarina
(UNESC), Criciúma, Brazil
2Translational Psychiatry Program,
Department of Psychiatry and Behavioral
Sciences, McGovern Medical School,
University of Texas Health Science Center
at Houston (UTHealth), Houston, TX, USA
3Center of Excellence on Mood Disorders,
Department of Psychiatry and Behavioral
Sciences, McGovern Medical School, The
University of Texas Health Science Center
at Houston (UTHealth), Houston, TX, USA
4Neuroscience Graduate Program, The
University of Texas Graduate School of
Biomedical Sciences at Houston, Houston,
TX, USA
Correspondence
Gislaine Z. Réus, Translational Psychiatry
Laboratory, University of Southern Santa
Catarina, Criciuma, SC, 88806-000, Brazil.
Email: gislainereus@unesc.net
Funding information
Texas Health Science Center; Pat
Rutherford Jr. Chair in Psychiatry; John
S. Dunn Foundation; Anne and Don Fizer
Foundation Endowment for Depression
Research; CNPq; FAPESC; Instituto
Cérebro e Mente; UNESC
Abstract
Major depressive disorder (MDD) is a psychiatric condition that affects a large num-
ber of people in the world, and the treatment existents do not work for all individuals
affected. Thus, it is believed that other systems or pathways which regulate brain net-
works involved in mood regulation and cognition are associated with MDD patho-
genesis. Studies in humans and animal models have been shown that in MDD there
are increased levels of inflammatory mediators, including cytokines and chemokines
in both periphery and central nervous system (CNS). In addition, microglial activa-
tion appears to be a key event that triggers changes in signaling cascades and gene
expression that would be determinant for the onset of depressive symptoms. Recent
researches also point out that changes in the gut microbiota would lead to a systemic
inflammation that in different ways would reach the CNS modulating inflammatory
pathways and especially the microglia, which could influence responses to treat-
ments. Moreover, pre- and probiotics have shown antidepressant responses and anti-
inflammatory effects. This review will focus on studies that show the relationship of
inflammation with the gut microbiota–brain axis and its relation with MDD.
KEYWORDS
gut microbiota–brain axis, inflammasome, kynurenine pathway, major depressive disorder,
neuroinflammation
2
|
CARLESSI ET AL.
1
|
INTRODUCTION
1.1
|
Depression
Mental disorders are one of the main causes of disability in
developed countries, overcoming illnesses such as coronary
diseases and cancer (Reeves et al., ; Whiteford et al., 2013).
Major depressive disorder (MDD) is the main psychiatric
disorder condition, and it is the leading cause of disability,
reaching the whole world (Ferrari et al., 2013). It is estimated
that MDD reaches 4.4% of the world population, correspond-
ing to 322 million people living with this disorder (World
Health Organization, 2017).
The symptoms of MDD include depressed mood, anhedo-
nia, irritability, difficulty concentrating, changes in appetite
and sleep, and others (Nestler et al., 2002). In addition, MDD
is associated with psychosocial and functional impairment,
cognitive deficits, increased risk of suicidal behavior, and in-
creased mortality (Amidfar, Réus, Quevedo, & Kim, 2018),
reaching close to 800,000 deaths per year due to suicides
(World Health Organization, 2017).
The pathophysiology of MDD has been not still completely
elucidated. The hypothesis of monoaminergic deficiency,
which a decrease in the levels of serotonin, noradrenaline,
and dopamine in the synaptic cleft is the most used theory
to explain depressive symptoms. Moreover, antidepressant
drugs used for MDD treatment act increase monoamine lev-
els (Berton & Nestler, 2006; Schildkraut, 1995).
The first generation of drugs used to treat MDD, known
as classical antidepressants, were monoamine oxidase in-
hibitors (MAOIs) and tricyclic agents (TCAs), following for
selective serotonin reuptake inhibitors (SSRIs; Réus et al.,
2016). Currently, MAOIs and TCAs are not the first options
in the treatment of MDD due to considerable side effects
(Elhwuegi, 2004), such as hypertensive crisis, tachycardia,
anxiety, orthostatic hypotension, myoclonus, and convul-
sion, among other effects (Aguiar et al., 2011). In fact, SSRIs
are the more prescribed antidepressant drugs due it milder
side effects, greater tolerance, and adherence to treatment
by individuals with MDD (Millan, 2006). However, classic
antidepressants have some limitations, such as the delay to
therapeutic response (around three-six weeks) and low remis-
sion rates (around 30%; Amidfar, Réus, Quevedo, Kim, &
Arbabi, 2017; Krishnan & Nestler, 2008; Machado-Vieira et
al., 2009). Thus, it is crucial to study new therapeutic strate-
gies, as well as new pathways to understand the pathophysiol-
ogy of MDD. In fact, other systems have been associated with
MDD, including genes involved in biological rhythm regula-
tion, and immune system (Delpech et al., 2016; Johnson &
Kaffman, 2018; Ketchesin, Becker-Krail, & McClung, 2018;
Réus et al., 2017).
There is evidence that inflammatory processes, including
proinflammatory cytokines levels and microglial activation,
could influence behavior and emotions (Delpech et al., 2016;
Johnson & Kaffman, 2018; Réus et al., 2017). Studies have
shown that proinflammatory cytokines from the periphery
can lead to a microglial activation which can induce neuroin-
flammation (Nakamura, Okada, Toyama, & Okano, 2005;
Zhang, Hu, Qian, O'Callaghan, & Hong, 2010). In fact, rats
exposed to maternal deprivation early in life exhibited de-
pressive-like behavior in adult life and also an increase in the
proinflammatory cytokines and microglial activation (Réus
et al., 2017, 2019) which supports the idea that neuroinflam-
mation is also one of the pathways related to the development
of MDD.
One of the new pathways that have been studied to un-
derstand the pathophysiology of MDD is the gut–brain axis,
which communicate in complex and bidirectional ways. Some
studies show that individuals with MDD have altered gut
microbiota compared to healthy subjects (Jiang et al., 2015;
Kelly et al., 2016; Zheng et al., 2016). Emerging evidence
has suggested that the gut microbiota has some involvement
with inflammation, brain development, and behavior (Bailey
& Coe, 1999; Buffington et al., 2016; Diaz Heijtz et al., 2011;
Evrensel & Ceylan, 2015; Hsiao et al., 2013; Scott, Pound,
Patrick, Eberhard, & Crookshank, 2017). There are several
factors that can cause a disturbance in the balance of the gut
microbiota, stress is one of them, and it stimulates inflam-
mation-to-brain mechanisms and leads to microglia activa-
tion (Maes, 2008). Mice exposed to chronic restraint stress
had depressive-like behavior and an increase in the neuroin-
flammatory markers; on the other hand, prolonged intake
(21days) of the probiotic Bifidobacterium adolescentis was
able to reverse these changes, suggesting that the antidepres-
sant effects of B.adolescentis are related to reducing inflam-
matory cytokines and rebalancing the gut microbiota (Guo et
al., 2019).
2
|
NEUROINFLAMMATION IN
MDD
Advances in neuroscience have linked chemokines to neu-
robiological processes important to psychiatric disorders,
such as synaptic transmission, plasticity, neurogenesis, and
communication (Stuart & Baune, 2014). In this sense, clini-
cal (Enache, Pariante, & Mondelli, 2019) and experimental
studies (Giridharan et al., 2019) have been related stress
and MDD to an increase in the immune system activity.
Neuroinflammation caused by the immune system activation
affects the central nervous system (CNS) through cytokines
releasing causing a dysregulation in brain activities and emo-
tions. Recent studies support the idea that cytokines affect
the brain signal patterns involved in the psychopathology of
MDD and in the mechanisms of antidepressant drugs. These
observations suggest that neuroinflammation and cytokines
|
3
CARLESSI ET AL.
might cause and/or maintain depressive symptoms (Jeon &
Kim, 2016). Trials studies with classical anti-inflammatory
drugs such as celecoxib (Abbasi, Hosseini, Modabbernia,
Ashrafi, & Akhondzadeh, 2012) and nonsteroidal anti-in-
flammatory drugs (NSAD; Köhler, Krogh, Mors, & Benros,
2016) demonstrated an improvement of depressive symp-
toms with anti-inflammatory plus antidepressant treatment
compared to antidepressants plus placebo. In this sense, other
studies have been testing new anti-inflammatory therapies in
order to modulate the microglial activation. He et al. (2019)
tested paricalcitol, a vitamin D2 analogue that has been dem-
onstrated to have anti-inflammatory effects by modulating
nuclear factor kappa B (NF-κB) signaling. In this study, they
concluded that paricalcitol could alleviate the depressive-
like behavior induced by systemic lipopolysaccharide (LPS)
injection in mice, associating this to a potential anti-inflam-
matory effect. LPS is a component of the cell membrane of
gram-negative bacteria and has been used to induce depres-
sive behavior in animal models. Stress provoked by LPS
causes activation of the immune system and increase in the
expression of indoleamine-2, 3-dioxygenase (IDO) leading
to depressive-like behavior (Fu et al., 2010).
The imbalance produced by neuroinflammation can also
deregulate the hypothalamic–pituitary–adrenal (HPA) axis.
Some neurotransmitters, such as acetylcholine, dopamine,
noradrenaline, and serotonin, regulate peripheral cytokines
through cortisol levels, promoting or secreting corticotro-
pin-releasing hormone (CRH) in the hypothalamus, and adre-
nocorticotropic hormone (ACTH) in the pituitary (Calogero,
Gallucci, Chrousos, & Gold, 1988). In the normal state, pe-
ripheral cytokines are hydrophilic and have large molecular
weights; then, they are unable to cross through the blood–
brain barrier (BBB). However, they can cross through the
BBB in pathological states, including in the MDD situation.
In fact, increased BBB permeability is reported in depression
(Jeon & Kim, 2016).
Liu et al. (2019) describe in their review that important re-
gions involved in MDD are also correlated with higher levels
of proinflammatory cytokines. Experimental evidences have
been linked changes in the chemokine network to depressive
behavior. In addition, studies have been shown a relationship
between depression and microglial activation and an increase
in the proinflammatory cytokines, including interleukin-5
(IL-5), IL-6, IL-7, IL-10, tumor necrosis factor-α (TNF-α),
and interferon gamma (INF-γ), in the brain of rats subjected
to maternal deprivation (Giridharan et al., 2019). The au-
thors support the hypothesis that neuroinflammation and
microglial activation could be involved with changes in the
brain-resident cells, including excitotoxicity in neurons and
astrocytes atrophy following early life stress, which would be
associated with the development of depressive behavior.
Clinical studies have been also reported an association of
neuroinflammation and MDD. Enache et al. (2019) showed,
in a meta-analyses review, that an increase in the IL-6 and
TNF-α levels in cerebrospinal fluid (CSF) and brain paren-
chyma could be involved to microglial activation and reduc-
tion of astrocytes and oligodendrocytes. Müller et al. (2019)
examined an association of MDD with inflammation due to
a result of childhood maltreatment and threatening experi-
ences in the past year. They found significantly higher levels
of IL-6 and IL-10 in children with these adversities in life.
Moreover, it increased cytokine levels were related to the de-
velopment of MDD in the adulthood. In this context, IL-6
in higher concentration on serum of MDD patients who was
abused in their childhood was also detected (Munjiza et al.,
2018). Interestingly findings reported that female is dispro-
portionally affected by stress-primed inflammation. It could
be related to why women suffer more than men from mood
disorders (Bekhbat et al., 2019).
Considering these findings, the use of anti-inflammatory
agents in MDD has been investigated with growing interest.
In this sense, a recent meta-analysis concluded that anti-in-
flammatory agents improved antidepressant treatment effects.
However, the study calls the attention to future interventions
more personalized including measures of inflammation and
the somatic comorbidity profile in the overall assessment and
evaluation of the depressed patient (Köhler-Forsberg et al.,
2019).
3
|
GUT MICROBIOTA
The gut microbiota has been shown to be one of the most im-
portant complex and bidirectional pathways between gut and
brain (Rhee, Pothoulakis, & Mayer, 2009). Other pathways
previously established include the autonomic nervous sys-
tem (ANS), the enteric nervous system (ENS), the neuroen-
docrine system, and the immune system (Foster & McVey
Neufeld, 2013). These pathways interact to form a complex
communication between the brain and gut. The brain influ-
ences the motor, sensory, and secretory modalities of the gas-
trointestinal tract; on the other hand, the gut influences brain
function, especially in areas of the brain involved to stress
regulation (Dinan & Cryan, 2012; Hannan, 2016). In fact, it
was demonstrated that germ-free (GF) mice, with microbiota
deficient from birth, had different parts of amygdala and hip-
pocampus with morphology impairment (Luczynski et al.,
2016). The function of insular brain area also appears to be
associated with microbiota diversity (Curtis et al., 2019).
The bidirectional communication between gut and brain
may be substantiated on the comorbidity between gastrointes-
tinal and neurodegenerative diseases, such as Parkinson, and
mood disorders, including MDD (Kennedy, Cryan, Dinan,
& Clarke, 2014; Smith & Parr-Brownlie, 2019; Whitehead,
Palsson, & Jones, 2002); for example, a significant number
of patients with irritable bowel syndrome (IBS) have MDD
4
|
CARLESSI ET AL.
and/or anxiety (Spiller & Garsed, 2009). Furthermore, medi-
cations used to relieve the symptoms of patients with IBS and
eating disorders include are low-dose of antidepressants such
as tricyclic antidepressants (TCAs) or selective serotonin re-
uptake inhibitors (SSRIs; Ruepert et al., 2011).
Gut bacteria are essential in regulating important as-
pects of host health, such as brain development and function
(Diaz Heijtz et al., 2011; Hsiao et al., 2013). The microbi-
ota is influenced by different external stimuli, such as diet,
use of antibiotics, stress, and infections (Bercik, Park, et
al., 2011a). These factors can cause an imbalance between
pathogenic and beneficial bacteria (Forsythe, Sudo, Dinan,
Taylor, & Bienenstock, 2010), stimulating the process called
dysbiosis. Dysbiosis alters the permeability of the gut barrier,
and then, bacteria and its metabolic products could cross to
the periphery and activate immune response due release of
proinflammatory cytokines (Kiliaan et al., 1998). Dysbiosis
can increase inflammatory cytokines and bacterial metabo-
lites that can alter the gut and BBB permeability inducing
neuroinflammation (Roy & Banerjee, 2019). A study com-
pared oral and intraperitoneal administration of antibiotics in
adult mice and found that only animals treated by oral via had
transient changes in the composition of the gut microbiota
(Bercik, Denou, et al., 2011b).
Stress can dysregulate the gut microbiota, stimulating
immunological and brain mechanisms, including microglial
activation (Maes, 2008). Microglial cells are involved in the
release of proinflammatory cytokines in the brain in stress-
ful situations and seem to be altered in MDD (Réus, Fries,
et al., 2015a; Walker, Nilsson, & Jones, 2013). However, a
healthy microbiota can regulate these stress responses by
using the synthesis of hormones and neurotransmitters essen-
tial to minimize the effects of stress on the body (Asano et
al., 2012). Desbonnet et al. (2010) identified in their study
that the stress induced by early maternal deprivation in ro-
dents leads to depressive-like behavior and immune and
monoaminergic systems alterations (Desbonnet et al., 2010).
However, these changes were attenuated with the treatment
of Bifidobacterium infantis probiotic, which is a gram-posi-
tive anaerobic bacterium, existent in the gastrointestinal tract,
and it is one of the main genera of Actinobacteria that form
the colon microbiota in mammals (Duranti et al., 2016). This
study suggests that the microbiota by direct or indirect mech-
anisms may play an important role in CNS regulation and
psychiatric disorders.
In the context of early life stress, other studies also have
been shown that stress or changes in microbiota in early
life could lead to changes for all life. For example, dietary
interventions (milk fat globule membrane [MFGM] and
a polydextrose/galactooligosaccharide prebiotic blend) in
maternally separated rats improved HPA axis, changes, and
behavior impairment, and influenced abundance at family
and genus level as well as influencing beta-diversity levels
of microbiota (O'Mahony et al., 2019). Moreover, early life
supplementation of a blend of two prebiotics, galactooligo-
saccharide (GOS) and polydextrose (PDX), and the glyco-
protein lactoferrin (LAC) attenuated stress-induced learned
helplessness (Mika et al., 2017). Also, GOS, PDX, and LAC
diet was able to attenuate stress-evoked decreases in mRNA
for the 5-HT1A autoreceptor in the dorsal raphe nucleus and
increased basal BDNF mRNA in the prefrontal cortex (Mika
et al., 2017).
The gut microbiota plays important functions and modu-
lations that have been investigated and discussed as a potent
neuropharmacological target for MDD and other psychiatric
disorders conditions. Probiotics are described as living mi-
croorganisms that, when ingested in adequate amounts, bring
benefits to host health (Hill et al., 2014). The main mecha-
nisms by which probiotics exert their functions are through
modulation of the immune system, production of antimicro-
bial substances, competitive exclusion of pathogenic micro-
organisms, increase of the epithelial barrier, and gut mucosal
adhesion (Bermudez-Brito, Plaza-Díaz, Muñoz-Quezada,
Gómez-Llorente, & Gil, 2012). In addition, probiotics could
modulate opioid and cannabinoid receptors in gut epithelial
cells (Sanders, 2011). Opioids also have been associated with
neuroinflammatory processes induced by glial cells, such as
astrocytes and microglia (Hofford, Russo, and Kiraly (2018).
Probiotics also act on calcium-dependent potassium
channels in gut sensory neurons (Rousseaux et al., 2007).
Moreover, probiotics play a key role in the balance of the
gut flora, restoring the composition of the microbiota to a
more favorable state (Choi, Lee, & Paik, 2015). Some studies
also suggest that an improvement in the symptoms associated
to neurological and psychiatric disorders, as well as improve
the metabolic state, inflammatory biomarkers and oxidative
stress, through the probiotic effects on CNS circuits are me-
diated by the bowel-microbiota-brain axis (Messaoudi et al.,
2011; Réus, Fries, et al., 2015a).
Evidence has been shown that the gut microbiota can
affect the brain (Smith, 2015). Some bacteria, such as
Lactobacillus and Bifidobacterium, secrete gamma-amino-
butyric acid (GABA), an inhibitory neurotransmitter in the
brain, which regulates physiological and psychological pro-
cesses, and it is involved in the pathophysiology of anxiety
and MDD (Schousboe & Waagepetersen, 2007). Studies
have been shown that B.infantis can increase the availabil-
ity of tryptophan and thus the concentration of serotonin in
the brain (Desbonnet, Garrett, Clarke, Bienenstock, & Dinan,
2008; Desbonnet et al., 2010). Streptococcus, Escherichia,
and Enterococcus also produce serotonin, whereas the bac-
teria Escherichia, Bacillus, and Saccharomyces produce nor-
epinephrine, and Bacillus and Serratia have the potential to
produce dopamine (Holzer & Farzi, 2014; Özogul, 2011).
The microbiota and probiotics also act through neuro-
active bacterial metabolites called short-chain fatty acids
|
5
CARLESSI ET AL.
(SCFAs), such as acetate, butyrate, and propionate (Overduin,
Schoterman, Calame, Schonewille, & Bruggencate, 2013).
SCFAs are produced by the gut microbiota through the fer-
mentation of complex polysaccharides and have been shown to
have immunomodulatory effects (Macfarlane & Macfarlane,
2003). For example, Bifidobacterium can produce SCFAs by
lowering the pH of the intestine, forming biological barriers
and secreting antimicrobial compounds to attenuate patho-
genic bacteria (Liao et al., 2016). Another important aspect is
the strain specificity. In fact, a study demonstrated that mice
with anxiety and inflammation induced by parasite had an
improvement in anxious behavior after the treatment with
Bifidobacterium longum NC3001, but not with Lactobacillus
rhamnosus NCC4007 (Bercik et al., 2010).
Different probiotics have been investigated for neurolog-
ical and psychiatric disorders; however, Lactobacillus and
Bifidobacterium genera have been shown to be more effective
(Kim, Yun, On, & Choi, 2018). Yang et al. (2017) demonstrated
in their study that the resilience of mice subjected to chronic
social defeat stress may be associated with Bifidobacterium in
the host intestine. In fact, the oral ingestion of Bifidobacterium
significantly increased the number of resilient mice after stress.
The administration of probiotics may result in the improvement
of depressive symptoms by increasing plasma levels of tryp-
tophan, decreased concentrations of serotonin metabolites in
the frontal cortex, and dopamine metabolites in the amygdaloid
cortex (Desbonnet et al., 2008).
It is well known that resilience plays a role in stress-re-
lated psychiatric disorders such as MDD (Réus, de Moura,
Silva, Resende, & Quevedo, 2018). There are increasing
reports showing the role of gut–brain axis in resilience ver-
sus susceptibility in rodents exposed to stress (Cathomas,
Murrough, Nestler, Han, & Russo, 2019; Dantzer, Cohen,
Russo, & Dinan, 2018). In fact, it was demonstrated that sus-
ceptible male rats, which were exposed to inescapable elec-
tric stress under the learned helplessness (LH) paradigm, had
abnormal microbiota composition, including Lactobacillus,
Clostridium cluster III, and Anaerofustis, when compared to
resilient rats (Zhang, Fujita, et al., 2019; Zhang, Guo, et al.,
2019).
4
|
MICROBIOTA AND
NEUROINFLAMMATION IN MDD
4.1
|
Association among microbiota,
periphery inflammation, and brain–blood
barrier
The gut microbiota plays a key role in the formation and func-
tion of the immune system. It is believed that the first con-
tact of the immune system with the microbiota occurs during
childbirth and that this interaction has a great influence on
the immune system throughout development (Palmer, Bik,
DiGiulio, Relman, & Brown, 2010). However, the compo-
nents of breast milk, such as live microorganisms, metabo-
lites, immunoglobulin A (IgA), immune cells, and cytokines,
are associated with host responses to this early colonization
and with the formation of the microbiota (Belkaid & Hand,
2014).
Studies indicate that the gut microbiota alters CNS func-
tions, leading to the development of mood and depressive
behavior (Neufeld, Kang, Bienenstock, & Jane, 2011). For
example, mice submitted to different stressors for five weeks
displayed depressive-like behavior and elevated levels of
IL-1 in the hippocampus (Goshen et al., 2008). In addition,
the human oral intake of probiotic B.infantis 35,624 is asso-
ciated with increased expression of IL-10 in peripheral blood
(Bilbo & Schwarz, 2012). Valkanova, Ebmeier, and Allan
(2013) reported increased levels of IL-1β, IL-6, and TNF-α
in the serum of depressed individuals. Therefore, the balance
between the brain and the gut can be modulated by the im-
mune system (Bengmark, 2013).
The immune system has a symbiotic relationship between
host and microbiota, and the disruption of this dynamic
interaction may lead to the development of CNS disor-
ders (Hooper, Littman, & Macpherson, 2012). Pathogen-
associated molecular patterns (PAMPs), such as LPS, are
recognized by pattern recognition receptors (Barton, 2008).
These receptors play an important role in the recognition of
microbial targets for inflammation, and one of the families
of these receptors is Toll-like receptors (TLRs), which are
expressed not only in innate immune cells but also in CNS-
resident cell populations, including neurons and glial cells
(Crack & Bray, 2007). TLRs recognize components of micro-
organisms, including bacteria, viruses, and fungi, and then
activate inflammatory and antimicrobial innate immune re-
sponses (Medzhitov, 2001).
The TLRs are divided into subfamilies that bind to each
other and become activated by specific binders; for exam-
ple, TLR4 recognizes LPS, TLR2 along with TLR1 or TLR6
recognizes a variety of PAMPs (Kawai & Akira, 2010),
and TLR5 recognizes bacterial flagellin (Akira, Uematsu,
& Takeuchi, 2006). After activation of a TLR, a signaling
cascade is initiated which results in activation of major in-
tracellular transcription factors such as interferon I (Giles &
Stagg, 2017) and NF-κB (Sanz & Moya-Perez, 2014). Once
activated, these cells produce numerous proinflammatory
cytokines, such as IL-1α, IL-1β, TNF-α, and IL-6 (Dantzer,
Konsman, Bluthe, & Kelley, 2000). These cytokines stim-
ulate the development of CD4+T helper 1 effector cells
(TH1) and TH17 cells. TH17 in turn produces IL-17A, IL-
17F, and IL-22, resulting in chronic inflammation (Maynard,
Elson, Hatton, & Weaver, 2012).
The gut microbiota can promote different subsets of
CD4+T cells through antigenic stimulation and activation of
6
|
CARLESSI ET AL.
immune signaling pathways. Animals treated with long-act-
ing antibiotics showed a significant reduction of Treg cells
and CD4+T cells (Cording, Fleissner, & Heimesaat, 2013).
Bacteroides fragilis promotes the development of TH1
cells via the polysaccharide A pathway (Mazmanian, Liu,
Tzianabos, & Kasper, 2005). Besides, a study demonstrated
that GF mice have a defect in Foxp3+ Tregs lymphocytes,
these cells are responsible for controlling the function and
proliferation of effector T cells, and the gut microbiota plays
an important role in the induction of Treg cells (Atarashi et
al., 2011). SFCAs also induce the proliferation of Foxp3+
Tregs by histone modification (Van Loosdregt et al., 2011).
The excessive production of inflammatory cytokines in the
periphery, such as IL-1α, IL-1β, TNF-α, and IL-6, could cross
the BBB and reach the brain-resident cells. In the brain, these
cytokines act on receptors expressed by neurons and glial
cells (Dantzer et al., 2000). One of the possible mechanisms
by which cytokines can cross the BBB is by transmembrane
diffusion, requiring low molecular weight and high liposol-
ubility to enter (Oldendorf, 1974). The intestine metabolic
products have this characteristic and therefore allow their
access through the BBB (Stilling, Dinan, & Cryan, 2014).
A study showed that GF mice had increased permeability of
BBB compared to pathogen-free mice with a normal gut flora,
and this permeability was associated with reduced expression
of junction proteins; on the other hand, the exposure of GF
adult mice to a gut microbiota free of pathogens leads to a
decrease in the BBB permeability and positively regulates the
expression of junction proteins (Braniste et al., 2014).
4.2
|
Gut microbiota, inflammasome, and
microglial activation
Studies have been hypothesized the inflammasome as a key
mediator of the neuroinflammatory responses due to stress,
and its dysregulation may be implicated in the pathophysiol-
ogy of MDD (Akosile et al., 2018; Wong et al., 2016; Zhang,
Fujita, et al., 2019; Zhang, Guo, et al., 2019). Inflammasome
activation, specially nucleotide binding and oligomeriza-
tion domain-like receptor family pyrin domain-containing
3 (NLRP3), can occur by damage-associated molecular pat-
terns (DAMPs) or PAMPs mediated by Toll-like receptor 4
(TLR4) in microglial cells or by stimulus such as oxidative
stress and recruitment and activation of caspase-1, besides
IL-1β and IL-18 (Kaufmann et al., 2017). Wong et al. (2016)
conducted a study using mice with genetic deficiency or
pharmacological inhibition of caspase-1 and demonstrated
that the caspase-1 inhibition was able to reduce depressive-
and anxiety-like behavior. Also, such effects were associated
with a protective effect on the stress response by a modula-
tion in the microbiota–gut–inflammasome–brain axis (Wong
et al., 2016). In the prefrontal cortex of rats chronically
stressed, it was found an increase in the NF-κB, NLRP3, and
IL-1β; microglial activation and astrocyte impairment were
also observed; and on other hand, the treatment with the
antidepressant fluoxetine was able to reverse these changes
(Pan, Chen, Zhang, & Kong, 2014). Westfall and Pasinetti
(2019) demonstrated that a combination of a dietary polyphe-
nolic with Lactobacillus plantarum and B.longum was able
to reduce anxiety- and depressive-like behavior induced by
chronic stress. In addition, L.plantarum and B.longum were
suggested to inhibit NLRP3-mediated generation of IL-1β in
microglia cells (Westfall & Pasinetti, 2019). By targeting the
gut microbiome using prebiotic intervention (FOS-Inulin),
it was demonstrated a reversion in age-induced changes,
mainly monocyte infiltration into the brain and microglial ac-
tivation (van de Boehme et al., 2019), suggesting that prebi-
otic-driven changes in gut microbiota composition could be
a novel strategy for the improvement of age-related neuroin-
flammatory diseases and brain function. Other study revealed
that immobilization stress-induced anxiety/depression and
colitis in mice were prevented by probiotics Lactobacillus re-
uteri NK33 and B.adolescentis NK98. NK33 and NK98 also
were able to reduce microglial activation in the hippocampus
and IL-6 and corticosterone in the blood (Jang, Lee, & Kim,
2019). A study conducted by Zhang et al. (2017) demon-
strated that anti-mouse IL-6 receptor antibody (MR16-1) ad-
ministration leads to antidepressant effects in a social defeat
stress model and improved Firmicutes/Bacteroidetes ratio in
susceptible mice, suggesting that IL-6 blockade induces an-
tidepressant effects by normalizing the altered composition
of gut microbiota. Needed, IL-6 plays an important role in
MDD. Recently, the antidepressant effects of ketamine, an
antagonist of N-methyl-D-aspartate (NMDA) receptor, were
associated with reduction in serum IL-6 levels in treatment-
resistant patients with MDD (Yang et al., 2015). In addition,
Getachew et al. (2018) revealed that after 24hr of ketamine
treatment Wistar rats had change in microbiota diversity.
More studies investigating ketamine effects in microbiota in
human with MDD and in animal models of depression could
be promising.
4.3
|
Gut microbiota and
kynurenine pathway
The kynurenine pathway (KP) plays an important role in
the development of psychiatric disorders, including schizo-
phrenia (Réus, Becker, et al., 2018; Zhu et al., 2019) and
MDD (Réus, Jansen, et al., 2015b). Moreover, the gut
microbiota seems to be related to the regulation of tryp-
tophan, which is a precursor of KP (Clarke et al., 2014).
Colonization of the gut microbiota in early life as well
as changes in its composition and diversity throughout
life influences the availability of tryptophan and plays an
|
7
CARLESSI ET AL.
important role in serotonergic signaling at the CNS level.
Modulation of host behavior by the gut microbiota can
occur through recruitment of tryptophan metabolism and
serotonergic signaling of the brain–gut axis (O'Mahony,
Clarke, Borre, Dinan, & Cryan, 2015). Microbial gut me-
tabolites, such as short-chain fatty acids, can promote the
production of serotonin through tryptophan, thus regulat-
ing this pathway and preventing the tryptophan conversion
to KP (Reigstad et al., 2015).
Tryptophan is an essential amino acid precursor of
the neurotransmitter serotonin and metabolites of the KP.
Through indoleamine-2,3-dioxygenase (IDO) enzymes,
found in all tissues, and tryptophan-2,3-dioxygenase
(TDO), located in the liver, a small part of tryptophan is
metabolized to serotonin and the remainder on the KP
(Salter & Pogson, 1985). The stimulation of inflammatory
cytokines during gut inflammation, especially interferon
gamma (IFN-γ) and IL-6, induces the production of IDO
and results in changes in the metabolism of tryptophan,
causing the shift of serotonin synthesis to the production
of kynurenine and its metabolites (Jürgens, Hainz, Fuchs,
Felzmann, & Heitger, 2009; Yeung, Terentis, King, &
Thomas, 2015). Cytokines, such as IFN-γ, IL-2, and TNF-
α, have been associated with a higher risk for MDD de-
velopment (Howren, Lamkin, & Suls, 2009), and intestinal
inflammation may lead to altered brain function and MDD
(Waclawiková & Aidy, 2018). Rodents treated with B.in-
fantis display increased tryptophan plasma concentrations
and reduced IDO activity (Desbonnet et al., 2008). In ad-
dition, the treatment with B.infantis promoted antidepres-
sant-like behavior (Desbonnet et al., 2008). GF rodents
have a decrease in the KP activity; on the other hand, a
healthy microbiota reposition in the same animal is able to
normalize the KP (Clarke et al., 2013).
The metabolites of kynurenine are divided into two path-
ways: (a) production of kynurenic acid (KYNA), which
is neuroprotector and NMDA receptor antagonist and (b)
production of quinolinic acid (QUIN), which is neurotoxic
FIGURE 1 Association between gut microbiota–brain axis and MDD. Stress situations can dysregulate the gut microbiota leading to
dysbiosis, decrease in SCFAs, and increase in the proinflammatory cytokines, mainly IL-6 and IFN-γ. This inflammatory status can induce gut
permeability and bacteria migration (leaky gut). Inflammatory cytokines induce increase in IDO, which can induce a impairment QUIN/KYNA
production. Kynurenine pathway toxic metabolites and inflammatory cytokines could lead to damage in BBB, inducing an increase in the IL-6,
IL-1β, and inflammasome NLRP3 in brain-resident cells. In the brain, the inflammatory process provokes microglial activation and astrocyte
atrophy and consequently more inflammation that can culminate in mood disorders, such as anxiety and MDD. Stress situation also can directly
affect brain-resident cells increasing inflammatory mediators and indirectly changing gut microbiota. On the other hand, prebiotics and probiotics
are able to change gut microbiota and intestinal barrier, thus indirectly decreasing inflammatory cytokines, toxic metabolic of kynurenine pathway,
and BBB permeability. BBB, brain–blood barrier; IDO, indoleamine-2, 3-dioxygenase; IFN-γ, interferon gamma; IL-6, interleukin-6; IL-1β,
interleukin-1β; KYNA, kynurenic acid; MDD; major depressive disorder, NLRP3, nucleotide binding and oligomerization domain-like receptor
family pyrin domain-containing 3; QUIN, quinolinic acid; SCFAs, short-chain fatty acids
8
|
CARLESSI ET AL.
and NMDA receptor agonist (Le Floc'h, Otten, & Merlot,
2011; Schwarcz, Bruno, Muchowski, & Wu, 2012). In the
brain, QUIN causes excitotoxicity through the stimulation of
NMDA receptors, as KYNA acts by neutralizing these ef-
fects (Szalardy et al., 2012). MDD is associated with exces-
sive production of QUIN along with the reduction of KYNA
(Savitz et al., 2015). Reduced levels of KYNA were found in
patients with MDD (Wurfel et al., 2017), while elevated lev-
els of QUIN were found in the CSF of patients with suicide
attempts (Erhardt et al., 2013). Moreover, mice submitted to
chronic stress had altered gut microbiota composition, a de-
crease in Lactobacillus, and an increase in the kynurenine
levels. Though, restoration of gut Lactobacillus levels was
sufficient to improve metabolic changes and behavioral ab-
normalities (Marin et al., 2017).
At the end of the KP, the catabolism proceeds to com-
plete oxidation forming either adenosine triphosphate (ATP)
or nicotinamide adenine dinucleotide (NAD). This complete
oxidation begins with the degradation of tryptophan by TDO
or IDO in KYN. Kynurenine is subsequently degraded to
3-hydroxyquinurenine (OHK) by quinurenine-3-mono-ox-
ygenase (KMO) or kinurenic acid (KYNA) by kynurenine
aminotransferases (KATs). OHK continues to be degraded
by kynureninase to produce 3-hydroxyanthranilic acid
(HAA). Finally, HAA can be degraded in ATP or in small
amounts of picolinic acid or quinolinic acid (QUIN) and
subsequently in NAD (Réus, Jansen, et al., 2015b).
In the brain, IDO is expressed by various brain cells, includ-
ing astrocytes and microglia (Grant, Naif, Espinosa, & Kapoor,
2000). During inflammation and increased degradation of
tryptophan, induced by inflammatory cytokines (Pemberton,
Kerr, Smythe, & Brew, 1997), excessive production of ky-
nurenine can be transported through the BBB into the brain.
The metabolites of this pathway contribute to neuroprotec-
tive and/ or neurodegenerative changes in the brain (Myint,
Schwarz, & Muller, 2012). Microglia and macrophages appear
to be involved in the production of QUIN and KYNA in the
astrocytes. During infection, infiltration of activated macro-
phages and microglia activation in the brain may result in the
contribution of neurotoxicity astrocyte-mediated (Guillemin et
al., 2001). MDD is associated with an inflammatory state that
may stimulate the production of the neuropathic pathway of
KP, maintenance glial cells in constant imbalance which could
contribute to the recurrent and chronic nature of MDD (Myint
& Kim, 2003; Réus, Jansen, et al., 2015b).
5
|
CONCLUSION
The gut microbiota–brain axis dysregulation is evident in
MDD. This association has been reported by human and
animal studies. Stress situations could lead to an impair-
ment in the gut microbiota that in turn lead a production of
inflammatory mediators, mainly IFN-γ and IL-6, and a de-
crease in SCFAs. Changes in the gut microbiota could impact
the gut barrier and to produce higher levels of inflammatory
cytokines in the blood. In the periphery, the kynurenine path-
ways, metabolic, mainly toxic metabolic, are influenced by
changes in the gut microbiota and inflammatory mediators.
From periphery, toxic products from a microbiota impair-
ment or proinflammatory cytokines could disrupt BBB and
cross to the brain increasing cytokines, such as IL-6, IL-1β,
and the inflammasome NLPR3 in the brain-resident cells.
Microglial and astrocytes are influenced, suffering activation
and atrophy, respectively (Figure 1). Changes in these glial
cells influence brain networks involved in memory, learning,
emotions, and mood regulation, what could be behind the
onset of depressive symptoms or anxiety.
ACKNOWLEDGEMENTS
The Translational Psychiatry Program (USA) is funded
by the Department of Psychiatry and Behavioral Sciences,
McGovern Medical School, The University of Texas Health
Science Center at Houston (UTHealth). The Center of
Excellence on Mood Disorders (USA) is funded by the Pat
Rutherford Jr. Chair in Psychiatry, John S. Dunn Foundation,
and Anne and Don Fizer Foundation Endowment for
Depression Research. Translational Psychiatry Laboratory
(Brazil) is one of the centers of the National Institute for
Molecular Medicine (INCT-MM) and one of the members of
the Center of Excellence in Applied Neurosciences of Santa
Catarina (NENASC). Its research is supported by grants
from CNPq (JQ and GZR), FAPESC (JQ and GZR), Instituto
Cérebro e Mente (JQ, AIZ and GZR), and UNESC (JQ, GZR,
and AIZ). JQ is a 1A CNPq Research Fellow.
CONFLICT OF INTEREST
JQ has provided clinical research support to Janssen
Pharmaceutical and Allergan; has served on the speaker bu-
reau for Daiichi Sankyo, and on other advisory boards and
speaker bureaus as expert witness or consultant; is a stock-
holder for Instituto de Neurociencias; and is a copyright
holder for Artmed Editora and Artmed Panamericana. ASC,
LAB, AIZ, and GZR have no conflict of interest.
AUTHOR CONTRIBUTIONS
ASC, LAB, AIZ, JQ, and GZR contributed with writing
manuscript. ASC prepared the figure. GZR contributed with
study design and manuscript revision.
DATA AVAILABILITY STATEMENT
The data associated with this review paper are available on
pubmed.gov
ORCID
Gislaine Z. Réus https://orcid.org/0000-0001-6073-0855
|
9
CARLESSI ET AL.
REFERENCES
Abbasi, S. H., Hosseini, F., Modabbernia, A., Ashrafi, M., &
Akhondzadeh, S. (2012). Effect of celecoxib add-on treatment on
symptoms and serum IL-6 concentrations in patients with major
depressive disorder: Randomized double-blind placebo-controlled
study. Journal of Affective Disorders, 141, 308–314. https ://doi.
org/10.1016/j.jad.2012.03.033
Aguiar, C. C., Castro, T. R., Carvalho, A. F., Vale, O. C., Sousa, F. C.,
& Vasconcelos, S. M. (2011). Drogas antidepressivas. Acta Médica
Portuguesa, 24, 91–98.
Akira, S., Uematsu, S., & Takeuchi, O. (2006). Pathogen recognition
and innate immunity. Cell, 4, 783–801.
Akosile, W., Voisey, J., Lawford, B., Colquhounc, D., Young, R. M., &
Mehta, D. (2018). The inflammasome NLRP12 is associated with
both depression and coronary artery disease in Vietnam veterans.
Psychiatry Research, 270, 775–779. https ://doi.org/10.1016/j.psych
res.2018.10.051
Amidfar, M., Réus, G. Z., Quevedo, J., & Kim, Y. K. (2018). The
role of memantine in the treatment of major depressive disorder:
Clinical efficacy and mechanisms of action. European Journal
of Pharmacology, 827, 103–111. https ://doi.org/10.1016/j.
ejphar.2018.03.023
Amidfar, M., Réus, G. Z., Quevedo, J., Kim, Y. K., & Arbabi, M. (2017).
Effect of co-administration of memantine and sertraline on the an-
tidepressant-like activity and brain-derived neurotrophic factor
(BDNF) levels in the rat brain. Brain Research Bulletin, 128, 29–33.
https ://doi.org/10.1016/j.brain resbu ll.2016.11.003
Asano, Y., Hiramoto, T., Nishino, R., Aiba, Y., Kimura, T., Yoshihara,
K., … Sudo, N. (2012). Critical role of gut microbiota in the produc-
tion of biologically active, free catecholamines in the gut lumen of
mice. American Journal of Physiology. Gastrointestinal and Liver
Physiology, 303, 1288–1295.
Atarashi, K., Tanoue, T., Shima, T., Imaoka, A., Kuwahara, T., Momose,
Y., … Honda, K. (2011). Induction of colonic regulatory T cells by
indigenous Clostridium species. Science, 331, 337–341. https ://doi.
org/10.1126/scien ce.1198469
Bailey, M. T., & Coe, C. L. (1999). Maternal separation disrupts the
integrity of the intestinal microflora in infant rhesus monkeys.
Developmental Psychobiology, 35, 146–155. https ://doi.org/10.1002/
(sici)1098-2302(19990 9)35:2<146::aid-dev7>3.0.co;2-g
Barton, G. M. (2008). A calculated response: Control of inflammation
by the innate immune system. Journal of Clinical Investigation, 118,
413–420. https ://doi.org/10.1172/jci34431
Bekhbat, M., Howell, P. A., Rowson, S. A., Kelly, S. D., Tansey, M. G.,
& Neigh, G. N. (2019). Chronic adolescent stress sex-specifically
alters central and peripheral neuro-immune reactivity in rats. Brain,
Behavior, and Immunity, 76, 248–257. https ://doi.org/10.1016/j.
bbi.2018.12.005
Belkaid, Y., & Hand, T. (2014). Role of the Microbiota in Immunity
and inflammation. Cell, 157, 121–141. https ://doi.org/10.1016/j.
cell.2014.03.011
Bengmark, S. (2013). Gut microbiota, immune development and
function. Pharmacological Research, 69, 87–113. https ://doi.
org/10.1016/j.phrs.2012.09.002
Bercik, P., Denou, E., Collins, J., Jackson, W., Lu, J., Jury, J., …
Collins, S. M. (2011b). The intestinal microbiota affect cen-
tral levels of brain-derived neurotropic factor and behavior in
mice. Gastroenterology, 141, 599–609. https ://doi.org/10.1053/j.
gastro.2011.04.052
Bercik, P., Park, A. J., Sinclair, D., Khoshdel, A., Lu, J., Huang, X.,
Verdu, E. F. (2011a). The anxiolytic effect of Bifidobacterium
longum NCC3001 involves vagal pathways for gut-brain communi-
cation. Neurogastroenterology and Motility, 23, 1132–1139. https ://
doi.org/10.1111/j.1365-2982.2011.01796.x
Bercik, P., Verdu, E. F., Foster, J. A., Macri, J., Potter, M., Huang, X.,
… Collins, S. M. (2010). Chronic gastrointestinal inflammation in-
duces anxiety-like behavior and alters central nervous system bio-
chemistry in mice. Gastroenterology, 139(2102–2112), e2101. https
://doi.org/10.1053/j.gastro.2010.06.063
Bermudez-Brito, M., Plaza-Díaz, J., Muñoz-Quezada, S., Gómez-
Llorente, C., & Gil, A. (2012). Probiotic mechanisms of action.
Annals of Nutrition & Metabolism, 61, 160–174. https ://doi.
org/10.1159/00034 2079
Berton, O., & Nestler, E. J. (2006). New approaches to antidepressant
drug discovery: Beyond monoamines. Nature Reviews Neuroscience,
7, 137–151. https ://doi.org/10.1038/nrn1846
Bilbo, S. D., & Schwarz, J. M. (2012). The immune system and de-
velopmental programming of brain and behavior. Frontiers in
Neuroendocrinology, 33, 267–286. https ://doi.org/10.1016/j.
yfrne.2012.08.006
Braniste, V., Al-Asmakh, M., Kowal, C., Anuar, F., Abbaspour, A., Tóth,
M., … Pettersson, S. (2014). The gut microbiota influences blood-
brain barrier permeability in mice. Science Translational Medicine,
6, 263ra158. https ://doi.org/10.1126/scitr anslm ed.3009759
Buffington, S. A. D., Prisco, G. V., Auchtung, T. A., Ajami, N. J.,
Petrosino, J. F., & Costa-Mattioli, M. (2016). Microbial reconsti-
tution reverses maternal diet-induced social and synaptic defi-
cits in offspring. Cell, 165, 1762–1775. https ://doi.org/10.1016/j.
cell.2016.06.001
Calogero, A. E., Gallucci, W. T., Chrousos, G. P., & Gold, P. W. (1988).
Catecholamine effects upon rat hypothalamic corticotropin-releas-
ing hormone secretion in vitro. Journal of Clinical Investigation, 82,
839–846. https ://doi.org/10.1172/jci11 3687
Cathomas, F., Murrough, J. W., Nestler, E. J., Han, M. H., & Russo, S.
J. (2019). Neurobiology of resilience: Interface between mind and
body. Biological Psychiatry, 15, 410–420.
Choi, H. J., Lee, N. K., & Paik, H. D. (2015). Health benefits of lactic
acid bacteria isolated from kimchi, with respect to immunomodula-
tory effects. Food Science and Biotechnology, 24, 783–789. https ://
doi.org/10.1007/s10068-015-0102-3
Clarke, G., Grenham, S., Scully, P., Fitzgerald, P., Moloney, R. D.,
Shanahan, F., … Cryan, J. F. (2013). The microbiome-gut-brain axis
during early life regulates the hippocampal serotonergic system in a
sex-dependent manner. Molecular Psychiatry, 18, 666–673. https ://
doi.org/10.1038/mp.2012.77
Clarke, G., Stilling, R. M., Kennedy, P. J., Stanton, C., Cryan, J. F.,
& Dinan, T. G. (2014). Minireview: Gut microbiota: The neglected
endocrine organ. Molecular Endocrinology, 28, 1221–1238. https ://
doi.org/10.1210/me.2014-1108
Cording, S., Fleissner, D., & Heimesaat, M. M. (2013). Commensal
microbiota drive proliferation of conventional and Foxp3+ regula-
tory CD4+ T cells in mesenteric lymph nodes and Peyer's patches.
European Journal of Clinical Microbiology and Infectious Diseases,
3, 1–10. https ://doi.org/10.1556/eujmi.3.2013.1.1
Crack, P. J., & Bray, P. J. (2007). Toll-like receptors in the brain and their
potential roles in neuropathology. Immunology and Cell Biology, 85,
476–480. https ://doi.org/10.1038/sj.icb.7100103
Curtis, K., Stewart, C. J., Robinson, M., Molfese, D. L., Gosnell, S. N.,
Kosten, T. R., … Salas, R. (2019). Insular resting state functional
10
|
CARLESSI ET AL.
connectivity is associated with gut microbiota diversity. European
Journal of Neuroscience, 50, 2446–2452. https ://doi.org/10.1111/
ejn.14305
Dantzer, R., Cohen, S., Russo, S. J., & Dinan, T. G. (2018). Resilience
and immunity. Brain, Behavior, and Immunity, 74, 28–42. https ://
doi.org/10.1016/j.bbi.2018.08.010
Dantzer, R., Konsman, J. P., Bluthe, R. M., & Kelley, K. W. (2000).
Neural and humoral pathways of communication from the immune
system to the brain: Parallel or convergent? Autonomic Neuroscience,
85, 60–65. https ://doi.org/10.1016/S1566-0702(00)00220-4
Delpech, J. C., Wei, L., Hao, J., Yu, X., Madore, C., Butovsky, O., &
Kaffman, A. (2016). Early life stress perturbs the maturation of
microglia in the developing hippocampus. Brain, Behavior, and
Immunity, 57, 79–93. https ://doi.org/10.1016/j.bbi.2016.06.006
Desbonnet, L., Garrett, L., Clarke, G., Bienenstock, J., & Dinan,
T. G. (2008). The probiotic Bifidobacteria infantis: An assess-
ment of potential antidepressant properties in the rat. Journal of
Psychiatric Research, 43, 164–174. https ://doi.org/10.1016/j.jpsyc
hires.2008.03.009
Desbonnet, L., Garrett, L., Clarke, G., Kiely, B., Cryan, J. F., & Dinan,
T. G. (2010). Effects of the probiotic bididobacterium infantis in the
maternal separation model of depression. Neuroscience, 170, 1179–
1188. https ://doi.org/10.1016/j.neuro scien ce.2010.08.005
Diaz Heijtz, R., Wang, S., Anuar, F., Qian, Y., Björkholm, B.,
Samuelsson, A., Pettersson, S. (2011). Normal gut microbiota
modulates brain development and behavior. Proceedings of the
National Academy of Sciences of the United States of America, 108,
3047–3052.
Dinan, T. G., & Cryan, J. F. (2012). Regulation of the stress response
by the gut microbiota: Implications for psychoneuroendocrinol-
ogy. Psychoneuroendocrinology, 37, 1369–1378. https ://doi.
org/10.1016/j.psyne uen.2012.03.007
Duranti, S., Milani, C., Lugli, G. A., Mancabelli, L., Turroni, F.,
Ferrario, C., … Ventura, M. (2016). Evaluation of genetic diver-
sity among strains of the human gut commensal Bifidobacterium
adolescentes. Scientific Reports, 6, 23971. https ://doi.org/10.1038/
srep2 3971
Elhwuegi, A. S. (2004). Central monoamines and their role in major de-
pression. Progress in Neuro-Psychopharmacology and Biological
Psychiatry, 28, 435–451. https ://doi.org/10.1016/j.pnpbp.2003.11.018
Enache, D., Pariante, C. M., & Mondelli, V. (2019). Markers of cen-
tral inflammation in major depressive disorder: A systematic review
and meta-analysis of studies examining cerebrospinal fluid, positron
emission tomography and post-mortem brain tissue. Brain, Behavior,
and Immunity, 81, 24–40. https ://doi.org/10.1016/j.bbi.2019.06.015
Erhardt, S., Lim, C. K., Linderholm, K. R., Janelidze, S., Lindqvist, D.,
Samuelsson, M., … Brundin, L. (2013). Connecting inflammation
with glutamate agonism in suicidality. Neuropsychopharmacology,
38, 743–752. https ://doi.org/10.1038/npp.2012.248
Evrensel, A., & Ceylan, M. E. (2015). The gut-brain axis: The missing
link in depression. Clinical Psychopharmacology and Neuroscience,
13, 239–244. https ://doi.org/10.9758/cpn.2015.13.3.239
Ferrari, A. J., Charlson, F. J., Norman, R. E., Patten, S. B., Freedman,
G., Murray, C. J., … Whiteford, H. A. (2013). Burden of depressive
disorders by country, sex, age, and year: Findings from the global
burden of disease study 2010. PLoS Medicine, 10, e1001547. https
://doi.org/10.1371/journ al.pmed.1001547
Forsythe, P., Sudo, N., Dinan, T., Taylor, V. H., & Bienenstock, J.
(2010). Mood and gut feelings. Brain, Behavior, and Immunity, 24,
9–16. https ://doi.org/10.1016/j.bbi.2009.05.058
Foster, J. A., & McVey Neufeld, K. A. (2013). Gut-brain axis: How
the microbiome influences anxiety and depression. Trends
in Neurosciences, 36, 305–312. https ://doi.org/10.1016/j.
tins.2013.01.005
Fu, X., Zunich, S. M., O'Connor, J. C., Kavelaars, A., Dantzer, R., &
Kelley, K. W. (2010). Central administration of lipopolysaccha-
ride induces depressive-like behavior in vivo and activates brain-
indoleamine 2,3 dioxygenase in murine organotypic hippocampal
slice cultures. Journal of Neuroinflammation, 7, 43. https ://doi.
org/10.1186/1742-2094-7-43
Getachew, B., Aubee, J. I., Schottenfeld, R. S., Csoka, A. B., Thompson,
K. M., & Tizabi, Y. (2018). Ketamine interactions with gut-micro-
biota in rats: Relevance to its antidepressant and anti-inflammatory
properties. BMC Microbiology, 22, 222. https ://doi.org/10.1186/
s12866-018-1373-7
Giles, E. M., & Stagg, A. J. (2017). Type 1 interferon in the human
intestine-A Co-ordinator of the immune response to the micro-
biota. Inflammatory Bowel Diseases, 23, 524–533. https ://doi.
org/10.1097/mib.00000 00000 001078
Giridharan, V. V., Réus, G. Z., Selvaraj, S., Scaini, G., Barichello, T., &
Quevedo, J. (2019). Maternal deprivation increases microglial acti-
vation and neuroinflammatory markers in the prefrontal cortex and
hippocampus of infant rats. Journal of Psychiatric Research, 115,
13–20. https ://doi.org/10.1016/j.jpsyc hires.2019.05.001
Goshen, I., Kreisel, T., Ben-Menachem-Zidon, O., Licht, T., Weidenfeld,
J., Ben-Hur, T., & Yirmiya, R. (2008). Brain interleukin-1 medi-
ates chronic stress-induced depression in mice via adrenocortical
activation and hippocampal neurogenesis suppression. Molecular
Psychiatry, 13, 717–728. https ://doi.org/10.1038/sj.mp.4002055
Grant, R. S., Naif, H., Espinosa, M., & Kapoor, V. (2000). IDO induc-
tion in IFN-gamma activated astroglia: A role in improving cell via-
bility during oxidative stress. Redox Report, 5, 101–104. https ://doi.
org/10.1179/13510 00001 01535357
Guillemin, G. J., Kerr, S. J., Smythe, G. A., Smith, D. G., Kapoor,
V., Armati, P. J., … Brew, B. J. (2001). Kynurenine pathway
metabolism in human astrocytes: A paradox for neuronal pro-
tection. Journal of Neurochemistry, 78, 842–853. https ://doi.
org/10.1046/j.1471-4159.2001.00498.x
Guo, Y., Xie, J. P., Deng, K., Li, X., Yuan, Y., Xuan, Q., … Luo, H.
(2019). Prophylactic effects of bifidobacterium adolescentis on
anxiety and depression-like phenotypes after chronic stress: A role
of the gut microbiota-inflammation axis. Frontiers in Behavioural
Neurosciences, 13, 126. https ://doi.org/10.3389/fnbeh.2019.00126
Hannan, A. J. (2016). Thinking with your stomach? Gut feelings on mi-
crobiome modulation of brain structure and function (Commentary
on Luczynski et al.). European Journal of Neuroscience, 44, 2652–
2653. https ://doi.org/10.1111/ejn.13399
He, M. C., Shi, Z., Sha, N. N., Chen, N., Peng, S. Y., Liao, D. F.,
Zhang, Y. (2019). Paricalcitol alleviates lipopolysaccharide-induced
depressive-like behavior by suppressing hypothalamic microglia ac-
tivation and neuroinflammation. Biochemical Pharmacology, 163,
1–8. https ://doi.org/10.1016/j.bcp.2019.01.021
Hill, C., Guarner, F., Reid, G., Gibson, G. R., Merenstein, D. J., Pot,
B., … Sanders, M. E. (2014). Expert consensus document the
International Scientific Association for Probiotics and Prebiotics
consensus statement on the scope and appropriate use of the term
probiotic. Nat Rev Gastroenterol Hepatol., 11(8), 506–514. https ://
doi.org/10.1038/nrgas tro.2014.66.
Hofford, R. S., Russo, S. J., & Kiraly, D. D. (2018). Neuroimmune
mechanisms of psychostimulant and opioid use disorders. European
|
11
CARLESSI ET AL.
Journal of Neuroscience, 50, 2562–2573. https ://doi.org/10.1111/
ejn.14143
Holzer, P., & Farzi, A. (2014). Neuropeptides and the microbiotagut-
brain axis. Advances in Experimental Medicine and Biology, 817,
195–219. https ://doi.org/10.1007/978-1-4939-0897-4_9
Hooper, L. V., Littman, D. R., & Macpherson, A. J. (2012). Interactions
between the microbiota and the immune system. Science, 336,
1268–1273. https ://doi.org/10.1126/scien ce.1223490
Howren, M. B., Lamkin, D. M., & Suls, J. (2009). Associations of de-
pression with C-reactive protein, IL-1, and IL-6: A meta-analysis.
Psychosomatic Medicine, 71, 171–186. https ://doi.org/10.1097/
PSY.0b013 e3181 907c1b
Hsiao, E. Y., McBride, S. W., Hsien, S., Sharon, G., Hyde, E. R., McCue,
T., … Mazmanian, S. K. (2013). Microbiota modulate behavioral
and physiological abnormalities associated with neurodevelop-
mental disorders. Cell, 155, 1451–1463. https ://doi.org/10.1016/j.
cell.2013.11.024
Jang, H. M., Lee, K. E., & Kim, D. H. (2019). The preventive and cura-
tive effects of Lactobacillus reuteri NK33 and Bifidobacterium ad-
olescentes NK98 on immobilization stress-induced anxiety/depres-
sion and colitis in mice. Nutrients, 11, 819. https ://doi.org/10.3390/
nu110 40819
Jeon, S. W., & Kim, Y. K. (2016). Neuroinflammation and cytokine ab-
normality in major depression: Cause or consequence in that illness?
World Journal of Psychiatry, 6, 283–293. https ://doi.org/10.5498/
wjp.v6.i3.283
Jiang, H., Ling, Z., Zhang, Y., Mao, H., Ma, Z., Yin, Y., … Ruan, B.
(2015). Altered fecal microbiota composition in patients with major
depressive disorder. Brain, Behavior, and Immunity, 48, 186–194.
https ://doi.org/10.1016/j.bbi.2015.03.016
Johnson, F. K., & Kaffman, A. (2018). Early life stress perturbs the
function of microglia in the developing rodent brain: New insights
and future challenges. Brain, Behavior, and Immunity, 69, 18–27.
https ://doi.org/10.1016/j.bbi.2017.06.008
Jürgens, B., Hainz, U., Fuchs, D., Felzmann, T., & Heitger, A. (2009).
Interferon-gamma-triggered indoleamine 2,3-dioxygenase compe-
tence in human monocyte-derived dendritic cells induces regulatory
activity in allogeneic T cells. Blood, 114, 3235–3243. https ://doi.
org/10.1182/blood-2008-12-195073
Kaufmann, F. N., Costa, A. P., Ghisleni, G., Diaz, A. P., Rodrigues,
A. L. S., Peluffo, H., & Kaster, M. P. (2017). NLRP3 inflam-
masome-driven pathways in depression: Clinical and preclinical
findings. Brain, Behavior, and Immunity, 64, 367–383. https ://doi.
org/10.1016/j.bbi.2017.03.002
Kawai, T., & Akira, S. (2010). The role of pattern-recognition re-
ceptors in innate immunity: Update on toll-like receptors. Nature
Immunology, 11, 373–8410. https ://doi.org/10.1038/ni.1863
Kelly, J. R., Borre, Y., O'Brien, C., Patterson, E., El Aidy, S., Deane,
J., … Dinan, T. G. (2016). Transferring the blues: Depression-
associated gut microbiota induces neurobehavioural changes in
the rat. Journal of Psychiatric Research, 82, 109–118. https ://doi.
org/10.1016/j.jpsyc hires.2016.07.019
Kennedy, P. J., Cryan, J. F., Dinan, T. G., & Clarke, G. (2014). Irritable
bowel syndrome: A microbiome-gut-brain axis disorder? World
Journal of Gastroenterology, 20, 14105–14125.
Ketchesin, K. D., Becker-Krail, D., & McClung, C. A. (2018).
Mood-related central and peripheral clocks. European Journal of
Neuroscience, https ://doi.org/10.1111/ejn.14253
Kiliaan, A. J., Saunders, P. R., Bijlsma, P. B., Berin, M. C., Taminiau,
J. A., Groot, J. A., & Perdue, M. H. (1998). Stress stimulates
transepithelial macromolecular uptake in rat jejunum. American
Journal of Physiology, 275, G1037–G1044. https ://doi.org/10.1152/
ajpgi.1998.275.5.G1037
Kim, N., Yun, M., On, Y. J., & Choi, H. J. (2018). Mind-altering
with the gut: Modulation of the gut-brain axis with probiotics.
Journal of Microbiology, 56, 172–182. https ://doi.org/10.1007/
s12275-018-8032-4
Köhler, O., Krogh, J., Mors, O., & Benros, M. E. (2016). Inflammation
in depression and the potential for anti-inflammatory treat-
ment. Current Neuropharmacology, 14, 732–742. https ://doi.
org/10.2174/15701 59x14 66615 12081 13700
Köhler-Forsberg, O., N. Lydholm, C., Hjorthøj, C., Nordentoft, M.,
Mors, O., & Benros, M. E. (2019). Efficacy of anti-inflammatory
treatment on major depressive disorder or depressive symptoms:
Meta-analysis of clinical trials. Acta Psychiatrica Scandinavica,
139, 404–419. https ://doi.org/10.1111/acps.13016
Krishnan, V., & Nestler, E. J. (2008). The molecular neurobiology of
depression. Nature, 455, 894–902. https ://doi.org/10.1038/natur
e07455
Le Floc'h, N., Otten, W., & Merlot, E. (2011). Tryptophan metabolism,
from nutrition to potential therapeutic applications. Amino Acids,
41, 1195–1205. https ://doi.org/10.1007/s00726-010-0752-7
Liao, L. Z., Zeng, B. H., Wang, W., Li, G. H., Wu, F., Wang, L.,
Fang, X. (2016). Impact of the consumption of tea polyphenols on
early atherosclerotic lesion formation and intestinal Bifidobacteria
in high-fat-fed ApoE/ mice. Frontiers in Nutrition, 3, 42. https ://
doi.org/10.3389/fnut.2016.00042
Liu, C. H., Zhang, G. Z., Li, B., Li, M., Woelfer, M., Walter, M., &
Wang, L. (2019). Role of inflammation in depression relapse.
Journal of Neuroinflammation, 16, 90. https ://doi.org/10.1186/
s12974-019-1475-7
Luczynski, P., Whelan, S. O., O'Sullivan, C., Clarke, G., Shanahan, F.,
Dinan, T. G., & Cryan, J. F. (2016). Adult microbiota-deficient mice
have distinct dendritic morphological changes: Differential effects in
the amygdala and hippocampus. European Journal of Neuroscience,
44, 2654–2666. https ://doi.org/10.1111/ejn.13291
Macfarlane, S., & Macfarlane, G. T. (2003). Regulation of short-chain
fatty acid production. The Proceedings of the Nutrition Society, 62,
67–72. https ://doi.org/10.1079/pns20 02207
Machado-Vieira, R., Salvadore, G., Ibrahim, L. A., Diaz-GranadoS,
N., Zarate, J., & Carlos, A. (2009). Targeting glutamatergic signal-
ing for the development of novel therapeutics for mood disorders.
Current Pharmaceutical Design, 15, 1595–1611.
Maes, M. (2008). The cytokine hypothesis of depression: Inflammation,
oxidative & nitrosative stress (IO & NS) and leaky gut as new tar-
gets for adjunctive treatments in depression. Neuro Endocrinology
Letters, 29, 287–291.
Marin, I. A., Goertz, J. E., Ren, T., Rich, S. S., Onengut-Gumuscu,
S., Farber, E., … Gaultier, A. (2017). Microbiota alteration is as-
sociated with the development of stress-induced despair behavior.
Scientific Reports, 7, 43859. https ://doi.org/10.1038/srep4 3859
Maynard, C. L., Elson, C. O., Hatton, R. D., & Weaver, C. T. (2012).
Reciprocal interactions of the intestinal microbiota and immune sys-
tem. Nature, 489, 231–241. https ://doi.org/10.1038/natur e11551
Mazmanian, S. K., Liu, C. H., Tzianabos, A. O., & Kasper, D. L. (2005).
An immunomodulatory molecule of symbiotic bacteria directs mat-
uration of the host immune system. Cell, 122, 107–118. https ://doi.
org/10.1016/j.cell.2005.05.007
Medzhitov, R. (2001). Toll-like receptors and innate immunity. Nature
Reviews Immunology, 1, 135–145. https ://doi.org/10.1038/35100529
12
|
CARLESSI ET AL.
Messaoudi, M., Lalonde, R., Violle, N., Javelot, H., Desor, D., Nejdi, A.,
… Cazaubiel, J. M. (2011). Assessment of psychotropic-like proper-
ties of a probiotic formulation (Lactobacillus helveticus R0052 and
Bifidobacterium longum R0175) in rats and human subjects. British
Journal of Nutrition, 105, 755–764. https ://doi.org/10.1017/s0007
11451 0004319
Mika, A., Day, H. E., Martinez, A., Rumian, N. L., Greenwood, B. N.,
Chichlowski, M., … Fleshner, M. (2017). Early life diets with pre-
biotics and bioactive milk fractions attenuate the impact of stress
on learned helplessness behaviours and alter gene expression within
neural circuits important for stress resistance. European Journal of
Neuroscience, 45, 342–357. https ://doi.org/10.1111/ejn.13444
Millan, M. J. (2006). Multi-target strategies for the improved treatment
of depressive states: Conceptual foundations and neuronal sub-
strates, drug discovery and therapeutic application. Pharmacology
& Therapeutics, 111, 135–370. https ://doi.org/10.1016/j.pharm
thera.2005.11.006
Müller, N., Krause, D., Barth, R., Myint, A. M., Weidinger, E.,
Stettinger, W., … Schwarz, M. J. (2019). Childhood Adversity and
Current Stress are related to Pro-and Antiinflammatory Cytokines
in Major Depression. Journal of Affective Disorders, 253, 270–276.
https ://doi.org/10.1016/j.jad.2019.04.088
Munjiza, A., Kostic, M., Pesic, D., Gajic, M., Markovic, I., & Tosevski,
D. L. (2018). Higher concentration of interleukin 6 - A possi-
ble link between major depressive disorder and childhood abuse.
Psychiatry Research, 264, 26–30. https ://doi.org/10.1016/j.psych
res.2018.03.072
Myint, A. M., & Kim, Y. K. (2003). Cytokine-serotonin interac-
tion through IDO: A neurodegeneration hypothesis of depres-
sion. Medical Hypotheses, 61, 519–525. https ://doi.org/10.1016/
S0306-9877(03)00207-X
Myint, A., Schwarz, M. J., & Muller, N. (2012). The role of the kynurenine
metabolism in major depression. Journal of Neural Transmission,
119, 245–251. https ://doi.org/10.1007/s00702-011-0741-3
Nakamura, M., Okada, S., Toyama, Y., & Okano, H. (2005). Role of
IL-6 in spinal cord injury in a mouse model. Clinical Reviews in
Allergy and Immunology, 28, 197–204. https ://doi.org/10.1385/
CRIAI :28:3:197
Nestler, E. J., Barrot, M., Dileone, R. J., Eisch, A. J., Gold, S. J., &
Monteggia, L. M. (2002). Neurobiology of depression. Neuron, 34,
13–25. https ://doi.org/10.1016/S0896-6273(02)00653-0
Neufeld, K. A., Kang, N., Bienenstock, J., & Jane, A. (2011). Effects
of intestinal microbiota on anxiety-like behavior. Communicative &
Integrative Biology, 4, 492–494. https ://doi.org/10.4161/cib.15702
Oldendorf, W. H. (1974). Lipid solubility and drug penetration of the
blood brain barrier. Proceedings of the Society for Experimental
Biology and Medicine, 147, 813–815.
O'Mahony, S. M., Clarke, G., Borre, Y. E., Dinan, T. G., & Cryan, J.
F. (2015). Serotonin, tryptophan metabolism and the brain-gut-mi-
crobiome axis. Behavioral Brain Research, 277, 32–48. https ://doi.
org/10.1016/j.bbr.2014.07.027
O'Mahony, S. M., Neufeld Mc Vey, K. A., Waworuntu, R. V., Pusceddu,
M. M., Manurung, S., Murphy, K., … Cryan, J. F. (2019). The en-
during effects of early-life stress on the microbiota–gut–brain axis
are buffered by dietary supplementation with milk fat globule mem-
brane and a prebiotic blend. European Journal of Neuroscience,
https ://doi.org/10.1111/ejn.14514
Overduin, J., Schoterman, M. H., Calame, W., Schonewille, A. J., & Ten
Bruggencate, S. J. (2013). Dietary galacto-oligosaccharides and cal-
cium: Effects on energy intake, fat-pad weight and satiety-related,
gastrointestinal hormones in rats. British Journal of Nutrition, 109,
1338–1348. https ://doi.org/10.1017/S0007 11451 2003066
Özogul, F. (2011). Effects of specific lactic acid bacteria species on
biogenic amine production by foodborne pathogen. International
Journal of Food Science & Technology, 46, 478–484. https ://doi.
org/10.1111/j.1365-2621.2010.02511.x
Palmer, C., Bik, E. M., DiGiulio, D. B., Relman, D. A., & Brown, P.
O. (2010). Development of the human infant intestinal microbiota.
PLoS Biolog., 5, e1566–e1573.
Pan, Y., Chen, X. Y., Zhang, Q. Y., & Kong, L. D. (2014). Microglial
NLRP3 inflammasome activation mediates IL-1β-related inflam-
mation in pré-frontal cortex of depressive rats. Brain, Behavior, and
Immunity, 41, 90–100. https ://doi.org/10.1016/j.bbi.2014.04.007
Pemberton, L. A., Kerr, S. J., Smythe, G., & Brew, B. J. (1997).
Quinolinic acid production by macrophages stimulated with
IFN-gamma, TNFalpha, and IFN-alpha. Journal of Interferon
and Cytokine Research, 17, 589–595. https ://doi.org/10.1089/
jir.1997.17.589
Reeves, W. C., Strine, T. W., Pratt, L. A., Thompson, W., Ahluwalia, I.,
Dhingra, S. S., … Safran, M. A. (2011). Mental illness surveillance
among adults in the United States. Morbidity and Mortality Weekly
Report, 60, 601–629.
Reigstad, C. S., Salmonson, C. E., Rainey, J. F., Szurszewski, J. H.,
Linden, D. R., Sonnenburg, J. L., Kashyap, P. C. (2015). Gut
microbes promote colonic serotonin production through an effect
of short-chain fatty acids on enterochromaffin cells. The FASEB
Journal, 29, 1395–1403. https ://doi.org/10.1096/fj.14-259598
Réus, G. Z., Abelaira, H. M., Tuon, T., Titus, S. E., Ignácio, Z. M.,
Rodrigues, A. L. S., & Quevedo, J. (2016). Glutamatergic NMDA
Receptor as Therapeutic Target for Depression. Advances in Protein
Chemistry and Structural Biology, 103, 169–202.
Réus, G. Z., Becker, I. R. T., Scaini, G., Petronilho, F., Oses, J. P.,
Kaddurah-Daouk, R., … Barichello, T. (2018). The inhibition
of the kynurenine pathway prevents behavioral disturbances and
oxidative stress in the brain of adult rats subjected to an animal
model of schizophrenia. Progress in Neuro-Psychopharmacology
and Biological Psychiatry, 2, 55–63. https ://doi.org/10.1016/j.
pnpbp.2017.10.009
Réus, G. Z., de Fernandes, G. C., Moura, A. B., Silva, R. H., de Darabas,
A. C., Souza, T. G., Quevedo, J. (2017). Early life experience
contributes to the developmental programming of depressive-like
behaviour, neuroinflammation and oxidative stress. Journal of
Psychiatric Research, 95, 196–207.
Réus, G. Z., de Moura, A. B., Silva, R. H., Resende, W. R., & Quevedo,
J. (2018). Resilience dysregulation in major depressive disorder:
Focus on glutamatergic imbalance and microglial activation. Current
Neuropharmacology, 5, 297–307. https ://doi.org/10.2174/15701
59X15 66617 06301 64715
Réus, G. Z., de Silva, R. H., Moura, A. B., Presa, J. F., Abelaira,
H. M., Abatti, M., Quevedo, J. (2019). Early Maternal
Deprivation Induces Microglial Activation, Alters Glial
Fibrillary Acidic Protein Immunoreactivity and Indoleamine
2,3-Dioxygenase during the Development of Offspring Rats.
Molecular Neurobiology, 56, 1096–1108. https ://doi.org/10.1007/
s12035-018-1161-2
Réus, G. Z., Fries, G. R., Stertz, L., Badawy, M., Passos, I. C.,
Barichello, T., … Quevedo, J. (2015a). The role of inflammation
and microglial activation in the pathophysiology of psychiatric dis-
orders. Neuroscience, 300, 141–154. https ://doi.org/10.1016/j.neuro
scien ce.2015.05.018
|
13
CARLESSI ET AL.
Réus, G. Z., Jansen, K., Titus, S., Carvalho, A. F., Gabbay, V., &
Quevedo, J. (2015b). Kynurenine pathway dysfunction in the patho-
physiology and treatment of depression: Evidences from animal and
human studies. Journal of Psychiatric Research, 68, 316–328. https
://doi.org/10.1016/j.jpsyc hires.2015.05.007
Rhee, S. H., Pothoulakis, C., & Mayer, E. A. (2009). Principles and
clinical implications of the brain-gut-enteric microbiota axis. Nature
Reviews Gastroenterology & Hepatology, 6, 306–314. https ://doi.
org/10.1038/nrgas tro.2009.35
Rousseaux, C., Thuru, X., Gelot, A., Barnich, N., Neut, C., Dubuquoy,
L., … Desreumaux, P. (2007). Lactobacillus acidophilus modulates
intestinal pain and induces opioid and cannabinoid receptors. Nature
Medicine, 13, 35–37.
Roy, S. S., & Banerjee, S. (2019). Gut microbiota in neurodegenerative
disorders. Journal of Neuroimmunology, 328, 98–104. https ://doi.
org/10.1016/j.jneur oim.2019.01.004
Ruepert, L., de Quartero, A. O., Wit, N. J. V., der Heijden, G. J., Rubin,
G., & Muris, J. W. (2011). Bulking agents, antispasmodics and anti-
depressants for the treatment of irritable bowel syndrome. Cochrane
Database Systematic Review, 10, 8.
Salter, M., & Pogson, C. I. (1985). The role of tryptophan 2,3-dioxygen-
ase in the hormonal control of tryptophan metabolism in isolated rat
liver cells. Effects of glucocorticoids and experimental diabetes. The
Biochemical Journal, 229, 499–504. https ://doi.org/10.1042/bj229 0499
Sanders, M. E. (2011). Impact of probiotics on colonizing microbiota of
the gut. Journal of Clinical Gastroenterology, 45, S115–S119. https
://doi.org/10.1097/MCG.0b013 e3182 27414a
Sanz, Y., & Moya-Perez, A. (2014). Microbiota, inflammation and obe-
sity. Advances in Experimental Medicine and Biology, 817, 291–
317. https ://doi.org/10.1007/978-1-4939-0897-4_14
Savitz, J., Drevets, W. C., Wurfel, B. E., Ford, B. N., Bellgowan, P. S.,
Victor, T. A., … Dantzer, R. (2015). Reduction of kynurenic acid
to quinolinic acid ratio in both the depressed and remitted phases
of major depressive disorder. Brain, Behavior, and Immunity, 46,
55–59. https ://doi.org/10.1016/j.bbi.2015.02.007
Schildkraut, J. J. (1995). The catecholamine hypothesis of affective dis-
orders: A review of supporting evidence. Journal of Neuropsychiatry
and Clinical Neurosciences, 7, 524–533. https ://doi.org/10.1176/
jnp.7.4.524
Schousboe, A., & Waagepetersen, H. S. (2007). GABA: Homeostatic
and pharmacological aspects. Progress in Brain Research, 160, 9–
19. https ://doi.org/10.1016/s0079-6123(06)60002-2
Schwarcz, R., Bruno, J. P., Muchowski, P. J., & Wu, H. Q. (2012).
Kynurenines in the mammalianbrain: When physiology meets pa-
thology. Nature Reviews Neuroscience, 13, 465–477. https ://doi.
org/10.1038/nrn3257
Scott, F. W., Pound, L. D., Patrick, C., Eberhard, C. E., & Crookshank, J.
A. (2017). Where genes meet environment-integration the role of gut
luminal contents, immunity and pancreas in type 1 diabetes. Transl
Res., 179, 183–198. https ://doi.org/10.1016/j.trsl.2016.09.001.
Smith, L. M., & Parr-Brownlie, L. C. (2019). A neuroscience perspec-
tive of the gut theory of Parkinson's disease. European Journal of
Neuroscience, 49, 817–823. https ://doi.org/10.1111/ejn.13869
Smith, P. A. (2015). The tantalizing links between gut microbes and the
brain. Nature, 526, 312–314. https ://doi.org/10.1038/526312a
Spiller, R., & Garsed, K. (2009). Post-infectious irritable bowel syn-
drome. Gastroenterology, 136, 1979–1988.
Stilling, R. M., Dinan, T. G., & Cryan, J. F. (2014). Microbial genes,
brain & behaviour - epigenetic regulation of the gut-brain axis.
Genes, Brain, and Behavior, 13, 69–86.
Stuart, M. J., & Baune, B. T. (2014). Chemokines and chemokine receptors
in mood disorders, schizophrenia, and cognitive impairment: A sys-
tematic review of biomarker studies. Neuroscience and Biobehavioral
Reviews, 42, 93–115. https ://doi.org/10.1016/j.neubi orev.2014.02.001
Szalardy, L., Zadori, D., Toldi, J., Fulop, F., Klivenyi, P., & Vecsei,
L. (2012). Manipulating kynurenic acid levels in the brain on the
edge between neuroprotectionand cognitive dysfunction. Current
Topics in Medicinal Chemistry, 12, 1797–1806. https ://doi.
org/10.2174/15680 26611 20906 1797
Valkanova, V., Ebmeier, K. P., & Allan, C. L. (2013). CRP, IL-6 and
depression: A systematic review and meta-analysis of longitudinal
studies. Journal of Affective Disorders, 150, 736–744. https ://doi.
org/10.1016/j.jad.2013.06.004
van de Boehme, M., Wouw, M., Bastiaanssen, T. F. S., Olavarría-
Ramírez, L., Lyons, K., Fouhy, F., … Cryan, J. F. (2019). Mid-life
microbiota crises: middle age is associated with pervasive neuroim-
mune alterations that are reversed by targeting the gut microbiome.
Molecular Psychiatry. https ://doi.org/10.1038/s41380-019-0425-1
Van Loosdregt, J., Brunen, D., Fleskens, V., Pals, C. E., Lam, E. W., &
Coffer, P. J. (2011). Rapid temporal control of Foxp3 protein degra-
dation by sirtuin-1. PLoS ONE, 6, e19047. https ://doi.org/10.1371/
journ al.pone.0019047
Waclawiková, B., & Aidy, S. E. (2018). Role of Microbiota and
Tryptophan Metabolites in the Remote Effect of Intestinal
Inflammation on Brain and Depression. Pharmaceuticals (Basel),
11, 63. https ://doi.org/10.3390/ph110 30063
Walker, F. R., Nilsson, M., & Jones, K. (2013). Acute and chronic
stress-induced disturbances of microglial plasticity, phenotype
and function. Current Drug Targets, 14, 1262–1276. https ://doi.
org/10.2174/13894 50111 31499 90208
Westfall, S., & Pasinetti, G. (2019). Design of a novel synbiotic formu-
lation to optimize gut-derived phenolic acid mediated gut-brain axis
signals for the treatment of stress-induced depression and anxiety
(OR23-03-19). Current Developments in Nutrition, 3(Supp 1). https
://doi.org/10.1093/cdn/nzz040.OR23-03-19
Whiteford, H., Degenhardt, L., Rehm, J., Baxter, A. J., Ferrari, J. A.,
Erskine, H. E., … Vos, T. (2013). Global burden of disease attrib-
utable to mental and substance use disorders: Findings from the
Global Burden of Disease Study 2010. Lancet, 382, 1575–1586.
https ://doi.org/10.1016/S0140-6736(13)61611-6
Whitehead, W. E., Palsson, O., & Jones, K. R. (2002). Systematic re-
view of the comorbidity of irritable bowel syndrome with other
disorders: What are the causes and implications? Gastroenterology,
122, 1140–1156. https ://doi.org/10.1053/gast.2002.32392
Wong, M. L., Inserra, A., Lewis, M. D., Mastronardi, C. A., Leong, L.,
Choo, J., … Licinio, J. (2016). Inflammasome signaling affects anx-
iety- and depressive like behavior and gut microbiomecomposition.
Molecular Psychiatry, 21, 797–805.
World Health Organization. (2017). Depression. WHO-MSD-MER-
2017.2-eng.pdf Access in: 09 mai. 2019.
Wurfel, B. E., Drevets, W. C., Bliss, S. A., McMillin, J. R., Suzuki, H.,
Ford, B. N., … Savitz, J. B. (2017). Serum kynurenic acid is reduced
in affective psychosis. Translational Psychiatry, 7, e1115. https ://
doi.org/10.1038/tp.2017.88
Yang, C., Fujita, Y., Ren, Q., Ma, M., Dong, C., & Hashimoto, K.
(2017). Bifidobacterium in the gut microbiota confer resilience to
chronic social defeat stress in mice. Scientific Reports, 7, 45942.
Yang, J. J., Wang, N., Yang, C., Shi, J. Y., Yu, H. Y., & Hashimoto,
K. (2015). Serum interleukin-6 is a predictive biomarker for ket-
amine's antidepressant effect in treatment-resistant patients with
14
|
CARLESSI ET AL.
major depression. Biological Psychiatry, 1, e19–e20. https ://doi.
org/10.1016/j.biops ych.2014.06.021
Yeung, A. W., Terentis, A. C., King, N. J., & Thomas, S. R. (2015).
Role of indoleamine 2,3-dioxygenase in health and disease. Clinical
Science (Lond), 129, 601–672.
Zhang, D., Hu, X., Qian, L., O'Callaghan, J. P., & Hong, J. S. (2010).
Astrogliosis in CNS pathologies: Is there a role for microglia?
Molecular Neurobiology, 41, 232–241.
Zhang, J. C., Yao, W., Dong, C., Yang, C., Ren, Q., Ma, M., &
Hashimoto, K. (2017). Blockade of interleukin-6 receptor in the
periphery promotes rapid and sustained antidepressant actions: A
possible role of gut-microbiota-brain axis. Transl. Psychiatry., 30,
1138. https ://doi.org/10.1038/tp.2017.112
Zhang, K., Fujita, Y., Chang, L., Qu, Y., Pu, Y., Wang, S., Hashimoto,
K. (2019). Abnormal composition of gut microbiota is associated
with resilience versus susceptibility to inescapable electric stress.
Translational Psychiatry, 17, 231.
Zhang, W. Y., Guo, Y. J., Han, W. X., Yang, M. Q., Wen, L. P., Wang,
K. Y., & Jiang, P. (2019). Curcumin relieves depressive-like behav-
iors via inhibition of the NLRP3 inflammasome and kynurenine
pathway in rats suffering from chronic unpredictable mild stress.
International Immunopharmacology, 67, 138–144. https ://doi.
org/10.1016/j.intimp.2018.12.012
Zheng, P., Zeng, B., Zhou, C., Liu, M., Fang, Z., Xu, X., … Xie, P.
(2016). Gut microbiome remodeling induces depressive-like be-
haviors through a pathway mediated by the host's metabolism.
Molecular Psychiatry, 21, 786–796.
Zhu, F., Guo, R., Wang, W., Ju, Y., Wang, Q., Ma, Q., … Ma, X. (2019).
Transplantation of microbiota from drug-free patients with schizo-
phrenia causes schizophrenia-like abnormal behaviors and dysregu-
lated kynurenine metabolism in mice. Molecular Psychiatry, https ://
doi.org/10.1038/s41380-019-0475-4
How to cite this article: Carlessi AS, Borba LA,
Zugno AI, Quevedo J, Réus GZ. Gut microbiota–brain
axis in depression: The role of neuroinflammation.
Eur J Neurosci. 2019;00:1–14. https ://doi.
org/10.1111/ejn.14631
... Natural products and specific nutrients such as vegetables, fruits, medicinal plants, tea, and curcumin were found to reduce oxidative stress and inflammation, enhance the nervous system's performance, and alter biomarkers and signaling pathways related to anxiety and depressive symptoms [44][45][46][47]. These molecular mechanisms can improve monoamine neurotransmitter production and decrease the hyperactivity of the hypothalamus-pituitary-adrenal (HPA) axis [48][49][50]. Moreover, they can modify the microbiota-gut-brain axis, which decreases oxidative stress and disorders connected to inflammation [48][49][50]. ...
... These molecular mechanisms can improve monoamine neurotransmitter production and decrease the hyperactivity of the hypothalamus-pituitary-adrenal (HPA) axis [48][49][50]. Moreover, they can modify the microbiota-gut-brain axis, which decreases oxidative stress and disorders connected to inflammation [48][49][50]. The overload of HPA axis is also associated with depressive symptoms. ...
Article
Full-text available
Background: In recent decades, the incidence of depression has gradually increased in the general population globally. Depression is also common during gestation and could result in detrimental gestational complications for both the mother and the fetus. The survey presented aimed to evaluate whether pregnant women’s perinatal depression could be associated with socio-demographic, anthropometry and lifestyle factors, and perinatal and postnatal outcomes. Methods: This is a cross-sectional survey conducted on 5314 pregnant women. Socio-demographic and lifestyle factors were recorded by relevant questionnaires via face-to-face interviews. Anthropometric parameters were measured by qualified personnel. Perinatal depressive symptomatology status was evaluated by Beck’s Depression Inventory (BDI-II) questionnaire. Results: Depressive symptoms throughout gestation were found in 35.1% of the enrolled women. Perinatal depression was significantly associated with lower educational and economic level, pre-pregnancy regular smoking and reduced levels of Mediterranean diet adherence levels, a higher prevalence of gestational diabetes and preterm birth, as well as a higher incidence of delivering by caesarean section and abnormal childbirth weight. Perinatal depression was also significantly associated with a higher prevalence of maternal postpartum depression and lower prevalence of exclusive breastfeeding practices, as well as with a higher incidence of childhood asthma. Conclusions: Pregnant women’s perinatal depression appears to be associated with various socio-demographic, anthropometry, and lifestyle characteristics and with a higher frequency of several adverse pregnancy complications. The present findings emphasize the importance of pregnant women’s perinatal mental health, highlighting the need to develop and apply public strategies and policies for psychological counseling and support of future mothers to minimize probable risk factors that may trigger perinatal depression. Novel well-organized, follow-up surveys of enhanced validity are highly recommended to establish more definitive conclusions.
... Even if it is unlikely that neurotransmitters produced by microbiota can reach the brain, with the exception of GABA, they can indirectly influence the brain activity through the enteric nervous system (176). Also, tryptophan metabolism pathway can be controlled by products generated by some bacteria, for example Bifidobacterium infantis (177,178). Specifically, SCFAs are not only involved in the BBB integrity, but they can also increase the tryptophan conversion rate into serotonin, indirectly influencing its amount in the brain (179); indeed, SCFAs resulted in being depleted in patients with MDD (180). Moreover, products of tryptophan called indole have been described as neuroactive signaling molecules able to regulate emotional behavior. ...
Article
Full-text available
Major depressive disorder (MDD) is a recurrent episodic mood disorder that represents the third leading cause of disability worldwide. In MDD, several factors can simultaneously contribute to its development, which complicates its diagnosis. According to practical guidelines, antidepressants are the first-line treatment for moderate to severe major depressive episodes. Traditional treatment strategies often follow a one-size-fits-all approach, resulting in suboptimal outcomes for many patients who fail to experience a response or recovery and develop the so-called “therapy-resistant depression”. The high biological and clinical inter-variability within patients and the lack of robust biomarkers hinder the finding of specific therapeutic targets, contributing to the high treatment failure rates. In this frame, precision medicine, a paradigm that tailors medical interventions to individual characteristics, would help allocate the most adequate and effective treatment for each patient while minimizing its side effects. In particular, multi-omic studies may unveil the intricate interplays between genetic predispositions and exposure to environmental factors through the study of epigenomics, transcriptomics, proteomics, metabolomics, gut microbiomics, and immunomics. The integration of the flow of multi-omic information into molecular pathways may produce better outcomes than the current psychopharmacological approach, which targets singular molecular factors mainly related to the monoamine systems, disregarding the complex network of our organism. The concept of system biomedicine involves the integration and analysis of enormous datasets generated with different technologies, creating a “patient fingerprint”, which defines the underlying biological mechanisms of every patient. This review, centered on precision medicine, explores the integration of multi-omic approaches as clinical tools for prediction in MDD at a single-patient level. It investigates how combining the existing technologies used for diagnostic, stratification, prognostic, and treatment-response biomarkers discovery with artificial intelligence can improve the assessment and treatment of MDD.
... These strains, when used either as a single or combination strain, can reduce depression in humans. Other than the regulation of the HPA-axis and sympatho-adrenal medullary (SAM)-axis and the inflammatory reflex, psychobiotics are also involved in cognitive functions of the brain, such as learning, memory, and behaviour [33], which is influenced by changes in the glial cells [34]. Psychobiotics administration achieves MGBA homeostasis, which is threatened by factors such as stress and anxiety; depressive symptoms are well managed when homeostasis is maintained [35]. ...
Article
Full-text available
The field of psychology has advanced over the years in treating psychiatric disorders such as major depressive disorder (MDD), Schizophrenia, and Alzheimer's disease (AD). Depression or clinical depression is a major mental health issue characterized by chronic sadness, hopelessness, and emptiness, which diminishes the patient's quality of life. According to WHO, an estimated 3.8% of the world's population experience depression, in which 15% of depressed patients eventually die by suicide. Recent studies in treating depressive patients have progressed with the usage of psychobiotics. Psychobiotics contain both probiotics and prebiotics, meaning psychobiotics possess the ability to introduce beneficial bacteria in the gut as well as support the growth of existing bacteria in the human gut. The gut-brain axis, which mediates the mechanism of action of psychobiotics in treating clinical depression, has been cleverly studied, and it provides promising results in the improvement of a patient's mental health status. Psychobiotics have proven their worth not only in upgrading the patient's mental health in psychological disorders but also in the enhancement of overall patient health by improving one's gut health.
... 1 Perturbance of the gut microbiota is implicated as a considerable factor in the development of a growing list of neurodegenerative diseases, neurological pathologies, and psychiatric disorders. [10][11][12][13][14][15] Studying the reciprocal impact between the gut microbiota and the health of the central nervous system (CNS) is a growing field in science, and implications are about to derive for new therapeutic approaches. 1 Microbiota-derived metabolites and other products, such as peptides and neuroactive substances, affect CNS development and function 1 and are expected to similarly impact on the peripheral nervous system (PNS). 16 Peripheral nerves originate from the neuroectoderm and elongate with the growth and development of mesodermal somites, which further develop into bone and muscle tissue. ...
Article
Full-text available
Gut microbiota is responsible for essential functions in human health. Several communication axes between gut microbiota and other organs via neural, endocrine, and immune pathways have been described, and perturbation of gut microbiota composition has been implicated in the onset and progression of an emerging number of diseases. Here, we analyzed peripheral nerves, dorsal root ganglia (DRG), and skeletal muscles of neonatal and young adult mice with the following gut microbiota status: a) germ-free (GF), b) gnotobiotic, selectively colonized with 12 specific gut bacterial strains (Oligo-Mouse-Microbiota, OMM12), or c) natural complex gut microbiota (CGM). Stereological and morphometric analyses revealed that the absence of gut microbiota impairs the development of somatic median nerves, resulting in smaller diameter and hypermyelinated axons, as well as in smaller unmyelinated fibers. Accordingly, DRG and sciatic nerve transcriptomic analyses highlighted a panel of differentially expressed developmental and myelination genes. Interestingly, the type III isoform of Neuregulin1 (NRG1), known to be a neuronal signal essential for Schwann cell myelination, was overexpressed in young adult GF mice, with consequent overexpression of the transcription factor Early Growth Response 2 (Egr2), a fundamental gene expressed by Schwann cells at the onset of myelination. Finally, GF status resulted in histologically atrophic skeletal muscles, impaired formation of neuromuscular junctions, and deregulated expression of related genes. In conclusion, we demonstrate for the first time a gut microbiota regulatory impact on proper development of the somatic peripheral nervous system and its functional connection to skeletal muscles, thus suggesting the existence of a novel ‘Gut Microbiota-Peripheral Nervous System-axis.’
... The intricate linkages in the microbiota-gut-brain axis [6][7][8][9][10] are comparable to the bidirectional relationship that exists between the kidney and the gut. An increasing body of evidence indicates the involvement of intestinal dysbiosis in the pathophysiology of ...
Article
Full-text available
(1) Background: Urinary tract infections (UTIs) are among otherwise healthy women represent a problem that requires additional understanding and approaches. Evidencing the link between dysbiosis and UTIs and the associated potential risk factors could lead to therapeutic approaches with increased efficiency under the conditions of reducing the risks associated with antibiotic treatments. The purpose of this study was to evaluate dysbiosis and other potential risk factors in women with a history of urinary tract infections; (2) Methods: Fecal dysbiosis tests were performed comparatively in two groups of women. The first group in-cluded women with recurrent urinary tract infections (rUTI) who had either two or more symp-tomatic episodes of UTI in the previous six months. The second group included women with spo-radic UTIs who did not have >1 UTI during a 12-month period and who did not have another UTI in the last 12 months; (3) Results: An association was shown between intestinal dysbiosis and recurrences of urinary tract infections. Increased body weight was associated with intestinal dysbiosis. Also, the lack of knowledge regarding the risk of using antibiotics and the benefits of probiotics was associated with both dysbiosis and recurrences of urinary tract infections; (4) Conclusions: Dysbiosis can have an impact on the recurrence of urinary tract infections. The risk factors for rUTI and dysbiosis in the sphere of lifestyle are potentially controllable, broadening the perspective for new approaches and changing the paradigm in the treatment of urinary tract infections.
... MFGM supplementation influences the gut microbiome and mood regulation [54][55][56]. The gut microbiota mediate the digestion of food and dietary supplements and influence the production of neurotransmitters [57]. ...
Article
Full-text available
The milk fat globule membrane (MFGM) contains bioactive proteins, carbohydrates, and lipids. Polar lipids found in the MFGM play a critical role in maintaining cell membrane integrity and neuronal signalling capacity, thereby supporting brain health. This review summarises the literature on the MFGM and its phospholipid constituents for improvement of mental health across three key stages of the human lifespan, i.e., infancy, adulthood, and older age. MFGM supplementation may improve mental health by reducing neuroinflammation and supporting neurotransmitter synthesis through the gut–brain axis. Fortification of infant formula with MFGMs is designed to mimic the composition of breastmilk and optimise early gut and central nervous system development. Early behavioural and emotional development sets the stage for future mental health. In adults, promising results suggest that MFGMs can reduce the negative consequences of situational stress. Preclinical models of age-related cognitive decline suggest a role for the MFGM in supporting brain health in older age and reducing depressive symptoms. While there is preclinical and clinical evidence to support the use of MFGM supplementation for improved mental health, human studies with mental health as the primary target outcome are sparce. Further high-quality clinical trials examining the potential of the MFGM for psychological health improvement are important.
... Importantly, both obesity and MetS have been found to be independently associated with depressive symptoms and inflammation. A possible pathophysiological overlap is being considered, with chronic low-grade inflammation and dysbiosis being suggested as possible connecting factors [5]. ...
Article
Full-text available
Probiotics may represent a safe and easy-to-use treatment option for depression or its metabolic comorbidities. However, it is not known whether metabolic features can influence the efficacy of probiotics treatments for depression. This trial involved a parallel-group, prospective, randomized, double-blind, controlled design. In total, 116 participants with depression received a probiotic preparation containing Lactobacillus helveticus Rosell®-52 and Bifidobacterium longum Rosell®-175 or placebo over 60 days. The psychometric data were assessed longitudinally at five time-points. Data for blood pressure, body weight, waist circumference, complete blood count, serum levels of C-reactive protein, cholesterol, triglycerides, and fasting glucose were measured at the beginning of the intervention period. There was no advantage of probiotics usage over placebo in the depression score overall (PRO vs. PLC: F(1.92) = 0.58; p = 0.45). However, we found a higher rate of minimum clinically important differences in patients supplemented with probiotics than those allocated to placebo generally (74.5 vs. 53.5%; X²(1,n = 94) = 4.53; p = 0.03; NNT = 4.03), as well as in the antidepressant-treated subgroup. Moreover, we found that the more advanced the pre-intervention metabolic abnormalities (such as overweight, excessive central adipose tissue, and liver steatosis), the lower the improvements in psychometric scores. A higher baseline stress level was correlated with better improvements. The current probiotic formulations may only be used as complementary treatments for depressive disorders. Metabolic abnormalities may require more complex treatments. ClinicalTrials.gov identifier: NCT04756544.
Article
The hippocampus is one of the most commonly studied brain regions in the context of depression. The volume of the hippocampus is significantly reduced in patients with depression, which severely disrupts hippocampal neuroplasticity. However, antidepressant therapies that target hippocampal neuroplasticity have not been identified as yet. Chinese medicine (CM) can slow the progression of depression, potentially by modulating hippocampal neuroplasticity. Xiaoyaosan (XYS) is a CM formula that has been clinically used for the treatment of depression. It is known to protect Gan (Liver) and Pi (Spleen) function, and may exert its antidepressant effects by regulating hippocampal neuroplasticity. In this review, we have summarized the association between depression and aberrant hippocampal neuroplasticity. Furthermore, we have discussed the researches published in the last 30 years on the effects of XYS on hippocampal neuroplasticity in order to elucidate the possible mechanisms underlying its therapeutic action against depression. The results of this review can aid future research on XYS for the treatment of depression.
Article
Full-text available
Depression is a highly prevalent psychological disorder characterized by persistent dysphoria, psychomotor retardation, insomnia, anhedonia, suicidal ideation, and a remarkable decrease in overall well-being. Despite the prevalence of accessible antidepressant therapies, many individuals do not achieve substantial improvement. Understanding the multifactorial pathophysiology and the heterogeneous nature of the disorder could lead the way toward better outcomes. Recent findings have elucidated the substantial impact of compromised blood-brain barrier (BBB) integrity on the manifestation of depression. BBB functions as an indispensable defense mechanism, tightly overseeing the transport of molecules from the periphery to preserve the integrity of the brain parenchyma. The dysfunction of the BBB has been implicated in a multitude of neurological disorders, and its disruption and consequent brain alterations could potentially serve as important factors in the pathogenesis and progression of depression. In this review, we extensively examine the pathophysiological relevance of the BBB and delve into the specific modifications of its components that underlie the complexities of depression. A particular focus has been placed on examining the effects of peripheral inflammation on the BBB in depression and elucidating the intricate interactions between the gut, BBB, and brain. Furthermore, this review encompasses significant updates on the assessment of BBB integrity and permeability, providing a comprehensive overview of the topic. Finally, we outline the therapeutic relevance and strategies based on BBB in depression, including COVID-19-associated BBB disruption and neuropsychiatric implications. Understanding the comprehensive pathogenic cascade of depression is crucial for shaping the trajectory of future research endeavors.
Article
Full-text available
Increasing evidence indicates that abnormalities in the composition of gut microbiota might play a role in stress-related disorders. In the learned helplessness (LH) paradigm, ~60–70% rats are susceptible to LH in the face of inescapable electric stress. The role of gut microbiota in susceptibility in the LH paradigm is unknown. In this study, male rats were exposed to inescapable electric stress under the LH paradigm. The compositions of gut microbiota and short-chain fatty acids were assessed in fecal samples from control rats, non-LH (resilient) rats, and LH (susceptible) rats. Members of the order Lactobacillales were present at significantly higher levels in the susceptible rats than in control and resilient rats. At the family level, the number of Lactobacillaceae in the susceptible rats was significantly higher than in control and resilient rats. At the genus level, the numbers of Lactobacillus, Clostridium cluster III, and Anaerofustis in susceptible rats were significantly higher than in control and resilient rats. Levels of acetic acid and propionic acid in the feces of susceptible rats were lower than in those of control and resilient rats; however, the levels of lactic acid in the susceptible rats were higher than those of control and resilient rats. There was a positive correlation between lactic acid and Lactobacillus levels among these three groups. These findings suggest that abnormal composition of the gut microbiota, including organisms such as Lactobacillus, contributes to susceptibility versus resilience to LH in rats subjected to inescapable electric foot shock. Therefore, it appears likely that brain–gut axis plays a role in stress susceptibility in the LH paradigm.
Article
Full-text available
Accumulating evidence suggests that gut microbiota plays a role in the pathogenesis of schizophrenia via the microbiota-gut-brain axis. This study sought to investigate whether transplantation of fecal microbiota from drug-free patients with schizophrenia into specific pathogen-free mice could cause schizophrenia-like behavioral abnormalities. The results revealed that transplantation of fecal microbiota from schizophrenic patients into antibiotic-treated mice caused behavioral abnormalities such as psychomotor hyperactivity, impaired learning and memory in the recipient animals. These mice also showed elevation of the kynurenine-kynurenic acid pathway of tryptophan degradation in both periphery and brain, as well as increased basal extracellular dopamine in prefrontal cortex and 5-hydroxytryptamine in hippocampus, compared with their counterparts receiving feces from healthy controls. Furthermore, colonic luminal filtrates from the mice transplanted with patients' fecal microbiota increased both kynurenic acid synthesis and kynurenine aminotransferase II activity in cultured hepatocytes and forebrain cortical slices. Sixty species of donor-derived bacteria showed significant difference between the mice colonized with the patients' and the controls' fecal microbiota, highlighting 78 differentially enriched functional modules including tryptophan biosynthesis function. In conclusion, our study suggests that the abnormalities in the composition of gut microbiota contribute to the pathogenesis of schizophrenia partially through the manipulation of tryptophan-kynurenine metabolism.
Article
Full-text available
Stress disturbs the balance of the gut microbiota and stimulates inflammation-to-brain mechanisms. Moreover, stress leads to anxiety and depressive disorders. Bifidobacterium adolescentis displays distinct anti-inflammatory effects. However, no report has focused on the anxiolytic and antidepressant effects of B. adolescentis related to the gut microbiome and the inflammation on chronic restraint stress (CRS) in mice. We found that pretreatment with B. adolescentis increased the time spent in the center of the open field apparatus, increased the percentage of entries into the open arms of the elevated plus-maze (EPM) and the percentage of time spent in the open arms of the EPM, and decreased the immobility duration in the tail suspension test as well as the forced swimming test (FST). Moreover, B. adolescentis increased the sequence proportion of Lactobacillus and reduced the sequence proportion of Bacteroides in feces. Furthermore, B. adolescentis markedly reduced the protein expression of interleukin-1β (IL-1β), tumor necrosis factor α (TNF-α), p-nuclear factor-kappa B (NF-κB) p65 and Iba1 and elevated brain derived neurotrophic factor (BDNF) expression in the hippocampus. We conclude that the anxiolytic and antidepressant effects of B. adolescentis are related to reducing inflammatory cytokines and rebalancing the gut microbiota.
Article
Full-text available
Objectives: Synbiotics, the combination of probiotics and prebiotics, may optimize the production of polyphenolic metabolites, and act as therapeutic agents for inflammation-induced depression. Recent evidence suggests that dysregulated immune activity increases susceptibility to depression and that bioactive polyphenolic metabolites can effectively reduce that inflammation. The problem remains that bioactive metabolite production is dependent on the gut microbiota, leading to significant interpersonal variation in the metabolites' therapeutic efficacy. The hypothesis of the study is that the synbiotic will standardize production and bioavailability of bioactive metabolites capable of suppressing innate immune biological signatures of depression. Methods: To standardize the production of bioactive metabolites, the synbiotic will be designed in an innovative in vitro model of the human gastrointestinal tract using a multivariate regression algorithm to predict which probiotic formulation produces the most effective bioactive metabolites. Following in vivo bioavailability and toxicity testing, the synbiotic's therapeutic efficacy was tested in a chronic unpredictable stress (CUS) mouse model of depression by measuring specific behaviors and changes to the gut microbiota populations. These changes were correlated to biological markers of depression modulated by the synbiotic-derived metabolites including neurobiological markers of depression and variations in innate immune markers, including interleukin-1β (IL-1β). Results: In this study, we show that a synbiotic combining a dietary polyphenolic preparation with L. plantarum and B. longum can potentiate the reduction in anxiety and depression in male mice subjected to a 28 day CUS protocol, as compared to polyphenolic treatment alone. Interestingly, we found that the synbiotic may mediate microglia inflammasome activation.This finding was reflected by inhibition of NLRP3-mediated generation of IL-1β in microglia. Conclusions: Collectively, these results support the potential role of a synbiotic in the potentiation of attenuation of psychological impairment in a model of depression through mechanisms that involved innate immune NLRP3 inflammation mediation in microglia. Funding sources: This project was funding a P50 CARBON Center grant from the NCCIH/ODS (Pasinetti, PD/PI).
Article
Full-text available
Male middle age is a transitional period where many physiological and psychological changes occur leading to cognitive and behavioural alterations, and a deterioration of brain function. However, the mechanisms underpinning such changes are unclear. The gut microbiome has been implicated as a key mediator in the communication between the gut and the brain, and in the regulation of brain homeostasis, including brain immune cell function. Thus, we tested whether targeting the gut microbiome by prebiotic supplementation may alter microglia activation and brain function in ageing. Male young adult (8 weeks) and middle-aged (10 months) C57BL/6 mice received diet enriched with a prebiotic (10% oligofructose-enriched inulin) or control chow for 14 weeks. Prebiotic supplementation differentially altered the gut microbiota profile in young and middle-aged mice with changes correlating with faecal metabolites. Functionally, this translated into a reversal of stress-induced immune priming in middle-aged mice. In addition, a reduction in ageing-induced infiltration of Ly-6Chi monocytes into the brain coupled with a reversal in ageing-related increases in a subset of activated microglia (Ly-6C+) was observed. Taken together, these data highlight a potential pathway by which targeting the gut microbiome with prebiotics can modulate the peripheral immune response and alter neuroinflammation in middle age. Our data highlight a novel strategy for the amelioration of age-related neuroinflammatory pathologies and brain function.
Article
Depression is a disorder of mood that causes strong impact on the patient and his family's quality of life. The increasing number of cases and its social consequences have made depression a great public health problem. Some depressant patients develop suicide thoughts and may try suicide later. The disease occurs in all ages and its prevalence is of 7.4%. Women aged 15 to 29 are more likely to be affected, whereas people aged 50 or older are less affected. The presence of depression worsens the prognostic of other clinical conditions when compared to not depressant patients. This paper describes the main drugs used in the treatment of depression and correlates the drug with the neurobiology of the disease. Aiming the study of the pharmacologic and therapeutic characteristics of antidepressant drugs, a literature review was performed using electronic databases (Pubmed and Lilacs), papers and books related to the theme. The main antidepressant drugs are classified according to their chemical structure or their action on neurotransmitters. An important point in the therapy is the understanding of the pharmacokinetics of the drugs. The choice of the drug must consider the symptoms, the patient's age, other drugs in use, the history of pharmacological treatments and so forth. No drug is significantly better than the other in the treatment of depression. The important thing in choosing a drug is to have the best therapeutic response, the reduction of symptoms, a good adherence to therapy, few side effects and secure drug interaction. By understanding the use of these drugs, it will be possible to give the patient and his family a better quality of life.
Article
Nutritional interventions targeting the microbiota‐gut‐brain axis are proposed to modulate stress‐induced dysfunction of physiological processes and brain development. Maternal separation (MS) in rats induces long‐term alterations to behavior, pain responses, gut microbiome and brain neurochemistry. In this study, the effects of dietary interventions (milk fat globule membrane (MFGM) and a polydextrose/galactooligosaccharide prebiotic blend (PDX/GOS)) were evaluated. Diets were provided from postnatal day 21 to both non‐separated (NS) and MS offspring. Spatial memory, visceral sensitivity and stress reactivity were assessed in adulthood. Gene transcripts associated with cognition and stress and the caecal microbiota composition were analysed. MS‐induced visceral hypersensitivity was ameliorated by MFGM and to greater extent with the combination of MFGM and prebiotic blend. Furthermore, spatial learning and memory were improved by prebiotics and MFGM alone and with the combination. The prebiotic blend and the combination of the prebiotics and MFGM appeared to facilitate return to baseline with regard to HPA axis response to the restraint stress which can be beneficial in times where coping mechanisms to stressful events are required. Interestingly, the combination of MFGM and prebiotic reduced the long‐term impact of MS on a marker of myelination in the prefrontal cortex. MS affected the microbiota at family level only while MFGM, the prebiotic blend and the combination influenced abundance at family and genus level as well as influencing beta diversity levels. In conclusion, intervention with MFGM and prebiotic blend significantly impacted the composition of the microbiota as well as ameliorating some of the long‐term effects of early‐life stress. This article is protected by copyright. All rights reserved.
Article
Background: Increased peripheral inflammation has been consistently reported in patients with major depressive disorder (MDD). However, only few studies have explored markers of central (brain) inflammation in patients with MDD. The aim of this study is to systematically review in vivo and post-mortem markers of central inflammation, including studies examining cerebrospinal fluid (CSF), positron emission tomography, and post-mortem brain tissues in subjects suffering with MDD compared with controls. Methods: PubMed and Medline databases were searched up to December 2018. We included studies measuring cerebrospinal fluid (CSF) cytokines and chemokines, positron emission tomography (PET) studies; and post-mortem studies measuring cytokines, chemokines and cell-specific markers of microglia and astrocytes, all in MDD. A meta-analysis was performed only for CSF and PET studies, as studies on post-mortem markers of inflammation had different cell-specific markers and analysed different brain regions. Results: A total of 69 studies met the inclusion criteria. CSF levels of IL-6 and TNF-α were higher in patients with MDD compared with controls (standardised mean difference SMD 0.37, 95%CI: 0.17-0.57 and SMD 0.58, 95%CI 0.26-0.90, respectively). CSF levels of IL-6 were increased in suicide attempters regardless of their psychiatric diagnosis. Translocator protein, a PET marker of central inflammation, was elevated in the anterior cingulate cortex and temporal cortex of patients with MDD compared with controls (SMD 0.78, 95%CI: 0.41-1.16 and SMD 0.52, 95%CI: 0.19-0.85 respectively). Abnormalities in CSF and PET inflammatory markers were not correlated with those in peripheral blood. In post-mortem studies, two studies found increased markers of microglia in MDD brains, while four studies found no MDD related changes. Of the studies investigating expression of cell-specific marker for astrocytes, thirteen studies reported a decreased expression of astrocytes specific markers, two studies reported increased expression of astrocytes specific markers, and eleven studies did not detect any difference. Four out of six studies reported decreased markers of oligodendrocytes in the prefrontal cortex. Post-mortem brain levels of tumor necrosis alpha (TNF-α) were also found increased in MDD. Conclusions: Our review suggests the presence of an increase in IL-6 and TNF-alpha levels in CSF and brain parenchyma, in the context of a possible increased microglia activity and reduction of astrocytes and oligodendrocytes markers in MDD. The reduced number of astrocytes may lead to compromised integrity of blood brain barrier with increased monocyte recruitment and infiltration, which is partly supported by post-mortem studies and by PET studies showing an increased TSPO expression in MDD.
Article
A relationship between neuroinflammation and the development of psychiatric disorder have been revealed by many studies in the past decade. Although studies have shown that stressors can induce long-term changes that may be related to behavioral responses, these alterations have been poorly studied soon after a stressor, such as maternal deprivation (MD). Thus, this study was designed to investigate the acute effect of experimental induction of MD on inflammatory and microglial activation markers in the brain of infant rats. Early MD was induced from postnatal day (PND) 1–10. On PND 10 the prefrontal cortex (PFC) and hippocampus from MD and control groups were removed to investigate microglial activation and neuroinflammatory markers. In the PFC the expressions of cluster of differentiation molecule 11B (CD11B), toll-like receptor (TLR)-2, and TLR-4 were increased in rats subjected to MD. The arginase expression was elevated in the PFC and hippocampus of maternally deprived rats. The cytokines interleukin-5 (IL-5), −6, −7, −10, tumor necrosis factor (TNF-α), and interferon gamma (INF-γ) were increased in the PFC of MD rats group. In the PFC the macrophage inflammatory proteins (MIP)-1α levels were reduced in MD rats group. In the hippocampus only the levels of TNF-α and INF-γ were elevated in infant rats subjected to MD. In conclusion, our results support the hypothesis that neuroinflammation and microglial activation, mainly in the PFC, could be involved with changes in the brain resident cells following MD, and these alterations could be associated to the development of psychiatric conditions late in life.
Article
Background: Stress during early childhood, for example as a result of maltreatment, can predict inflammation in adulthood. The association of depression with inflammation and current and long-term stress resulting from childhood maltreatment and threatening experiences in the past year has not yet been studied. Therefore, we assessed these variables in a group of patients with major depressive disorder (MDD) and measured levels of the pro-inflammatory cytokine IL-6 and the anti-inflammatory cytokine IL-10. High levels of IL-6 are associated with depression and of IL-10 with stress. Methods: We included 44 patients who fulfilled DSM-IV diagnostic criteria for MDD and 44 age- and gender-matched healthy controls. We used Cohen's Perceived Stress Scale (PSS), the list of life-threatening experiences questionnaire (LTE-Q) and the childhood trauma questionnaire (CTQ) to assess the level of stress and analyzed IL-6 and IL-10 cytokines in venous blood plasma. Results: The patient group showed significantly higher scores on the maltreatment scale LTE-Q (2.7 vs. 1.1; P = 0.001, and the stress scales CTQ (emotional abuse; P = 0.048 and physical neglect; P = 0.002) and PSS (35.2 vs 15.5; P < 0.001) as well as significantly higher levels of IL-6 (1.5pg/ml vs. 0.9pg/ml; P = 0.012). They also had significantly higher levels of IL-10 (1.1pg/ml vs. 0.7pg/ml; P < 0.001). Higher actual stress levels were associated with childhood maltreatment and higher IL-6 (tau = 0.004) and IL-10 (tau = 0.027) levels. Limitations: The results need to be replicated in a larger sample, and the study did not evaluate causal relationships. Although the assessment of childhood trauma was retrospective, the CTQ is a well-established assessment instrument. Conclusions: The patients with MDD in this study showed an immune activation in response to stress. This study highlights the association of childhood trauma and current and long-term stress with an increased immune activation in MDD.